Molecular Mechanisms Involved in Insulin- and Leptin - TSpace

200
Molecular Mechanisms Involved in Insulin- and Leptin- mediated Regulation of Hypothalamic Proglucagon Gene Expression and Action of Glucagon-like Peptides on Hypothalamic Neuropeptides by Prasad S. Dalvi A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Physiology University of Toronto © Copyright by Prasad S. Dalvi 2012

Transcript of Molecular Mechanisms Involved in Insulin- and Leptin - TSpace

Molecular Mechanisms Involved in Insulin- and Leptin-

mediated Regulation of Hypothalamic Proglucagon Gene

Expression and Action of Glucagon-like Peptides on

Hypothalamic Neuropeptides

by

Prasad S. Dalvi

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Graduate Department of Physiology

University of Toronto

© Copyright by Prasad S. Dalvi 2012

ii

Molecular Mechanisms Involved in Insulin- and Leptin-mediated

Regulation of Hypothalamic Proglucagon Gene Expression and Action

of Glucagon-like Peptides on Hypothalamic Neuropeptides

Prasad S. Dalvi

Doctor of Philosophy

Department of Physiology

University of Toronto

2012

Abstract

The hypothalamus is a central regulator of energy homeostasis. Recently, proglucagon-

derived peptides have emerged as potential appetite regulators. The proglucagon gene is

expressed in the periphery and also in selective hypothalamic neurons. The regulation of

hypothalamic proglucagon by two key regulators of energy balance, insulin and leptin, remains

unstudied. Central glucagon-like peptide (GLP)-1 receptor (GLP-1R) activation by exendin-4, a

long-acting GLP-1R agonist, induces anorexia; however, the specific hypothalamic neuronal

populations activated by exendin-4 remain largely unknown. The role of GLP-2 as a central

appetite regulator is poorly understood. In this thesis, using murine hypothalamic cell lines and

mice as experimental models, mechanisms involved in the direct regulation of proglucagon gene

by insulin and leptin were studied, and the actions of exendin-4 and GLP-2 on hypothalamic

neuropeptides were determined.

It was found that insulin and leptin regulate hypothalamic proglucagon mRNA by

activating Akt and signal transducer and activator of transcription 3, respectively. Insulin and

leptin did not regulate human proglucagon promoter regions, but affected proglucagon mRNA

stability. In mice, intracerebroventricular exendin-4 and GLP-2 induced anorexia, activated

proopiomelanocortin- and neuropeptide Y-expressing neurons in the arcuate nucleus and

iii

neurotensin- and ghrelin-expressing neurons in major hypothalamic appetite-regulating regions.

In the hypothalamic neuronal models, exendin-4 and GLP-2 activated cAMP-response element-

binding protein/activating transcription factor-1, and regulated neurotensin and ghrelin mRNA

levels via a protein kinase A-dependent mechanism. Overall, the in vivo and in vitro findings

suggest that these neuropeptides may serve as potential downstream mediators of exendin-4 and

GLP-2 action.

This research demonstrates direct regulation of hypothalamic proglucagon by insulin and

leptin in vitro, and reports a previously unrecognized link between central GLP-1R and GLP-2R

activation and regulation of hypothalamic neuropeptides. A better understanding of the

regulation of hypothalamic proglucagon and central GLP-1R and GLP-2R activation is important

to further expand our knowledge of feeding circuits.

iv

This thesis is dedicated

to

my mother, Jayanti, and father, Shridhar,

my sisters, Shailaja and Ranjana, and brothers-in-law, Bapu and Anant

my mother-in-law, Leena, and late father-in-law, Ramchandra

my daughter, Sanika, and son, Ojus,

& my wife, Pooja

for their endless love, support and encouragement

v

Preface

This thesis is submitted for the degree of Doctor of Philosophy at the University of

Toronto. The research described herein was conducted between January 2007 and April 2012

under the supervision of Dr. Denise D. Belsham in the Department of Physiology, University of

Toronto, ON, Canada. Financial stipend was provided by Banting and Best Diabetes Centre,

Ontario Government, Natural Sciences and Engineering Research Council of Canada and Dr.

Denise D. Belsham.

This research is, to the best of my knowledge, original, except where acknowledgements

and references are made to previous work. This research has been presented in the following

publications:

Prasad S. Dalvi, Frederick D. Erbiceanu, David M. Irwin, Denise D. Belsham. Direct

Regulation of the Proglucagon Gene by Insulin, Leptin and cAMP in Embryonic versus Adult

Hypothalamic Neurons. Molecular Endocrinology (2012) 26:1339-1355

Prasad S. Dalvi, Anaies Nazarians-Armavil, Matthew J. Purser and Denise D. Belsham.

Glucagon-like peptide-1 receptor agonist, Exendin-4, regulates feeding-associated neuropeptides

in hypothalamic neurons in vivo and in vitro. Endocrinology (2012) 153:2208-2222

Prasad S. Dalvi and Denise D. Belsham. Glucagon-like peptide-2 regulates

hypothalamic neurons expressing neuropeptides linked to appetite control in vivo and in vitro.

Endocrinology (2012) 153:2385-2397

.

vi

Acknowledgments

First and foremost, my heart-felt gratitude goes to my supervisor, Dr. Denise D. Belsham,

who hired me as a Master’s student at a very critical period in my life and has ever since guided

me through my doctorate degree. I am grateful to her for her time, constant support and

guidance. Her enthusiasm and encouragement were a strong motivational force throughout the

course of my graduate studies. The past five and a half years in the lab have been thoroughly

enjoyable and an amazing learning experience that has been profoundly rewarding and has truly

played an important part in shaping me as a cellular biology researcher.

I also sincerely thank Dr. Patricia Brubaker, Dr. Tianru Jin and Dr. Allen Volchuk for

being a part of my supervisory committee and for their invaluable insight towards the project. I

am also grateful to Dr. Martin Ralph, Dr. David Irwin and their teams for the great collaborative

efforts. I also greatly appreciate the support from the Department of Physiology, Banting and

Best Diabetes Centre, Ontario Government, Natural Sciences and Engineering Research Council

of Canada and other funding agencies for supporting my research.

My deepest appreciation goes towards the past and present members of the Belsham Lab

for their invaluable technical help at the bench and helping me orient to the Canadian way of life.

I am forever grateful to Jennifer Chalmers, Margaret Koletar and Luisa Centeno for teaching me

the ABCs of many scientific methods. It has been a great pleasure working with all.

Last but definitely not least, this work would not have been possible without my

incredibly supportive family, in-laws, teachers, mentors, friends and well-wishers spread across

the globe. Their endless love, teachings, support and encouragement have meant the world to me

and I thank them from the bottom of my heart for their endless love, support and encouragement!

vii

Table of Contents

List of Figures........................................................................................................................ xi

List of Abbreviations............................................................................................................. xiv

Chapter 1: Background and Significance........................................................................... 1

1.1 Introduction........................................................................................................... 3

1.2 Regulation of Energy Homeostasis........................................................................ 4

1.2.1 Hypothalamus......................................................................................... 5

1.2.1.1 Arcuate nucleus............................................................................... 5

1.2.1.2 Paraventricular nucleus................................................................... 7

1.2.1.3 Ventromedial hypothalamic nucleus............................................... 9

1.2.1.4 Dorsomedial hypothalamic nucleus................................................ 9

1.2.1.5 Lateral hypothalamus..................................................................... 9

1.2.2 The Hypothalamic Neuropeptides.......................................................... 10

1.2.2.1 Neuropeptide Y……….....................................................................

1.2.2.2 Pro-opiomelanocortin......................................................................

1.2.2.3 Neurotensin.....................................................................................

10

11

12

1.2.2.4 Ghrelin............................................................................................. 13

1.3 Proglucagon-derived peptides (PGDPs)............................................................... 15

1.3.1 Biosynthesis............................................................................................ 15

1.3.2 Role of PGDPs as appetite regulators.................................................... 17

1.3.3 Expression of PGDPs and their receptors in the hypothalamus............. 18

1.3.4 Signaling pathways activated by GLP-1R and GLP-2R stimulation...... 20

1.4 Insulin and insulin receptor signaling...................................................................

1.4.1 Insulin......................................................................................................

1.4.2 Insulin receptor signaling........................................................................

1.4.3 Insulin regulation of proglucagon...........................................................

1.5 Leptin and leptin receptor signaling......................................................................

1.5.1 Leptin......................................................................................................

1.5.2 Leptin receptor signaling.........................................................................

23

23

23

24

25

25

26

viii

1.5.3 Leptin regulation of proglucagon............................................................

1.6 Hypothalamic Neuronal Cell Models....................................................................

27

28

1.6.1 Generation of immortalized hypothalamic cell lines..............................

1.6.2 Immortalized embryonic hypothalamic cell lines, mHypoE-XX...........

28

29

1.6.3 Immortalized adult hypothalamic cell lines, mHypoA-XX.................... 30

1.7 Rationale, Hypotheses and Specific Aims..............................................................

1.7.1 Rationale.................................................................................................

1.7.2 Hypotheses..............................................................................................

1.7.3 Specific aims...........................................................................................

30

30

32

32

Chapter 2: Materials and Methods......................................................................................

2.1 Cell culture and reagents.......................................................................................

2.2 One-step reverse transcriptase-polymerase chain reaction..................................

2.3 Radioimmunoassay for adenosine 3’,5’-cyclic monophosphate...........................

2.4 Quantitative reverse transcription-polymerase chain reaction.............................

2.5 SDS-polyacrylamide gel electrophoresis and western blot analysis....................

2.6 Reporter gene plasmids.........................................................................................

2.7 Transient transfections..........................................................................................

2.8 Animal (mouse) experiments..................................................................................

2.9 Intracerebroventricular (i.c.v.) microinjections for feeding study........................

2.10 Assessment of neuronal activation by c-Fos immunohistochemistry………..........

2.11 Experimental normalization..................................................................................

2.12 Statistical analysis.................................................................................................

35

36

36

37

38

39

40

41

41

42

43

44

45

Chapter 3: Regulation of the Proglucagon mRNA levels by Insulin and Leptin in

Embryonic versus Adult Hypothalamic Neurons...............................................................

3.1 Abstract..................................................................................................................

46

48

3.2 Introduction........................................................................................................... 48

ix

3.3 Results....................................................................................................................

3.3.1 Characterization of the expression profile of the hypothalamic cell

lines.................................................................................................................

3.3.2 Activation of signaling pathways by insulin and leptin in the

hypothalamic neuronal cells............................................................................

3.3.3 Regulation of proglucagon mRNA transcript expression by insulin

and leptin..........................................................................................................

3.3.4 Reversal of insulin-mediated regulation of proglucagon mRNA

expression by PI3K inhibitors..........................................................................

3.3.5 Reversal of leptin-mediated regulation of proglucagon mRNA

expression by JAK2/STAT3 inhibitors............................................................

3.3.6 Regulation of human or rat proglucagon 5’ flanking promoter

constructs by insulin and leptin.......................................................................

3.3.7 Regulation of mRNA stability by insulin and leptin in mHypoA-2/10

and mHypoE-39 neuronal cells........................................................................

3.3.8 In silico analysis of murine proglucagon mRNA sequence for

microRNA binding sites and RNA-binding protein sites................................

3.4 Discussion..............................................................................................................

50

50

52

54

54

58

61

62

66

67

Chapter 4: Glucagon-like Peptide-1 Receptor Agonist, Exendin-4, Regulates Feeding-

associated Neuropeptides in Hypothalamic Neurons in vivo and in vitro.........................

4.1 Abstract..................................................................................................................

4.2 Introduction...........................................................................................................

4.3 Results....................................................................................................................

4.3.1 Effect of i.c.v. exendin-4 on food and water intake, and animal weight

4.3.2 Effect of i.c.v. exendin-4 on activation of hypothalamic nuclei and

neuropeptidergic neurons.................................................................................

4.3.3 Expression of GLP-1R and appetite-regulating neuropeptides in adult

mHypoA-2/30 and embryonic mHypoE-36/1 neurons....................................

4.3.4 Activation of CREB/ATF-1 and c-Fos by exendin-4 in the

hypothalamic neuronal cells.............................................................................

74

76

77

78

78

80

85

88

x

4.3.5 Regulation of neurotensin and ghrelin mRNA transcript expression by

exendin-4.........................................................................................................

4.3.6 Reversal of exendin-4 regulation of neurotensin and ghrelin by PKA

inhibitors..........................................................................................................

4.4 Discussion..............................................................................................................

Chapter 5: Glucagon-like Peptide-2 Regulates Hypothalamic Neurons Expressing

Neuropeptides Linked to Appetite Control in vivo and in vitro.........................................

5.1 Abstract..................................................................................................................

5.2 Introduction...........................................................................................................

5.3 Results....................................................................................................................

5.3.1 Effect of i.c.v. h(Gly2)GLP-2 on food and water intake, and animal

weight..............................................................................................................

5.3.2 Effect of h(Gly2)GLP-2 on activation of hypothalamic nuclei and

neuropeptidergic neurons.................................................................................

5.3.3 Expression of GLP-2R and appetite-regulating neuropeptides in

adult mHypoA-2/30 cells.................................................................................

5.3.4 Activation of CREB/ATF-1 and c-Fos by h(Gly2)GLP-2 in the

hypothalamic neuronal cells.............................................................................

5.3.5 Regulation of neurotensin and ghrelin mRNA transcript expression by

h(Gly2)GLP-2 ..................................................................................................

5.3.6 PKA inhibitors reverse the h(Gly2)GLP-2-induced up-regulation of

neurotensin and ghrelin mRNA transcript levels………..................................

5.4 Discussion..............................................................................................................

Chapter 6: Discussion............................................................................................................

6.1 Overall conclusions...............................................................................................

6.2 Limitations.............................................................................................................

6.3 Future directions....................................................................................................

Chapter 7: References...........................................................................................................

91

93

95

102

104

104

106

106

108

113

115

115

118

119

125

126

136

142

146

xi

List of Figures

Figure 1. Schematic diagrams of the hypothalamic and brainstem regions that express

neuropeptides involved in energy homeostasis (coronal sections) ........................

8

Figure 2. The products of preproglucagon cleavage. Schematic diagrams of the central

regions showing gene expression of proglucagon and GLP-1R and GLP-2R.....

19

Figure 3. Summary of GLP-1R signaling............................................................................... 21

Figure 4. Summary of GLP-2R signaling............................................................................... 22

Figure 5. Schematic of the proposed regulation of proglucagon-expressing neurons by

insulin and leptin in the hypothalamus...................................................................

33

Figure 6. Summary of the proposed signal transduction mechanisms triggered by GLP-1R

and GLP-2R activation in the hypothalamus.........................................................

34

Figure 7. Characterization of the expression profile of the proglucagon-expressing

hypothalamic cell lines............................................................................................

51

Figure 8. Insulin activates Akt and leptin activates STAT3 in the hypothalamic neuronal

cells.........................................................................................................................

53

Figure 9. Insulin and leptin regulate proglucagon mRNA expression in the hypothalamic

neuronal cells...........................................................................................................

55

Figure 10. Regulation of proglucagon mRNA expression by insulin via activation of the

PI3K/Akt pathway..................................................................................................

56

Figure 11. Regulation of proglucagon mRNA expression by leptin via activation of the

JAK2/STAT3 pathway...........................................................................................

Figure 12. Insulin and leptin do not affect the transcription of proglucagon promoter

constructs, but regulate mRNA stability.................................................................

Figure 13. In silico analysis of murine proglucagon mRNA sequence for miRNA binding

sites and RNA-binding protein sites.......................................................................

Figure 14. Intracerebroventricular injection of exendin-4 was effective to reduce food and

water intake, and body weight in mice....................................................................

Figure 15. Exendin-4 activates hypothalamic neurons...........................................................

Figure 16. Acute exendin-4 treatment increases the number of hypothalamic neurons co-

expressing c-Fos-immunoreactivity (ir) with α-MSH-, NPY-, neurotensin- or

ghrelin-ir..................................................................................................................

59

63

68

79

81

83

xii

Figure 17. Graphical representation showing the number of neurons expressing c-Fos and

neuropeptide-immunoreactivity in the ARC, VMH, DMH, PVN, LH, PeV,

and the internuclear space between the VMH and DMH of the saline- or

exendin-4-treated mouse hypothalamus..................................................................

Figure 18. Expression profile of GLP-1 receptor and appetite-regulating neuropeptides

in the hypothalamic neuronal cell lines..................................................................

Figure 19. Exendin-4 induces c-Fos activation and CREB/ATF-1 phosphorylation in the

hypothalamic neuronal cell lines.............................................................................

Figure 20. Regulation of neurotensin and ghrelin mRNA expression by exendin-4 in the

mHypoA-2/30 and mHypoE36/1 neuronal cell lines..............................................

Figure 21. Regulation of neurotensin and ghrelin mRNA expression via activation of the

protein kinase A pathway........................................................................................

Figure 22. Intracerebroventricular injection of h(Gly2)GLP-2 inhibits food and

water intake, and induces weight loss in a dose-dependent manner in mice..........

Figure 23. Acute h(Gly2)GLP-2 treatment activates hypothalamic appetite-regulating

nuclei.......................................................................................................................

Figure 24. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity (ir) in the

hypothalamic neurons expressing α-MSH-, NPY-, neurotensin- or ghrelin-ir.......

Figure 25. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity in the

hypothalamic neurons expressing NPY-, neurotensin- or ghrelin-ir......................

Figure 26. Graphical representation of neurons expressing c-Fos- and neuropeptide-

immunoreactivity in the ARC, VMH, DMH, PVN, LH and internuclear space

between the DMH and LH of the saline- or h(Gly2)GLP-2-treated mouse

hypothalamus..........................................................................................................

Figure 27. Expression profile of GLP-2R and appetite-regulating neuropeptides in the

hypothalamic neuronal cell lines.............................................................................

Figure 28. Acute h(Gly2)GLP-2 treatment induces c-Fos activation and CREB/ATF-1

phosphorylation in the hypothalamic GLP-2R-positive mHypoA-2/30 neuronal

cells.........................................................................................................................

Figure 29. Regulation of neurotensin and ghrelin mRNA expression by h(Gly2)GLP-2 in

the mHypoA-2/30 neuronal cell line in protein kinase A-dependent manner…....

84

86

89

92

94

107

109

110

111

112

114

116

117

xiii

Figure 30. Summary of the mechanisms activated by insulin and leptin to regulate

proglucagon gene expression in the mHypoA-2/10 and mHypoE-39 neuronal

cells.........................................................................................................................

Figure 31. Summary of the mechanisms activated by exendin-4 to regulate neurotensin

and ghrelin gene expression in the mHypoA-2/30 and mHypoE-36/1 neuronal

cells.........................................................................................................................

Figure 32. Summary of the mechanisms activated by GLP-2 to regulate neurotensin and

ghrelin gene expression in the mHypoE-36/1 neuronal cells..................................

128

130

132

xiv

List of Abbreviations

aa

AgRP

AMP

cAMP

ANOVA

AP-1

ATF-1

ARC

ATP

AMPK

BBB

BCA

BDNF

CART

CNS

CNTF

CRE

CREB

CRH

DAB

DMEM

DMH

DMSO

DNA

cDNA

amino acid

agouti-related peptide

adenosine monophosphate

cyclic adenosine monophosphate

analysis of variance

activator protein-1

activating transcription factor-1

arcuate nucleus of the hypothalamus

adenosine-5’- triphosphate

AMP-dependent kinase

blood-brain barrier

bicinchoninic acid

brain-derived neurotrophic factor

cocaine- and amphetamine-regulated transcript

central nervous system

ciliary neurotrophic factor

cAMP-response element

cAMP-response element-binding protein

corticotrophin-releasing hormone

diaminobenzidine

dulbecco’s modified Eagle medium

dorsomedial nucleus of the hypothalamus

dimethyl sulfoxide

deoxyribonucleic acid

complementary deoxyribonucleic acid

xv

DPP-4

DRB

Epac

ERK

FBS

GABA

GHS-R

Ghrl−/−

GLPs

GLP-1

GLP-1R

GLP-2

GLP-2R

GnRH

GRPP

h(Gly2)GLP-2

GUE

HPA

i.c.v.

IGF-1

IP

-ir

IBMX

IHC

IR

IRS

dipeptidyl peptidase-4

5, 6-Dichlorobenzimidazole 1-β-D-ribofuranoside

exchange protein activated by cAMP

extracellular signal-regulated kinases

fetal bovine serum

-aminobutyric acid

G protein beta subunit

ghrelin receptor (growth hormone secretagogue receptor)

ghrelin-deficient

glucagon-like peptides

glucagon-like peptide-1

GLP-1 receptor

glucagon-like peptide-2

GLP-2 receptor

gonadotropin-releasing hormone

glicentin-related pancreatic peptide

human (Gly2)GLP-2(1-33)

proglucagon gene upstream enhancer element

hypothalamic-pituitary-adrenal axis

intracerebroventricular

insulin-like growth factor-1

intervening peptide

-immunoreactive

3-isobutyl -1-methylxanthine

immunohistochemistry

insulin receptor

insulin receptor substrate

xvi

JAK

LH

MAPK

MCH

MC3R

MC4R

ME

MPGF

α-MSH

NPY

NT-AS

NTS

Ntsr

Ob-Rb

PBS

PC1/3

PC2

PeV

PGDPs

PI3K

PIP3

PKA

PKC

PKI

PVN

mRNA

miRNA

janus kinase

lateral hypothalamus

mitogen-activated protein kinase

melanin-concentrating hormone

melanocortin-3 receptor

melanocortin-4 receptor

median eminence

major proglucagon fragment

α--melanocyte stimulating hormone

Neuropeptide Y

neurotensin antiserum

nucleus of the solitary tract

neurotensin receptor

leptin receptor

phosphate buffered saline

prohormone convertase 1/3

prohormone convertase 2

periventricular nucleus of the hypothalamus

proglucagon-derived peptides

phosphoinositide-3-kinase

phosphatidylinositol-3,4,5-triphosphate

protein kinase A

protein kinase C

PKI 14-22 amide, myristoylated

paraventricular nucleus of the hypothalamus

messenger ribonucleic acid

microRNA

xvii

siRNA

RBPDB

RT-PCR

qRT-PCR

SDS-PAGE

STAT

SV

TRH

VMH

WHO

small interfering RNA

RNA-binding protein database

reverse transcription-polymerase chain reaction

quantitative reverse transcription-polymerase chain reaction

sodium dodecyl sulfate polyacrylamide gel electrophoresis

signal transducer and activator of transcription

simian virus

thyrotropin-releasing hormone

ventromedial nucleus of the hypothalamus

World Health Organization

1

Chapter 1

Background and Significance

2

Published figure:

Figure 1. Schematic diagrams of the hypothalamic and brainstem regions that express

neuropeptides involved in energy homeostasis (coronal sections).

Prasad S. Dalvi, Anaies Nazarians-Armavil, Stephanie Tung, and Denise D. Belsham.

Immortalized neurons for the study of hypothalamic function. Am J Physiol Regul Integr Comp

Physiol (2011) 300:R1030-1052.

Permissions were obtained to reproduce the copyrighted material.

3

1 Background and Significance

1.1 Introduction

Over the past two decades, an escalating pandemic of obesity has affected millions of

people worldwide (1, 2). According to Statistics Canada, more than half of Canadians are

overweight or obese, with almost 23% of the population falling in the category of "obese".

Obesity leads to numerous complications, such as type 2 diabetes, hypertension, stroke and heart

disease, that contribute to the increase in global burden of morbidity and mortality among all age

groups of human population (3). Increased appetite, indiscriminate and enhanced high calorie

consumption coupled with sedentary and stressful life-styles, and significant contributions of

genetic factors are responsible for dysregulation of energy balance that leads to the development

of obesity (4, 5). Since dieting alone is unable to control body weight in obese individuals, more

research is focused on the development of appetite-regulating drugs for which it is imperative to

understand mechanisms underlying central regulation of appetite and energy balance. The long-

term objective of this research project is to find new insights into the control mechanisms that are

critical to understand regulation of appetite and energy homeostasis. The main control centre of

appetite and energy homeostasis in the brain is the hypothalamus, which is comprised of multiple

neuronal subtypes expressing either appetite-stimulating (orexigenic) or appetite-suppressing

(anorexigenic) neuropeptides. Among them, the proglucagon-derived peptides (PGDPs) have

emerged as potential regulators of feeding behavior, particularly glucagon-like peptide (GLP)-1

and GLP-2, which induce anorexigenic effects via stimulation of GLP-1 receptor (GLP-1R) and

GLP-2R, respectively (6, 7). Two key anorexigenic hormones, insulin, and leptin, are known as

adiposity signals and maintain energy homeostasis through the regulation of hypothalamic

neurons (8-10). Several key studies have demonstrated the importance of central insulin actions

in regulating energy homeostasis (11-14). Similarly, leptin, an adipocyte-derived hormone acts in

the hypothalamus to reduce feeding and body weight (15-17). Insulin and leptin regulation of the

hypothalamic proglucagon gene expression, and also the cellular mechanisms triggered by GLP-

1 and GLP-2 in the process of appetite regulation are poorly understood, mainly due to

complexity of the in vivo architecture of the hypothalamus and the lack of representative

neuronal cell models required for these studies. Recently, our laboratory has generated an

4

extensive collection of immortalized, clonal hypothalamic neuronal cell lines that can be used to

determine the effects of hormones and neuropeptides on specific neuroendocrine cell types (18,

19). Some of these cell models express the proglucagon gene and can be used to study

mechanisms involved in proglucagon gene regulation. Importantly, we have the first

hypothalamic cell lines that endogenously express GLP-1R and GLP-2R, and can be used to

study hormone-responsiveness, at the level of the individual neuropeptide gene and through

analysis of signal transduction events. Using these cells as in vitro models and mice as in vivo

models in these studies, the short-term objectives of this research project were to study

regulation of proglucagon gene expression in the hypothalamic neurons, to assess the effect of

long-acting GLP-1R agonist exendin-4 and GLP-2R agonist h(Gly2)GLP-2[1-33] on modulation

of hypothalamic appetite-regulating neuropeptides and to dissect the mechanisms involved in

this process. First, using the hypothalamic cell models, potential action of insulin and leptin, two

key regulators of food intake and energy balance, on proglucagon mRNA transcript levels in the

hypothalamic cell models was investigated. Second, the transcriptional mechanisms involved in

the regulation of proglucagon gene expression were defined. Finally, the action of exendin-4 and

h(Gly2)GLP-2[1-33] on the hypothalamic neuropeptides neurotensin and ghrelin, and the

signaling mechanisms involved were investigated.

1.2 Regulation of energy homeostasis

The central nervous system (CNS) regulates energy balance via an evolutionarily

conserved homeostatic system. Over the last two decades, extensive research has been conducted

to investigate neuronal pathways, neuropeptides, neurotransmitters and modulators, and signal

transduction mechanisms involved in appetite regulation and energy balance. In the CNS, the

hypothalamus is a primary control site that regulates central neuroendocrine functions including

but not limited to thermoregulation, fluid and electrolyte balance, reproduction, circadian

regulation, stress response, and energy homeostasis (20). Afferent signals from the periphery are

integrated in the neuroendocrine hypothalamus and processed into efferent signals to modulate

food intake and energy expenditure in order to maintain energy balance (21). Dysregulation of

this homeostatic system in favor of increased energy intake may lead to metabolic disorders that

further result in obesity and related complications such as type 2 diabetes (22).

5

1.2.1 Hypothalamus

The hypothalamus plays a vital role in maintaining the energy homeostasis (23). It

comprises of several nuclei that regulate energy intake and expenditure. The main nuclei

involved in appetite regulation are the arcuate nucleus (ARC), paraventricular nucleus (PVN),

ventromedial nucleus (VMH), dorsomedial nucleus (DMH), and lateral hypothalamus (LH)

(Figure 1). The ARC, PVN, and DMH are the most important hypothalamic nuclei involved in

the regulation of energy balance, and the hypothalamic VMH and LH are involved in the control

satiety and hunger, respectively (24, 25). These hypothalamic nuclei contain specific neuronal

phenotypes to form a complex network of orexigenic and anorexigenic circuits that regulate

energy homeostasis (20). Numerous neuropeptides, including but not limited to neuropeptide Y

(NPY), agouti-related peptide (AgRP), melanin-concentrating hormone (MCH), α-melanocyte

stimulating hormone (α-MSH)/pro-opiomelanocortin (POMC), corticotrophin-releasing hormone

(CRH), cocaine- and amphetamine-regulated transcript (CART), brain-derived neurotrophic

factor (BDNF), orexin A and B, glucagon-like peptides (GLPs), galanin, ghrelin, and neurotensin

mediate orexigenic and anorexigenic processes in the hypothalamus (6). Recently, the PGDPs,

specifically, glucagon-like peptide-1 (GLP-1), glucagon-like peptide-2 (GLP-2) and

oxyntomodulin, have been investigated as potential regulators of feeding behavior (6). Several

studies have shown that central administration of PGDPs controls satiety and reduces food intake

(26-28), however further research is warranted to dissect the mechanism(s) utilized by PGDPs to

induce anorectic signals and their exact role in appetite regulation (7).

1.2.1.1 Arcuate nucleus

The ARC encloses the third ventricle and lies immediately above the median eminence

(ME). The ARC-ME area represents an important region because the lack of blood-brain barrier

(BBB) in ME permits the interplay between peripheral organs and the brain to take place in this

small area (29). Due to its multitude of neuropeptidergic neurons, it is a key hypothalamic

nucleus in the regulation of appetite. The two major neuronal populations in the ARC implicated

in the regulation of feeding are orexigenic NPY and AgRP and anorexigenic POMC. NPY is a

powerful central appetite-stimulant and acts upon its own receptors, Y1, Y2, Y4, Y5 (30), but

exerts its orexigenic effect predominantly via the Y1 and Y5 receptors. The majority of the

6

hypothalamic NPY neurons found within ARC also co-express AgRP (31, 32). AgRP is part of

the melanocortin system and is an antagonist to the melanocortin-3 receptor (MC3R) and

melanocortin-4 receptor (MC4R) (33). Thus, when the NPY/AgRP neuron is activated, NPY

stimulates orexigenic pathways and AgRP inhibits the anorexigenic melanocortin pathway.

Hypothalamic POMC neurons are primarily located in the lateral ARC region (34, 35).

Cleavage of POMC by endoproteases yields active peptides including melanocortins that are

essential for normal energy homeostasis (10). Feeding stimulates hypothalamic POMC mRNA

expression, and mice that are obese due to leptin deficiency or insensitivity exhibit decreased

hypothalamic POMC mRNA expression (35-37). Mice and humans entirely lacking POMC

develop hyperphagia and severe early-onset obesity (38). Thus, hypothalamic POMC neurons

mediate catabolic responses such as decreased food intake and increased energy expenditure. In

fact, the POMC peptide product α-MSH has been observed to have significant catabolic effects

through its action on MC4R (39, 40). The majority of POMC neurons in the ARC also co-

express CART mRNA. Animal studies have shown that central administration of CART inhibits

food intake, whereas central injection of CART antiserum increases food intake (41).

The analysis of the hormonal control of NPY/AgRP and POMC neurons has progressed

immensely. Both neuronal populations express insulin and leptin receptors and are targeted by

the respective hormones (12, 42, 43). A large body of evidence has revealed an important role

for leptin-activated STAT3 signaling in the arcuate POMC/AgRP neurons in control of energy

homeostasis. Disruption of neuron-specific STAT3 results in hyperphagia, obesity, diabetes and

infertility (44). Further, it was demonstrated that insulin- and leptin-evoked phosphoinositide-3-

kinase (PI3K) activation resulted in Akt-mediated disinhibition of POMC transcription (45, 46).

The projections from the “first order” NPY/AgRP and POMC neurons extensively

communicate with the “second-order” neurons, located in other hypothalamic areas involved in

appetite regulation, such as the PVN, VMH, DMH, LH, and the nucleus of the solitary tract

(NTS) in the brainstem (47). The axons of these neurons project to CRH-, thyrotropin-releasing

hormone (TRH)- and oxytocin-expressing neurons located in the PVN, and MCH- and

orexin/hypocretin-expressing neurons of the LH and perifornical area. The ARC is the chief

hypothalamic area involved in the control of food intake, therefore, when satiety or adiposity

7

signals reach the ARC, anorexigenic peptides are released which activate a catabolic circuit to

suppress feeding. In contrast, when satiety or adiposity signals are low in the brain, the activation

of anabolic pathway leads to the release of orexigenic peptides indicating the urgency to

replenish fuel stores by stimulating food intake.

1.2.1.2 Paraventricular nucleus

The PVN is located on either side of the superior part of the third ventricle in the anterior

hypothalamus. The PVN is a highly vascularised region of the hypothalamus and is protected by

the BBB, although its neuroendocrine neurons extend to ME and the posterior pituitary which

lack the BBB. It contains a dense cluster of heterogeneous neurons that are activated by various

stimuli including stress and physiological changes (48, 49). The magnocellular neurosecretory

neurons of the PVN project directly to the neurohypophysis (posterior pituitary) where they

release oxytocin or vasopressin into the general circulation (50, 51). The parvocellular

neurosecretory neurons of the PVN project to the ME and transmit CRH, TRH, gonadotropin-

releasing hormone (GnRH), growth hormone-releasing hormone, somatostatin and dopamine

into the portal blood vessels of the adenohypophysis (anterior pituitary) (52). There are other

PVN neurons that express neuropeptides such as ghrelin, CART, nociceptin and neurotensin (53-

56), which may be involved in the hypothalamic regulation of appetite and autonomic functions

in the brainstem and spinal cord. Evidence from many experiments suggests that the PVN is a

crucial site for the action of many peptides involved in appetite regulation including but not

limited to NPY, ghrelin, orexin-A, leptin, GLP-1 (57). It has been shown that microinjection of

CRH and leptin in the PVN attenuated fasting-induced feeding (58, 59). The NPY/AgRP neurons

of the ARC innervate TRH neurons in the PVN and inhibit pro-TRH gene expression, whereas

α-MSH/POMC projections stimulate TRH (60-62). Thus by integrating numerous neuronal

pathways implicated in energy balance such as projections from the NPY/AgRP and POMC

neurons of the ARC as well as projections from the suprachiasmatic nucleus and the NTS the

PVN plays an important role in the integration of nutritional signals with the hypothalamo-

pituitary axis and other downstream endocrine glands (49, 63).

8

Figure 1. Schematic diagrams of the hypothalamic and brainstem regions that express

neuropeptides involved in energy homeostasis (coronal sections). AgRP, agouti-related

peptide; ARC, arcuate nucleus; AP, area postrema; BBB, blood-brain barrier; CART, cocaine

and amphetamine-regulated transcript; CB1, endocannabinoid receptor 1; CRH corticotropin-

releasing factor; DMH, dorsomedial hypothalamic nucleus; GLP-1, glucagon-like peptide-1;

GLP-2, glucagon-like peptide-2; LH, lateral hypothalamus; MCH, melanin-concentrating

hormone; ME, median eminence; NPY, NPY; NT, neurotensin; NTS, nucleus tractus solitarius;

ObRb, leptin receptor; OC, optic chiasm; OXM, oxyntomodulin; OXY, oxytocin; Pit, pituitary

gland; POMC, pro-opiomelanocortin; PVN, paraventricular nucleus; SCN, suprachiasmatic

nucleus; 3V, third cerebral ventricle; TRH, thyrotropin-releasing hormone; VMH, ventromedial

hypothalamic nucleus. Orange arrows point to the coronal sections of the corresponding brain

regions. Dashed arrows indicate the central sites of action for regulating factors.

9

1.2.1.3 Ventromedial hypothalamic nucleus

The VMH of the hypothalamus, located adjacent to the ARC, receives NPY, AgRP, β-

endorphin and POMC/CART projections from the ARC (41, 64-67), and sends efferent

projections to the DMH, PVN and brain stem regions such as NTS. The neurons in the VMH

have long been hypothesized to play a major role in metabolic regulation since bilateral VMH

lesions cause hyperphagia and obesity (68-70). The VMH contains a large population of

glucoresponsive neurons which show increased activity during oral hyperglycemia (71, 72).

Recent work has demonstrated that BDNF is highly expressed in the VMH and central

administration of BDNF reduces feeding and causes loss of body weight (73). The VMH BDNF

neurons are regulated by food deprivation (74) and ARC POMC neurons via MC4R activate

them to decrease food intake (74). It has been found that the VMH expresses receptors to several

signalling molecules including leptin, which is a potent anorexigenic signal to inhibit appetite

and feeding and stimulate energy expenditure and weight reduction (75).

1.2.1.4 Dorsomedial hypothalamic nucleus

The DMH receives a high level of ARC NPY and α-MSH/POMC neuronal terminals (32,

76), whereas it sends projections to communicate with VMH, PVN and LH to integrate and

process information from these regions (15, 77). It has been shown that the electrolytic

destruction of the DMH disrupts feeding causing hyperphagia and obesity that suggests a role for

DMH in appetite regulation (78). Other findings have confirmed that DMH may act as a site of

action for leptin and interaction between NPY and leptin (79, 80). Normally, the DMH contains a

few NPY neurons (81, 82), however in diet-induced obese, in genetic obese mice and during

suckling-induced hyperphagia NPY gene expression is increased in the DMH (15, 83-85).

1.2.1.5 Lateral hypothalamus

The LH extends rostrally from the mesencephalic tegmentum to the lateral preoptic area

in dorsal and lateral aspect to VMH (77). It contains sparsely distributed subpopulation of

neurons expressing orexigenic MCH and orexins (orexin A and B or hypocretin-1 and -2) (86,

87) which are extensively synapsed by projections from NPY/AgRP and α-MSH/POMC neurons

from the ARC. MCH-immunopositive fibers further project to the cortex, brainstem and spinal

10

cord (88), and orexin/hypocretin neurons exert their effects via extensive communication with

the ARC, PVN and NTS (86, 89). The mechanism by which MCH and orexin/hypocretin

neurons regulate energy homeostasis is yet unclear. Apart from these neurons, large number of

glucose-sensing neurons are present in the LH that may integrate peripheral signals and mediate

the marked hyperphagia which is normally induced by hypoglycemia (24).

1.2.2 The Hypothalamic Neuropeptides

1.2.2.1 Neuropeptide Y

NPY is a 36-amino acid (aa) peptide and is one of the most potent neuropeptides known

to induce feeding in animals. NPY is one of the most abundant peptides found in the brain and

NPY-expressing neurons are widely distrubuted throughout the CNS. In the hypothalamus, NPY

is abundantly synthesized by the NPY-expressing neurons of the ARC and are known as the key

orexigenic neurons (90). These arcuate NPY neurons send projections towards the PVN as well

as the DMH (91). The widespread distribution of NPY facilitates functional diversity, including

cardiovascular regulation, seizure and cognition, stress, modulation of neuroendocrine systems

and appetite regulation. NPY mediates its biological actions through a portfolio of G-protein

coupled receptors, Y1-Y6 and second messenger systems in different downstream effector

neuronal types (30, 92, 93). Several studies suggest that the Y1 and Y5 receptors are important in

mediating the orexigenic effects of NPY in rats (94).

Negative energy states such as fasting, starvation, food deprivation, diet restriction or

hunger induce an increase in NPY neurons in the hypothalamus and NPY mRNA expression in

various brain nuclei (77, 95-101), whereas positive energy state following refeeding normalizes

both NPY secretion and NPY mRNA levels (102). NPY neurons are regulated by numerous

metabolic factors such as glucose, free fatty acids, insulin, leptin, ghrelin and other hormones

such as glucocorticoids (103-107). It has been shown that the satiety hormones insulin and leptin

decrease NPY expression (103, 104), and acute exposure to glucose inhibits NPY neurons (122).

In contrast, the orexigenic hormone ghrelin directly stimulates glucose-sensitive ARC NPY

neurons (103, 105-107). Glucocorticoid receptors are found on almost all NPY neurons in the

ARC (108), and glucocorticoids upregulate NPY expression in the hypothalamus (109).

11

In the genetically obese Zucker rats, NPY mRNA is increased in the ARC nucleus

causing hyperphagia in these rats (110). Similarly, hypothalamic NPY gene expression is

upregulated and peptide content in the ARC nucleus is significantly increased in the obese leptin-

deficient ob/ob mice (102, 111). Importantly, in adult rats central administration of NPY

stimulates food intake, whereas ablation of NPY/AgRP neurons in young mice reduces food

intake and causes severe starvation highlighting the importance of these neurons to energy

balance (112-114). However, neonatal NPY/AgRP ablation does not elicit any changes in

feeding behavior (112). Similarly, the phenotype of NPY knockout mice does not alter from that

of wild-type mice (115). The absence of any evident effect of NPY deletion on phenotype of

NPY knockout mouse model can be attributed to the compensatory adaptive mechanisms that

may develop during early developmental stages. Nevertheless, it has been demonstrated that

NPY ablation in ob/ob mice attenuates obesity through decreased feeding and weight gain

suggesting that the NPY neurons play a crucial role in the development of obesity (116).

1.2.2.2 Pro-opiomelanocortin

POMC neurons are expressed in the ARC of the hypothalamus and the anterior and

intermediate lobes of the pituitary gland. POMC is a precursor polypeptide that undergoes

extensive, tissue-specific, post-translational processing via cleavage by enzymes, such as

prohormone convertases, carboxypeptidase E, peptidyl α-amidating monooxygenase, N-

acetyltransferase, and prolycarboxypeptidase. The POMC-derived peptides have important roles

in the regulation of appetite and energy homeostasis (117, 118). Mutations in the POMC gene in

humans have been associated with hyperphagia and early onset obesity (119-121), and POMC-

deficient mice are severely obese (38, 122).

In the hypothalamus, enzymatic cleavage of POMC produces melanocortin peptides

including α-MSH which exert their effects via melanocortin receptors. The melanocortin

receptors MC3R and MC4R are expressed in the brain, but the MC4R is highly expressed in the

hypothalamus, particularly the PVN (123). It is now clear that the α-MSH released from ARC

POMC neurons activates downstream MC4Rs to inhibit food intake (124), whereas AgRP from

the ARC, being the endogenous inverse agonist and antagonist of the MC3R and MC4R,

reverses the α-MSH-mediated inhibition of feeding and stimulates food intake and a lower

12

metabolic rate (31). The physiological role of POMC-derived peptides in the brain is of

particular interest in obesity research as it has been shown that α-MSH together with the MC3R

and MC4R is central to the regulation of appetite and energy homeostasis. In vivo studies have

shown that POMC-derived peptides have potent anorexigenic effects, but α-MSH has the most

potent anorexigenic effects as it induces severe hypophagia and causes the greatest reduction in

body weight when administered to POMC-null mice (125). These findings highlight the

importance of this precursor polypeptide in the homeostatic regulation of feeding and body

weight (125). Further, mutations in the MC4R cause obesity, while agonists of the receptor

function suppress feeding (126-128). The MC4R-deficient patients are hyperphagic and up to 6%

of early-onset morbidly obese patients in British Caucasians were found to have MC4R

mutations (129, 130).

Recently, it was found that POMC is colocalized with GLP-1R in ARC neurons, and

another recent study indicates that the GLP-1R activation in the ARC POMC neurons is involved

in regulation of glucose homeostasis, but not feeding (131). Whether GLP-1 has any direct effect

on POMC expression or secretion of POMC-derived peptides is yet to be determined.

1.2.2.3 Neurotensin

The coordinated regulation of food intake and energy expenditure takes place in the

hypothalamus. Among the multitude of central signaling pathways and neuropeptides involved in

appetite regulation, hypothalamic neurotensin inhibits feeding due to its anorexigenic action

(132, 133). Neurotensin-expressing neurons are located mainly in the parvocellular portion of the

PVN, in the ARC, PeV, and LH. Moderate numbers of cell bodies containing neurotensin-like

immunoreactivity can also be found in the anterior hypothalamus, the DMH, and the posterior

hypothalamus (56). Neurotensin mediates its effect through neurotensin receptors (Ntsr) (134),

which are expressed in the hypothalamic ARC and DMH nuclei (135-138). Ntsr deficiency

moderately increases food intake and body weight, and blocks neurotensin-induced anorexia in

mice, implicating neurotensin-Ntsr signaling pathway in feeding and body weight regulation

(139).

Leptin is one of the most important adipose-derived hormones that reduces appetite by

regulating several central and peripheral signaling pathways involved in the regulation of energy

13

homeostasis (140). Recently, it was found that leptin directly stimulates neurotensin gene

expression in the hypothalamic cell lines that express both leptin receptor and neurotensin (141).

Furthermore, it has been suggested that leptin’s inhibitory regulation on feeding is mediated at

least partly through neurotensin-Ntsr signaling pathway (142, 143). All these findings confirm

that leptin targets neurotensin neurons to mediate its effect on appetite regulation. Interestingly,

in another study, it was found that when GLP-1R activation was blocked by exendin-(9-39),

leptin’s anorectic action was abolished (144). Further, exendin-4 treatment decreased food intake

and body weight in leptin-deficient ob/ob mice and fatty Zucker rats, indicating that leptin

deficiency is overcome by GLP-1R activation to induce leptin’s anorectic effect (145). Based on

these findings it can be strongly suggested that leptin may interact with intermediate proglucagon

neurons that subsequently excite downstream negative energy balance by stimulating

anorexigenic neurons to induce appetite suppression; neurotensin neurons can be the potential

downstream targets for the action of central GLPs. As yet, no studies have been performed to

link hypothalamic GLPs or their receptors with neurotensin.

1.2.2.4 Ghrelin

Ghrelin is a potent orexigenic hormone that strongly influences generation of hunger and

therefore is known as the “hunger hormone”. Its plasma concentrations increase before meals

and during fasting, and decrease after ingestion of food. Ghrelin is predominantly produced in

the stomach and stimulates appetite by its action on the ARC of the hypothalamus via activation

of growth hormone secretagogue receptor (GHS-R). Two isoforms of ghrelin have been

identified: active (acyl ghrelin) and inactive (des-acyl ghrelin). The enzyme ghrelin O-

acyltransferase converts proghrelin peptide into active acyl ghrelin (146, 147). Des-acyl ghrelin

does not bind to the GHS-R, and its biological roles are uncertain in the absence of an identified

equivalent receptor (148). Apart from the production of mature ghrelin, the proghrelin peptide is

also post-translationally processed in the stomach to generate an entirely different peptide,

obestatin, that apparently activates an orphan G-protein-coupled receptor, GPR39 (149).

However, the physiological function of obestatin and its receptor remains unclear. Although,

initial research demonstrated that obestatin had a potent anorectic effect in mice and rats (149),

subsequently these findings were disputed (150, 151).

14

Human and animal studies have demonstrated that the activation of the GHS-R results in

increased food intake and increased adiposity (152-154). Peripheral and central administration of

ghrelin increases feeding and promotes weight gain (153, 154). It is reported that endogenous

centrally released ghrelin is also involved in the regulation of food intake and body weight (153,

155). Existence of ghrelin-expressing neurons in the hypothalamus is debated, however, ghrelin-

expressing neurons have been shown to be present in the internuclear spaces between the PVN,

ARC, VMH, and DMH hypothalamic nuclei, the perifornical region, and the ependymal layer of

the third ventricle (105). These neurons interact with key hypothalamic neurons involved in

appetite regulation, mainly the NPY/AgRP and POMC neurons of the ARC suggesting that

hypothalamic ghrelin may stimulate orexigenic or inhibit anorexigenic neuropeptides (105, 156).

Although a series of pharmacological and clinical studies suggested that ghrelin was an

endogenous regulator of energy balance, recent genetic studies using adult ghrelin-deficient

(Ghrl−/−) mice resulting from targeted ghrelin gene disruption fail to demonstrate any significant

defects in normal energy balance, food intake, and adiposity on a standard diet (157, 158). Such

lack of a metabolic phenotype may be attributed to compensatory processes during early

developmental phases or existence of redundant pathways controlling energy balance.

Nevertheless, absence of ghrelin in Ghrl−/− mice protects them from the early-onset obesity

induced by early exposure to a high-fat diet (159). Importantly, simultaneous deletion of ghrelin

and its receptor leads to decreased body weight, increased energy expenditure, and increased

motor activity on a standard diet demonstrating that ghrelin has a physiological role in the

regulation of energy homeostasis (160). Another finding that the mice lacking GHS-R are

resistant to the development of diet-induced obesity demonstrates the importance of GHS-R in

the regulation of energy expenditure and body weight (161).

Regulation of ghrelin-expressing neurons in the hypothalamus is not well studied.

Although it has been shown that GLP-1 inhibited ghrelin-induced feeding and exendin-4 reduced

plasma levels of peripheral ghrelin (162, 163), whether hypothalamic ghrelin expression is

regulated by GLP-1R or -2R activation is not known. It can be speculated that the anorexigenic

action of the centrally-originated GLPs is mediated via inhibition of hypothalamic ghrelin

neurons; however, no studies have been conducted to elucidate this mechanism. Further

investigation is needed to explore the possibility of this interaction because inhibition of the

15

ghrelin-expressing neurons may lead to reduced food intake, and consequently to reductions in

body weight and adiposity. Therapeutic intervention by inhibiting ghrelin/GHS-R pathway may

be used to prevent or treat obesity (164).

1.3 Proglucagon-derived peptides

1.3.1 Biosynthesis and regulation

The proglucagon gene was isolated from rodents and humans in the early 1980s (165-

170). The proglucagon gene is expressed in the pancreatic α-cells, intestinal L-cells and a small

number of neurons in the CNS, specifically the hypothalamus and the brain stem, and a single

copy of proglucagon gene generates the PGDPs (166, 168, 171, 172). Identical preproglucagon

mRNA transcripts are expressed in these cells, however, the tissue- or cell-specific post-

translational processing of proglucagon leads to the biosynthesis of diverse PGDPs (173, 174)

(Figure 2A). Glicentin, oxyntomodulin, GLP-1, GLP-2, intervening peptide (IP)-1 and IP-2 are

synthesized in the intestinal endocrine L-cells which are located mainly in the distal ileum and

colon, whereas glucagon, glicentin-related pancreatic peptide (GRPP), major proglucagon

fragment (MPGF) and IP-1 are predominantly produced by pancreatic α-cells (171, 173-176).

This diversification is achieved through proteolytic processing by cell- or tissue-specific

prohormone convertases (174, 177-179). Currently, there are 9 known mammalian prohormone

convertases, all of which are thought to be responsible for the cleavage of many prohormones

and proneuropeptides to generate shorter forms by targeting the carboxyl terminus of a single or

paired basic residues of these prohormones and proneuropeptides (177). In the case of precursor

proglucagon, each PGDP is flanked by pairs of basic residues that present canonical

prohoromone cleavage sites for different prohormone convertases to exert their catalytic

functions (171).

The tissue-specific expression of the different convertases is responsible for the unique

protein profile of PGDPs. Prohormone convertase 2 (PC2) is the most abundant convertase in the

pancreatic α-cells and co-localizes with glucagon in the islet secretory granules (180). In

contrast, intestinal L-cells highly express prohormone convertase 1/3 (PC1/3) that is responsible

for the generation of the intestinal PGDPs including GLP-1 and GLP-2 (174, 180-182). Both

16

PC2 and PC1/3 are capable of cleaving proglucagon to generate the N-terminal intermediate

glicentin and C-terminal intermediate MPGF (181); however PC2 is required for additional

processing of glicentin to GRPP, glucagon and IP-1 (173, 181), and PC1/3 is required to produce

GLP-1 and GLP-2 from MPGF (173, 174, 180-182). The post-translational processing of

proglucagon in the CNS is mostly similar to that observed in the intestinal L-cells (183). Original

studies investigating expression pattern of hypothalamic PGDPs showed that glicentin,

oxyntomodulin and glucagon are abundantly present in fetal rat hypothalamus, whereas in adult

rat hypothalamus glicentin and oxyntomodulin are present in greater amounts than glucagon and

GLP-1 suggesting that the processing of proglucagon in the hypothalamus varies with the brain

development (183, 184).

Regulation of proglucagon gene expression in the pancreas and intestine has been

extensively studied (185). Extensive investigations of mechanisms underlying proglucagon gene

transcription have led to the identification of a minimum promoter region (G1) and four enhancer

elements (G2–G5) in the proglucagon gene promoter (186, 187). Transgenic mouse studies have

indicated that approximately 1.3 kb of rat proglucagon gene 5’flanking sequences are sufficient

to direct pancreatic α-cell- and brain-specific rat proglucagon gene expression (188). In contrast,

similar studies suggest that a much larger region (2.3 kb) of rat proglucagon 5’-flanking

sequences is required for proglucagon gene expression in the intestine (189), indicating that

DNA sequences located between -2.3 and -1.3 kb in the rat proglucagon promoter contain

specific elements required for intestinal proglucagon gene expression. These sequences situated

between -2.3 and -1.3 kb have been designated as the proglucagon gene upstream enhancer

element (GUE) that is composed of multiple positive and negative cis-acting DNA regulatory

elements involved in regulating tissue-specific proglucagon gene transcription (190). A

combination of cell transfection and transgenic reporter studies have reported that approximately

1.6 kb of human proglucagon gene 5’-flanking sequences can direct proglucagon gene

transcription to the brain and intestine, but for pancreas-specific proglucagon gene expression the

sequences within the first 6 kb of the human proglucagon gene 5’-flanking region are required

(191). These studies indicate that specific trans- and cis-regulatory elements are required for

tissue-specific proglucagon gene transcription in rodents and humans.

17

Several studies have demonstrated that proglucagon gene transcription can be regulated

by cyclic adenosine monophosphate (cAMP), amino acids, and a number of homeodomain

protein transcription factors such as isl-1, cdx-2/3, pax-6, HNF-3α, HNF-3β, and brn 4 (185,

192-196). Further, FoxO1 has been identified as a critical regulator of proglucagon gene

expression by insulin in pancreatic αTC1-9 cells (197). Recently, the Wnt/TCF-4 pathway was

shown to be involved in the regulation of proglucagon gene expression by insulin in

enteroendocrine cells (198). Furthermore, pathological levels of insulin were found to stimulate

intestinal proglucagon mRNA and GLP-1 production (199), in contrast to insulin’s known

inhibitory effect on proglucagon gene expression in pancreatic α-cells (200, 201). Although it is

known that the activators of cAMP/protein kinase A (PKA) pathway stimulate synthesis and

secretion of the PGDPs in the hypothalamus (202), and excitatory amino acid glutamate

stimulates PGDP secretion through a protein kinase C-dependent pathway (203), other internal

as well as external factors involved in regulating these neurons are not fully identified. At

present, the mechanisms involved in hypothalamic proglucagon gene regulation remain largely

unknown.

1.3.2 Role of PGDPs as appetite regulators

The appetite-related effects of PGDPs are well documented; particularly GLP-1, GLP-2

and oxyntomodulin are involved in inhibition of food intake (26, 204, 205). The actions of GLP-

1, GLP-2 and oxyntomodulin are exerted through receptors that belong to the family of 7-

transmembrane-spanning G-protein-coupled receptors, members of glucagon receptor

superfamily. GLP-1 and oxyntomodulin act via the GLP-1R, and GLP-2 specifically activates

the GLP-2R (206, 207).

GLP-1, GLP-2 and oxyntomodulin are mainly released from intestinal epithelial

endocrine L-cells in response to food intake. GLP-1 is inactivated by the ubiquitous enzyme

dipeptidyl peptidase-4 (DPP-4) in less than 2 minutes. Therefore, the long acting DPP-4 resistant

GLP-1R agonist exendin-4 has been developed (208). Acute administration of GLP-1 or

exendin-4 suppresses food intake inducing satiety, and chronic administration results in weight

loss (26, 209-211). GLP-2, similar to GLP-1, is also inactivated by DPP-4 thereby having a

relatively short biological half-life of approximately 7 minutes. Consequently, a degradation-

18

resistant analog (Gly2)GLP-2 (Teduglutide) is used in biomedical research (212). GLP-2

administration results in inhibition of gastric emptying and reduced intestinal motility (204).

Oxyntomodulin also reduces gastro-intestinal motiliy contributing to the “ileal brake” effect.

Acute central administration of oxyntomodulin inhibits food intake and reduces body weight in

rodents (205). Thus, it can be concluded that central GLP-1R and GLP-2R agonism constitutes a

potential pharmacological tool to reduce food intake and enhance energy expenditure.

1.3.3 Expression of PGDPs and their receptors in the hypothalamus

In the brain, proglucagon neurons are localized in the NTS and the hypothalamus (Figure

2B). The NTS preproglucagon neurons that process proglucagon to GLP-1, GLP-2 and

oxyntomodulin, project mainly to the two hypothalamic nuclei involved in appetite regulation -

the PVN and the DMH (213, 214). GLP-1-immunoreactive (ir) nerve fibres are densely located

in the ARC, PVN and DMH (215), whereas GLP-2-ir nerve fibers are found mainly in the

ventral part of the DMH and also in the ARC and PVN (214). The GLP-1R is widely expressed

in all hypothalamic areas receiving GLP-1-ir fibers i.e. the ARC, PVN and DMH nuclei and also

the VMH and LH (216, 217), but GLP-2R expression in the hypothalamus is much more limited

and confined to the DMH and VMH nuclei (214, 218). Thus, the localization of proglucagon-

neurons as well as GLP-1R- and GLP-2R-expressing neurons in the hypothalamus is very well

defined (Figure 2B); however, it is not yet completely known which PGDPs, whether centrally-

or peripherally-derived PGDPs, activate GLP-1R and GLP-2R in the hypothalamic nuclei that

regulate energy homeostasis. It has been demonstrated that GLP-1 and long-acting GLP-1R

agonist exendin-4, can cross BBB (219, 220), and it is possible that similar to GLP-1, GLP-2 can

also gain access to the brain from the periphery. However, it is a possibility that due to their

rapid inactivation by DPP-4, gut-derived GLP-1 and GLP-2 may not even cross the BBB, and

therefore only centrally-derived GLP-1 and GLP-2 are involved in their action on the

hypothalamus. Also it is yet to be confirmed whether differential distribution of GLP-1R and

GLP-2R in the hypothalamus is responsible for different roles played by central GLP-1 and

GLP-2 in appetite regulation and energy homeostasis.

19

A

B

3V

3V

LH LH

GLP-1R GLP-2R

ARC

VMH

DMH

PVN

LH

VMH

DMH

Proglucagon

Hypothalamus

NTS

Figure 2. A: The products of preproglucagon cleavage. GLP-1 and GLP-2, glucagon-like

peptides 1 and 2; GRPP, glicentin-related pancreatic peptide; IP-1 and IP-2, intervening peptides

1 and 2; MPGF, major proglucagon fragment; PC, prohormone convertase. B: Schematic

diagrams of the central regions showing gene expression of proglucagon and GLP-1

receptor (GLP-1R) and GLP-2R. ARC, arcuate nucleus; AP, area postrema; DMH,

dorsomedial hypothalamic nucleus; LH, lateral hypothalamus; NTS, nucleus tractus solitarius;

Pit, pituitary gland; PVN, paraventricular nucleus; 3V, third cerebral ventricle; VMH,

ventromedial hypothalamic nucleus.

20

1.3.4 Signaling pathways activated by GLP-1R and GLP-2R stimulation

The GLP-1R is expressed in GLP-1 responsive tissues, including pancreatic β-cells, the

gastrointestinal tract and the brain. GLP-1R is activated not only by GLP-1, but also by

oxyntomodulin to mediate its anorexigenic actions (221). GLP-1R activation is inhibited by

exendin-9–39, a known GLP-1R antagonist that blocks actions of GLP-1 and oxyntomodulin.

The signaling pathways triggered by GLP-1R activation in peripheral tissues have been

extensively studied (Figure 3). In the pancreatic β-cells, GLP-1R stimulation by GLP-1 activated

cAMP/PKA, exchange protein activated by cAMP (Epac), PI3K and mitogen-activated protein

kinase (MAPK) pathways (222-227). Oxyntomodulin also mimics many of the actions of GLP-1

on the islet β-cells. Similar to GLP-1, oxyntomodulin, via functional GLP-1R, activates cAMP

formation in murine islets and INS-1 cells (228).

GLP-2R expression is restricted to the gastrointestinal tract and the CNS, with limited

expression in other peripheral organs (229). GLP-2R recognizes specifically GLP-2 and is

inhibited by GLP-2(3-33). Complete information about GLP-2R activation is not available due to

the lack of cell models that endogenously express GLP-2R (230). In heterologous cell lines

transfected with rat or human GLP-2R, cAMP-dependent pathways are activated by GLP-2 (207,

231) (Figure 4). Recently GLP-2 has been shown to activate Akt in the intestinal epithelium

(232) and also the wnt/β-catenin signaling pathway in the intestinal crypt cells (233).

The signaling pathways activated by GLP-1R and GLP-2R stimulation in the CNS,

particularly in the hypothalamus are not well studied. The available majority of the data

highlighting the effects of GLP-1R and GLP-2R stimulation is based on studies performed in cell

lines derived from peripheral tissues. Until now, it is due to the lack of representative neuronal

cell models that the signal transduction pathways activated by GLP-1R and GLP-2R stimulation

in the hypothalamic neurons have not been studied completely. To address this issue, for the first

time, clonal, immortalized, hypothalamic, neuronal cell lines that endogenously express GLP-1R

and GLP-2R will be used as research models (18, 19, 234, 235). Recently, using these cell lines,

it was shown that GLP-1 regulates expression of hypothalamic insulin (234), however, regulation

of other neuropeptides by PGDPs remains largely unknown.

21

Figure 3. Proposed GLP-1R signaling in the hypothalamus. GLP-1R activation in the

hypothalamic neurons may trigger key signal transduction pathways to regulate gene expression

of hypothalamic neuropeptides. In the peripheral tissues, GLP-1R activates cAMP/PKA/

CREB/ATF-1, c-Fos and MAPK/ERK pathways, and other signaling pathways, such as

cAMP/Epac, Wnt/β-catenin and PI3K/Akt can also be activated.

22

Figure 4. Proposed GLP-2R signaling in the hypothalamus. GLP-2R activation in the

hypothalamic neurons may trigger key signal transduction pathways to regulate gene expression

of hypothalamic neuropeptides. In the peripheral tissues, GLP-2R activates cAMP/PKA/CREB

pathway, and other signaling pathways, such as Wnt/β-catenin, c-Fos and PI3K/Akt can also be

activated.

23

1.4 Insulin and insulin receptor signaling

1.4.1 Insulin

The discovery and isolation of insulin at the University of Toronto by Banting, Best,

Collip, and McLeod in the last century was one of the greatest events in the history of medicine.

Since then insulin has become the life-saving therapy for patients suffering from type 1 diabetes

and certain patients with type 2 diabetes (236). Insulin is a polypeptide hormone produced within

the β-cells of the islets of Langerhans. Insulin is generated from a 110-aa single chain

preprohormone, that is initially processed to proinsulin, through removal of 24 aa, then to the

mature 51 aa insulin and C-peptide through enzymatic cleavage by the prohormone convertases

PC1/3 and PC2, and carboxypeptidase E (237). Once mature insulin is synthesized, it is

packaged into granules that are secreted in response to glucose entry into β-cells. Higher glucose

metabolism in the β-cells increases cellular ATP that causes closing of ATP-sensitive potassium

channels, preventing potassium ion outflow, leading to membrane depolarization and calcium

influx via voltage dependent calcium channels (238). The increased cytosolic calcium leads to

exocytosis of the secretory granules, releasing insulin into circulation (239). The liver, skeletal

muscle and adipose tissue express insulin receptor (IR) and are the key peripheral targets of

insulin that mediate its anabolic actions (238).

1.4.2 Insulin receptor signaling

The key actions of insulin in peripheral tissues include increased glucose and amino acid

uptake, fatty acid synthesis, glycogen synthesis; and also decreased lipolysis, proteolysis, and

gluconeogenesis (238). Insulin action is mediated by a membrane receptor that belongs to the

tyrosine kinase receptor superfamily (240, 241). The IR protein is a heterotetrameric protein

composed of two 135 kDa alpha subunits that bind insulin and two 95 kDa beta subunits

containing the transmembrane and tyrosine kinase domains linked by disulfide bonds (242, 243).

Through alternative splicing of the primary transcript, the IR protein has two isoforms known as

the A and B isoforms (240, 241, 244). Isoform A lacks exon 11, which codes for a 12-aa

fragment near the C-terminal end of the α-chain (245), and differential expression of the two

24

isoforms allows for preferential activation of either the IR or insulin-like growth factor-1

receptor pathways (246). Only the IR isoform A is found in the brain (247).

It is well established that IRs are located in the hypothalamus (248, 249), and central

injection of insulin potently reduces food intake and body weight (250). Insulin exerts its central

effects predominantly through the PI3K pathway. The IR remains in an inactive dimerized state

when insulin is not bound (243). Insulin binding to the receptor causes autophosphorylation of

the tyrosine residues that allow docking of the insulin receptor substrate (IRS) family of proteins,

which mediate the IR signaling (251, 252). The IRS proteins are then activated through tyrosine

phosphorylation and can recruit PI3K to the cell membrane for its activation (253). Activation of

PI3K leads to the phosphorylation of membrane bound phosphatidylinositol-3,4-bisphosphate to

generate phosphatidylinositol-3,4,5-triphosphate (PIP3). The binding of PIP3 to the N-terminus

pleckstrin homology domain anchors Akt to the plasma membrane and allows its

phosphorylation and activation by phosphoinositide-dependent kinase-1. Activation of Akt can

then regulate numerous downstream signaling molecules in the periphery as well as in the CNS

(254, 255). Insulin action in the CNS is mediated via modulation of neuropeptide transcription

and release (256). Although IRS1 and IRS2, are often linked to the PI3K-Akt pathway, which is

responsible for the most of the metabolic actions of insulin, the MAPK pathway, works in

conjunction with the PI3K pathway to control cell proliferation, apoptosis and differentiation

(257, 258). Among multiple signalling pathways activated by insulin, the PI3K-Akt pathway

remains the main focus in central nervous system studies. Several investigations have

demonstrated that insulin activates PI3K in neurons (259, 260), and that PI3K inhibitors can

block the ability of insulin to regulate food intake (260).

1.4.3 Insulin regulation of proglucagon

The regulatory mechanisms underlying the differential and opposite actions of insulin on

the pancreatic and intestinal proglucagon gene expression have been extensively investigated

(185). It was demonstrated that insulin inhibited islet proglucagon gene expression via regulation

of gene transcription (200, 201), but another study found that proglucagon gene expression was

insensitive to changes in plasma glucose and insulin concentrations, however, hyperinsulinemia

in the presence of hyperglycemia lowered glucagon secretion from pancreatic α-cells (261).

25

Further, it was found that Akt was sufficient to mimic the inhibitory effect of insulin on

proglucagon expression in the pancreatic α-cells (262). Insulin may inhibit pancreatic

proglucagon by causing exit of FoxO1 from the nucleus to the cytoplasm (197). The repression

of proglucagon is physiologically important because it may result in decreased production of

glucagon, the major counter-regulatory hormone of insulin secreted by islet α-cells. In contrast to

inhibitory effect of insulin on proglucagon expression in pancreatic α-cells (200, 201),

pathological levels of insulin stimulate intestinal proglucagon mRNA and GLP-1 production via

activation of the bipartite transcription factor β-catenin/T cell factor, the major effector of the

canonical Wnt signaling pathway (199). Additionally, in the MKR mice, a non-obese model of

chronic insulin resistance and hyperinsulinemia, intestinal proglucagon mRNA expression and

GLP-1 content were found to be significantly increased (199). A recent finding indicated that

insulin stimulated GLP-1 secretion from the intestinal L-cells in a glucose-dependent manner

(263). Furthermore, similar to the proglucagon gene transcription in pancreatic α-cells, insulin

was demonstrated to utilize the same cis- and trans-regulatory elements in stimulating intestinal

proglucagon expression in the intestinal L-cells, supporting the existence of crosstalk between

insulin and Wnt signaling pathways (199). In the pancreatic α-cells, the repressive effect of

insulin on proglucagon gene expression relies on Akt activity (262), however, in the intestinal L

cells, it is likely that the Wnt signaling pathway, but not the Akt pathway (Akt-independent

mechanism), is involved in insulin-stimulated proglucagon promoter transcription (199).

Recently, it was found that Epac is involved in cAMP-stimulated proglucagon expression and

hormone production in pancreatic and intestinal endocrine cells thus demonstrating the PKA-

independent regulation of proglucagon in the peripheral tissues (264, 265). Currently, no studies

have been conducted to elucidate how insulin regulates hypothalamic proglucagon and the

underlying mechanisms.

1.5 Leptin and leptin receptor signaling

1.5.1 Leptin

Leptin is secreted by adipocytes and was cloned from the ob (leptin) gene in 1994 by

Friedman and his colleagues (17). Leptin is essential to maintain physiological energy

homeostasis. Knockout mice with mutations in the ob gene or db (leptin receptor) gene are

26

morbidly obese. Further, congenital leptin deficiency is associated with severe early-onset

obesity in humans (266). Administration of leptin to leptin deficient ob/ob mice and humans

normalizes body weight and neuroendocrine status (39, 267, 268), indicating the importance of

leptin in energy homeostasis. Leptin plasma concentrations are directly proportional to adiposity

levels and therefore it is known as an adiposity signal (269, 270). Although leptin does not exert

any demonstrable effect on energy metabolism in non-obese humans (271), recent studies

involving leptin administration to humans support its critical role in regulating neuroendocrine

and immune functions as well as insulin resistance in states of energy deficit (272).

The gene for the leptin receptor (Ob-R) is encoded by the db gene and is spliced into

several receptor isoforms. There is one long leptin receptor isoform: Ob-Rb, and several shorter

isoforms spliced at the C-terminal coding region: Ob-Ra, Ob-Rc, Ob-Rd, Ob-Re and Ob-Rf

(273-278). All of the leptin receptor isoforms have an extracellular leptin-binding domain, but

only Ob-Rb has a full-length intracellular domain required for signal transduction (269, 276). It

is well established that leptin receptors are located in the CNS, and within the rodent and human

hypothalamus, Ob-Rb, is highly expressed in the ARC, DMH, VMH, and is also detected in the

hypothalamic PeV, LH and PVN (273-275, 279). In the ARC, Ob-Rb is predominately expressed

in the NPY/AgRP and POMC neurons (43, 273, 280). Leptin gains access to the brain by

crossing the BBB via a saturable transport mechanism (281). It is known that leptin regulates

feeding and energy balance through interaction with complex neuronal circuits comprised of

anorexigenic and orexigenic neuropeptides. Central administration of leptin potently reduces

food intake and body weight in rhesus macaque (282). Lack of functional leptin receptors or

signaling in mice results in severe obesity (283, 284), while brain administration of leptin to

ob/ob mice can overturn the obese phenotype (37, 39, 267). These and other findings indicate

that leptin action within the hypothalamus via NPY/AgRP and POMC neurons is integral to the

maintenance of energy homeostasis (37, 285, 286).

1.5.2 Leptin receptor signaling

Leptin receptors belong to the class 1 cytokine receptor family that utilizes several

cytosolic signaling proteins to induce changes in gene transcription and metabolism (287).

Leptin activation of Ob-Rb predominantly triggers the janus kinase 2 (JAK2)/signal transducer

27

and activator of transcription 3 (STAT3) pathway. JAK2, a tyrosine kinase, is constitutively

associated with Ob-Rb and is rapidly activated upon leptin binding to its receptor. Activation of

JAK2 allows for the recruitment and phosphorylation of STAT3 by binding to tyrosine residue

1138 of Ob-Rb (288-292). Phosphorylated and activated STAT3 then dissociates from the

receptor, homodimerizes in the cytoplasm and ultimately translocates to the nucleus to regulate

gene transcription (293). Apart from activation of JAK2/STAT3 pathway, it has also been

demonstrated that leptin induces PI3K activation in the hypothalamus that is linked to the

anorexigenic action of central leptin, although the mechanisms involved remain unknown (294).

Another molecular target of leptin signaling is the AMP-dependent kinase (AMPK) (295, 296).

AMPK is activated as a result of an increase in the AMP/ATP ratio during low cellular energy

levels. In the hypothalamus, AMPK activation increases food intake and leptin has been shown

to decrease the phosphorylation of hypothalamic AMPK (297). Finally, evidence indicates that

leptin can also activate MAPK-extracellular signal-regulated kinases (ERK) pathway in the

hypothalamus (291, 298-300).

1.5.3 Leptin regulation of proglucagon

The leptin receptor Ob-Rb has been demonstrated to colocalize with neurons expressing

GLP-1-ir (301). Thus, it has been postulated that leptin may exert feeding regulation in part

through activation of GLP-1 neurons, possibly through stimulation of proglucagon gene

expression and production of GLP-1 (302, 303). In support, a recent study concluded that intact

GLP-1 signaling is necessary to mediate the effect of leptin on food intake in rats, indicating that

leptin and GLP-1 act in concert to control the activity of feeding centers (144, 303, 304). A

number of reports suggest that leptin interacts with proglucagon-expressing neurons in mice and

increases hypothalamic GLP-1 content as well as proglucagon mRNA levels in brainstem

neurons (302, 303, 305, 306). Leptin likely regulates the expression of proglucagon mRNA

through the STAT signaling pathway in the NTS neurons (307). Recently, it was found that

leptin downregulates proglucagon mRNA in αTC1-9 cells via STAT3 pathway at 12 h post-

treatment (308). Currently, the regulation of hypothalamic proglucagon neurons by leptin

remains largely unknown.

28

1.6 Hypothalamic neuronal cell models

Over the past two decades, hormones and peptides involved in energy homeostasis and

development of obesity have been investigated through the study of many animal models

including lesion, diet and genetic models (309). However, the function of metabolic hormones,

neuropeptides and neuromodulators in the developing and mature hypothalamus, and their role in

the hypothalamic regulation in energy homeostasis remains far from clear, particularly in

appetite control. The use of primary hypothalamic culture to study hypothalamic functions is

quite challenging and disadvantageous due to their short lifespan and heterogeneous nature as

primary cultures include a mixture of neuronal and glial cell populations, often with a minimal

number of healthy, peptide-secreting neurons. Similarly, studies on the whole brain can be

complicated due to the complex architecture of the brain and the heterogeneity of the

hypothalamic neurons (310). Thus, many questions about the function of hypothalamic

neuropeptidergic neurons in the central nervous system remain unanswered. Immortalized, clonal

hypothalamic cell models may help to elucidate the action of neuropeptides in the hypothalamus

to determine their role in energy homeostasis.

1.6.1 Generation of Immortalized hypothalamic cell lines

The hypothalamic neurons have unique phenotypes and express single or multiple

neuropeptides (310). These neuropeptides are controlled by endogenous or exogenous stimuli. In

the past, bilateral stereotactic or electrical stimulation or lesioning were used to study the

functions of specific hypothalamic neurons; however hypothalamic neuronal subtypes, and

afferent or efferent neuronal extensions could be excessively stimulated or destroyed by these

methods that could potentially lead to erroneous and unreliable outcomes. Due to the multitude

of synaptic inputs received from other adjacent neurons, classical in vivo approach is not useful

to investigate any direct action of an agent on specific hypothalamic neuronal subtypes, or on

neuropeptide gene regulation, synthesis or secretion. More recently, genetically-modified rodent

models with deletion of the hypothalamic neuropeptides or complete ablation of

neuropeptidergic neurons have been used; however, these models are enormously challenging to

generate as it is not quite feasible to knock-down exclusively hypothalamus-specific neurons. In

contrast, cell lines originating from hypothalamic tumours or immortalized from primary

29

hypothalamic cultures, represent a simpler model that lacks the complex network of neuronal

inputs, multitude of connections and signalling, and makes it feasible to investigate those areas

found to be challenging to investigate in vivo.

Immortalized cell lines can be homogeneous, clonal population of specific neuronal

subtypes and can be maintained in a regulated environment with fewer unregulated variables

present in animal models that may interfere with the direct action of the chemicals under

investigation. Cell models can therefore be used to investigate regulation of specific genes or

proteins, which is challenging to perform in vivo due to the numerous neuronal phenotypes

present in a given hypothalamic area. To circumvent challenging issues with the contemporary

experimental models, scientists have tried to generate immortalized cell models with moderate

success (311-313). Using the retroviral transfer of simian virus (SV) 40-T-Antigen (18, 313), our

laboratory has successfully generated an array of immortalized cell models from the mouse

hypothalamus (18, 313). Many of these clonal cell models express specific neuropeptides and

receptors, and have been extensively used in biomedical research [reviewed in (235, 314-316)].

1.6.2 Immortalized embryonic hypothalamic cell lines, mHypoE-XX

Historically, there have been unsuccessful attempts to generate hypothalamic cell lines

(317, 318). The lack of representative neuronal cell models encouraged our research laboratory

to generate embryonic, clonal hypothalamic mouse cell lines (18). We utilized the retroviral

SV40-T-Antigen transfer technique to immortalize primary hypothalamic cultures obtained from

fetal mice on E15, E17 and E18 and generated a heterogeneous mixed population of neurons.

These mixed neuronal cultures were further subcloned to obtain several single, homogeneous,

clonal cell populations. These cell models are designated as mHypoE-‘clone number’ (mouse,

hypothalamic, embryonic – ‘clone number’) to distinguish them from other newly created cell

lines and to avoid any confusion. Each cell line, out of over 60 cell models, exhibits a distinct

neuronal phenotype, although all of the cell models commonly express mature neuronal markers

such as neuron-specific enolase and neurofilament, but not neuroglia-specific glial fibrillary

acidic protein. The cells also possess markers of the neurosecretory machinery, such as syntaxin,

a large number of neurosecretory granules and exhibit an intracellular calcium response after

depolarization by potassium chloride (18). These cell lines endogenously express hormone

30

receptors and neuropeptides involved in many hypothalamic functions. These cell lines have

been used as representative neuronal models to dissect complex cellular signaling mechanisms

involved in hypothalamic control of physiological processes [reviewed in (235, 314-316) ].

1.6.3 Immortalized adult hypothalamic cell lines, mHypoA-XX

The embryonic hypothalamic models provide a useful tool in understanding cellular

biology of specific neuroendocrine neurons, however, it is imperative to known how the

embryonic neurons differ from the adult neurons in their functions under normal and

pathological circumstances. Historically, immortalization of adult hypothalamic neurons

appeared to be impossible due to their fully differentiated, non-proliferating nature. Thus, only

proliferating cells such as embryonic neurons could be immortalized using SV40-T-antigen

transfer (312, 313). Recent findings indicate that neurogenesis occurs throughout the

hypothalamus at low levels and is not restricted to a specific neuronal phenotype or nuclei in the

CNS (319-324). A significant finding by Kokoeva et al. that ciliary neurotrophic factor (CNTF)

induces de novo neurogenesis in the hypothalamus enabled us to devise a novel method to

immortalize primary hypothalamic culture (19). Our laboratory used CNTF to induce adult

neuronal proliferation and successfully immortalized adult mouse hypothalamic neurons (235,

314, 315, 319). So far, we have generated over 50 cell lines from adult mice and labelled them as

mHypoA-‘clone number’ (mouse, hypothalamic, adult-‘clone number’).

1.7 Rationale, Hypotheses and Specific Aims

1.7.1 Rationale

At present, the mechanisms underlying proglucagon gene expression in pancreas and

intestine have been extensively studied (185), however the regulation of the hypothalamic

proglucagon by insulin or leptin remains unstudied (Figure 5). It has been demonstrated that in

the periphery as well as in the CNS, insulin mainly acts via classic PI3K/Akt pathway and leptin

activates classic JAK2/STAT3 pathway. Currently, it remains unknown, whether both hormones

utilize the same or different signaling pathways to regulate proglucagon mRNA levels in the

hypothalamic neurons. Therefore, in the first part of the present thesis, we investigated whether

31

hypothalamic proglucagon mRNA transcript expression is regulated by insulin in an Akt-

dependent manner, and by leptin via activation of JAK2/STAT3 pathway.

The exact mechanisms of action of GLP-1 and GLP-2 in the hypothalamus are not

completely clear; particularly regulation of hypothalamic neuropeptides by GLP-1R and GLP-2R

activation needs further attention (Figure 6). Even more importantly, this area needs to be

thoroughly investigated as drugs preventing GLP-1 inactivation and agonists of GLP-1R or GLP-

2R are already approved for clinical use or undergoing clinical trials for metabolic or

inflammatory disorders (325, 326). It is known that GLP-1R and GLP-2R agonists activate

classic cAMP/PKA pathway among several other pathways in the periphery and in the brainstem

neurons (Figures 3 and 4) (327), however, the signaling pathways activated by GLP-1R and

GLP-2R agonists in the hypothalamic neuropeptidergic neurons remain unknown. Therefore, in

the second part of this thesis, we investigated whether hypothalamic neuropeptides are regulated

by GLP-1R and GLP-2R agonists via activation of cAMP/PKA pathway. We focused on the

regulation of hypothalamic neurotensin and ghrelin, because, in contrast to the extensively

investigated NPY- and POMC-mediated regulation of energy homeostasis, these neuropeptides

have received relatively little attention, despite their potential role in the hypothalamic regulation

of feeding and energy homeostasis (6, 139, 141-143, 153, 155).

Many questions about the regulation and action of GLPs in the CNS remain unanswered

due to the lack of representative neuronal experimental models. Recently generated embryonic

and adult immortalized, clonal, hypothalamic mouse cell lines, endogenously expressing

proglucagon and functional GLP-1R and GLP-2R, will serve as relevant neuronal models to

investigate regulation and potential action of GLPs in the hypothalamus. Initially, using these

novel murine hypothalamic cell models, insulin- and leptin-mediated regulation of the

hypothalamic proglucagon-expressing neurons was studied. Subsequently, the effects of

exendin-4 and h(Gly2)GLP-2[1-33] on the hypothalamic neuropeptides that have remained

elusive until now were investigated. Dissecting the mechanisms underlying proglucagon gene

regulation and the effects of GLP-1R and GLP-2R activation on hypothalamic neuropeptides,

such as neurotensin and ghrelin, will lead to a better understanding of the role of proglucagon

gene and GLPs in hypothalamic control of many vital functions, including appetite regulation

32

and energy homeostasis. In the long run, these findings should facilitate the development of safe

and effective drugs for the treatment of metabolic or inflammatory disorders, including obesity.

1.7.2 Hypotheses

The first hypothesis is that hypothalamic proglucagon mRNA transcript expression is

upregulated by insulin in an Akt-dependent manner, and by leptin via activation of JAK2/STAT3

pathway in the hypothalamic proglucagon-expressing neuronal cell models (Figure 5).

The second hypothesis is that long-acting GLP-1R agonist exendin-4 and GLP-2R

agonist h(Gly2)GLP-2[1-33] upregulate hypothalamic anorexigenic neurotensin and

downregulate orexigenic ghrelin mRNA expression in a PKA-dependent manner in the

hypothalamic GLP-1R and -2R-expressing neuronal cell models (Figure 6).

1.7.3 Specific aims

The hypotheses were examined in three aims:

Aim 1: To study the regulation of proglucagon mRNA expression by insulin and leptin using

hypothalamic proglucagon-, IR- and ObRb-expressing neuronal cell models.

Western blot analysis was performed to study activation of signal-transduction pathways,

real-time qRT-PCR to detect changes in mRNA expression, and transfection of proglucagon

promoter-reporter constructs to determine regulation of gene transcription.

Aim 2: To determine potential neuropeptidergic neurons activated by exendin-4 in the mouse

hypothalamus [by immunohistochemistry (IHC)] and to investigate downstream signal

transduction pathways activated by exendin-4 to regulate neurotensin and ghrelin mRNA

expression in the hypothalamic GLP-1R-expressing neuronal cell models (by Western blot and

real-time qRT-PCR analysis).

Aim 3: To study the actions of h(Gly2)GLP-2 on potential downstream neuropeptidergic neurons

in vivo using mouse model (by IHC) and investigate downstream signal transduction pathways

involved in regulation of neurotensin and ghrelin mRNA in the hypothalamic GLP-2R-

expressing neuronal cell model (by Western blot and real-time qRT-PCR analysis).

33

Figure 5. Schematic of the proposed regulation of proglucagon-expressing (GLU) neurons

by insulin and leptin in the hypothalamus. The first hypothesis is that insulin and leptin

directly, via activation of specific signaling pathways, regulate proglucagon-expressing neurons

located in the hypothalamic nuclei involved in energy homeostasis. This regulation itself, as well

as further activation of second-order neurons via GLP-1 or GLP-2 action may lead to a decrease

in food intake and an increase in energy expenditure.

34

Figure 6. Summary of the proposed signal transduction mechanisms triggered by GLP-1R

and GLP-2R activation in the hypothalamus. The second hypothesis is that acute GLP-1R and

GLP-2R activation directly regulates NPY-, POMC-, neurotensin- and ghrelin-expressing

neurons located in the hypothalamic nuclei regulating energy homeostasis. This regulation itself,

as well as further activation of second-order neurons leads to a decrease in food intake and an

increase in energy expenditure.

35

Chapter 2

Materials and Methods

36

2 Materials and Methods

2.1 Cell culture and reagents

Immortalized neuronal cells (mHypoA-2/10, mHypoA-2/30, mHypoE-36/1, and

mHypoE-39) were grown in monolayer in Dulbecco’s modified Eagle medium (DMEM, Sigma,

Canada), supplemented with 4.5 mg/ml glucose, 5% fetal bovine serum (Hyclone Laboratories,

USA), and 1% penicillin/streptomycin (Gibco, Canada), and maintained at 37°C in an

atmosphere of 5% CO2 (18). For the mRNA expression study, cell culture medium was replaced

with DMEM containing 0.5% fetal bovine serum (FBS) and 1% penicillin/streptomycin for a

minimum of 12 h prior to treatments. For the protein expression study, cell culture medium was

replaced with FBS-free DMEM containing 1% penicillin/streptomycin for a minimum of 12 h

prior to treatments. Insulin was a generous gift from Novo Nordisk Canada Inc. (Mississauga,

ON, Canada). Recombinant murine leptin was obtained from Dr. A. F. Parlow, National

Hormone and Peptide Program, Torrance, CA. Exendin-4, GLP-2(1-33) and h(Gly2)GLP-2 were

purchased from American Peptide, USA. 3-isobutyl -1-methylxanthine (IBMX), actinomycin D,

wortmannin, LY 294002 hydrochloride, cucurbitacin I and SD 1008 were obtained from Tocris

Bioscience, USA. Forskolin and 5, 6-Dichlorobenzimidazole 1-β-D-ribofuranoside (DRB), an

inhibitor of transcription, were purchased from Sigma-Aldrich, Canada. The protein kinase A

(PKA) inhibitor H89 was the product of Calbiochem (EMD Biosciences), San Diego, CA or

Tocris Bioscience, USA. Expression vectors pGL2 and pRL-CMV were purchased from

Promega (Madison, WI). The G protein beta (Gβ) antibody was purchased from Santa Cruz

Biotechnology (Santa Cruz, CA). Antibodies against total Akt, phospho-Akt, total STAT3,

phospho-STAT3, total CREB, phospho CREB/ATF-1, c-Fos (for Western blot analysis) and

secondary antibodies were obtained from Cell Signaling Technology Inc. (Danvers, MA, USA).

2.2 One-step reverse transcriptase-polymerase chain reaction (RT-

PCR)

One-step RT-PCR was carried out to screen the cells for proglucagon, ghrelin,

neurotensin, NPY, POMC, IR, ObRb, GLP-1R and GLP-2R mRNA transcripts using One-Step

37

RT-PCR Kit (Qiagen, Mississauga, ON). The primer pairs used for the RT-PCR were designed

using mouse sequences in GenBank to span introns to minimize the possibility of genomic DNA

contamination. The primer pairs used are as follows: IR-F: 5’-gtgataccagagcataggag-3’, IR-R:

5’-ctgttcggaacctgatgac-3’; Ob-Rb-F: 5’-atgacgcagtgtactgctg-3’, Ob-Rb-R: 5’-

gtggcgagtcaagtgaacct-3’, NPY-F: 5’-taggtaacaagcgaatgggg-3’, NPY-R: 5’-acatggaagggtcttcaagc-

3’, -actin-F: 5’-gctccggcatgtgca-3’, -actin-R: 5’-aggatcttcatgaggtagt-3’, Proglucagon-F: 5’-

tgaagaccatttactttgtggct-3’, Proglucagon-R: 5’-tggtggcaagattgtccagaat-3’ (328), Ghrelin-F: 5’-

agcatgctctggatggacatg-3’, Ghrelin-R: 5’-aggcctgtccgtggttacttgt-3’, Neurotensin-F: 5’-

ataggaatgaaccttcagctg-3’, Neurotensin-R: 5’-gtaggaggccctcttgagtat-3’, POMC-F: 5’-

atgccgagattctgctacagtcg-3’, POMC-R: 5’-ttcatctccgttgccaggaaacac-3’, GLP-1R-F: 5’-

tttgatgactatgcctgctgg-3’, GLP-1R-R: 5’-agcccatcccactggtgtt-3’, GLP-2R-S: 5’-

tgctggtttccatcaagcaa-3’, GLP-2R-AS: 5’- atcagctgcaaggtggacaa-3’, Histone-F: 5’-

gcaagagtgcgccctctactg-3’, Histone-R: 5’-ggcctcacttgcctcgtgcaa-3’.

Total RNA was isolated from the immortalized hypothalamic cells, 3T3, α-TC and

GluTag cells using the guanidinium thiocyanate phenol chloroform extraction method (329). One

step RT-PCR was performed using the one step RT-PCR kit. Briefly, total 200 ng of RNA was

used from all samples with 1X one step RT-PCR buffer, one step enzyme mix, 0.4 mM dNTPs,

0.6 mM of sense primer and 0.6 mM of anti-sense primer in a final volume of 25 μL. The RT

protocol used for all genes was 50°C for 30 min, 95°C for 15 min, followed by amplification at

95°C for 15 s, 60°C for 15 s and 72°C for 1 min for 40 cycles with final incubation for 7 min at

72°C. All PCR-amplified products were visualized on 2% agarose gel containing ethidium

bromide under ultraviolet light with the Kodak IS2000 digital imaging centre. The PCR

amplicons were verified by purification and sequencing (The Centre for Applied Genomics,

Toronto, Canada).

2.3 Radioimmunoassay for cAMP analysis

The mHypoA-2/30 and mHypoE-36/1 cells were split into 24-well plates until 80% to

90% confluent. After overnight incubation in serum-free DMEM, cells were washed twice with

1X phosphate buffered saline (PBS) and pretreated for 5 minutes with vehicle, 1 µM exendin-9-

39 (a GLP-1R antagonist) or GLP-2 (3-33) (a GLP-2R antagonist) alone prior to a 10-minute

38

treatment with vehicle, forskolin (1 or 10 µM), exendin-4 (10 or 50 nM), GLP-2 (10 or 50 nM)

or h(Gly2)GLP-2 (10 nM) and incubated at 37 °C.

All drugs including vehicle were diluted in OptiMEM (Invitrogen Life Technologies,

ON) containing 100 mM IBMX, a competitive nonselective phosphodiesterase inhibitor that

promotes accumulation of intracellular cAMP. Forskolin was used as a positive control. After

incubation at 37 °C, 1 ml anhydrous ethanol at -20°C (final concentration, 77% ethanol) was

added to the cells to terminate the reaction. The plates were allowed to sit at -20°C for 24 hours.

The extracts were collected, centrifuged and the supernatant was analyzed for cAMP content by

using cAMP-RIA kit (Biomedical Technologies Inc., USA). cAMP levels were calculated

relative to total amount of protein, as determined by the bicinchoninic acid (BCA) protein assay

kit (Thermo Scientific, IL, USA).

2.4 Quantitative reverse transcription-polymerase chain reaction

(qRT-PCR)

The mHypoA-2/10 and mHypoE-39 cells were treated with vehicle, insulin (10 nM) or

leptin (10 nM), and harvested at the indicated time points. The mHypoA-2/30 and mHypoE-36/1

cells were harvested at the indicated time points following treatment with vehicle, exendin-4 (10

nM) or h(Gly2)GLP-2 (10 nM) over 24 h time period. For the regulation of gene expression

study, long-acting GLP-2, h(Gly2)GLP-2, was used. The EC50 of h(Gly

2)GLP-2 for mouse GLP-

2R is identical to that of rat GLP-2 for both the rat GLP-2R and hGLP-2R, suggesting a high

degree of similarity between their biological actions in different species (328). For the

experiments using inhibitors, the cells were pre-treated with either 25 µM LY294002, 1 µM

wortmannin, 5 µM cucurbitacin I, 10 µM SD1008, 1 μM PKI 14-22 amide, 5 μM H89, 10 µg/ml

actinomycin D, 60 µM DRB or dimethyl sulfoxide (DMSO) vehicle for 45 min to 1 h, and then

treated with either insulin (10 nM), leptin (10 nM), exendin-4 (10 nM) or h(Gly2)GLP-2 (10

nM) prior to total RNA isolation at the indicated time points.

Total RNA from the cells was isolated using guanidinium thiocyanate-phenol-chloroform

extraction method (329), and real-time qRT-PCR was performed with 2.0 µg of total RNA using

using SuperScript II and random primer as described in the Superscript II cDNA Synthesis Kit

39

(Invitrogen). Real-time qRT-PCR was performed with SYBR green PCR master mix according

to the manufacturer’s instructions (Applied Biosystems Inc., Streetsville, Ontario, Canada), and

run on the Applied Biosystems Prism 7000 real-time PCR machine (19). Total 200 ng of

template was used in total 10 µl reaction mixture containing 0.3X SYBR green dye, 1X ROX,

1X buffer, 3 mM MgCl2, 0.2 mM dNTP, and 0.5U Platinum Taq (all from Invitrogen Life

Technologies, Burlington, ON). The real-time qRT-PCR conditions for all genes were as

follows: 45 cycles, 15 sec at 95°C and 1 min at 65°C.

The gene primer sequences were purchased from Integrated DNA Technologies

(Coralville, IA) and are as follows: -actin-SYBR-F: 5'-cttccccacgccatcttg-3', -actin-SYBR-R:

5'-cccgttcagtcaggatcttcat-3', proglucagon-SYBR-F: 5’-gaggagaaccccagatcattcc-3’, proglucagon-

SYBR-R: 5’-gtggcgtttgtcttcattcatc-3’, ghrelin-SYBR-F: 5’-ggaggagctggagatcaggtt-3’, ghrelin-

SYBR-R: 5’-ggcccggccatgctgct-3’ (Primer Express software, Applied Biosystems). TaqMan

Gene Expression assay containing neurotensin-specific primers and probe (Applied Biosystems)

was used to determine neurotensin mRNA expression levels.

Data were represented as Ct values, defined as the threshold cycle of PCR at which

amplified product was first detected, and analyzed using ABI Prism 7000 SDS software package

(Applied Biosystems). Samples were run in triplicate and average Ct was analyzed. Copy

number of amplified proglucagon, ghrelin or neurotensin gene was standardized to -actin using

the standard curve method (ABI Prism 7000 Users Bulletin). The final fold differences in

mRNA expression were relative to the corresponding time-matched control.

2.5 SDS-polyacrylamide gel electrophoresis and western blot analysis

The hypothalamic cells were grown to 80-90% confluency and serum-starved overnight.

The mHypoA-2/10 and mHypoE-39 cells were treated with either vehicle, insulin (10 nM),

leptin (10 nM) over a 60 minute time course. The mHypoA-2/30 and mHypoE-36/1 cells were

treated with either vehicle, exendin-4 (10 nM) or h(Gly2)GLP-2 (10 nM) over a 6 h time course.

The cells were harvested at the indicated time points, and total protein was isolated using a 1X

lysis buffer [20 mM Tris-HCl (pH 7.5), 150 mM NaCl, 1 mM Na2EDTA, 1 mM EGTA, 1%

Triton, 2.5 mM sodium pyrophosphate, 1 mM β-glycerophosphate, 1 mM Na3VO4, 1 μg/ml

40

leupeptin] supplemented with 1mM PMSF as described previously (141). Protein concentration

was determined using the BCA protein assay kit (Thermo Scientific, IL, USA). Total protein (20

µg) was resolved on 8% SDS-PAGE gels and blotted onto Immun-Blot polyvinyl difluoride

membrane (Bio-Rad, CA, USA). The blots were blocked with 5% bovine serum albumin in tris-

buffered saline with 0.1% Tween-20 for 1 h, and then incubated overnight at 4ºC with rabbit

polyclonal anti-mouse primary antibodies against phospho-Akt (Ser473, 1:1000), phospho-

STAT3 (Tyr705, 1:1000), total Akt (1:1000), total STAT3 (1:1000), Phospho-CREB (Ser133,

1:500), total CREB (1:500), c-Fos (1:500) and Gβ (1:1000). Following incubation with

horseradish peroxidase-labeled secondary goat anti-rabbit IgG at a 1:5000 dilution for 1 h,

enhanced chemiluminescence (ECL Advance kit; GE Healthcare, USA) was added and the

immunoreactive bands were visualized using Kodak IS2000 digital imaging system. The

intensity of bands was quantified by densitometry analysis using Kodak 1D Image Analysis

Software 3.6 (Eastman Kodak Company, Rochester, NY, USA). For the inhibitor experiments,

the mHypoA-2/10 and mHypoE-39 cells were pre-treated with either 25 µM LY294002, 1 µM

wortmannin, 5 µM cucurbitacin I, 10 µM SD1008, or DMSO for 1 h, and then treated with 10

nM insulin or leptin for 15 min.

2.6 Reporter gene plasmids

Reporter constructs based on the pGL2 vector were constructed by transferring the inserts

from the reporter gene plasmids based on the pGL3 vector containing 312 and 476 bases of the

rat proglucagon promoter and human proglucagon promoter constructs containing 332 and 602

bases that have previously been described (330). Human proglucagon promoter construct

containing 829 bases of promoter was generated using a PCR based strategy as previously

described (330), for which the following human proglucagon gene primers were used (the

numbers in the names of primers identify the location of the 5' end of each primer relative to the

mRNA start site): -829F: 5’-gcgggtaccgcctgtgtgtccagtcacaaaac-3’, and +58R: 5’-

gcaagcttagagcaagccctctttgggaac-3’.

41

2.7 Transient transfections

The mouse hypothalamic mHypoE-39 and mhypoE-20/2 cell lines were grown in 24-well

plates in DMEM without antibiotics and supplemented with 4.5 mg/ml glucose and 5% fetal

bovine serum to 70-80% confluency. Prior to the transfection, the medium was changed to

medium with reduced serum (OptiMEM). The luciferase reporter gene constructs were

transfected into the cell lines using Lipofectamine 2000 (Invitrogen) as per the manufacturer’s

protocol, with 0.8 µg of DNA per well in a 24-well plate. All luciferase assays were performed

in triplicate. Using three different passages of the cell lines, at least three separate transfections

per plasmid were performed. For each transfection experiment, individual plasmids were

transfected in three separate wells. The cells were incubated for 4-6 h with DNA-Lipofectamine

2000 complex, and then washed and incubated in fresh media for 18 h before being treated with

vehicle, insulin (10 nM), leptin (10 nM), or forskolin/IBMX (10 µM). The cells were harvested

at 12 h post-treatment, and luciferase assay was performed to detect luciferase activity using

Firefly & Renilla Luciferase Assay Kit (Biotum, Inc., USA) and a Lumat LB 9501 luminometer

(EG&G Berthold, wellesley, MA). Protein concentrations were determined using the BCA

Protein Assay Kit (Thermo Scientific, Rockford, IL). Controls for transfections included the

promoterless pGL2 basic vector (Promega), and the internal control reporter pRL-CMV vector

(containing enhancer to express Renilla luciferase; Promega).

2.8 Animal (mouse) experiments

Adult male C57BL/6 (8-10 wk old, 25-30 g) mice were housed in individual

polycarbonate cages in a temperature-controlled vivarium on a 12:12 hour light:dark cycle. All

animals were given ad libitum access to standard rodent chow and water at all times. Using flat-

skull coordinates from bregma (antero-posterior -0.825 mm, medio-lateral 0 mm, dorso-ventral -

4.8 mm) a 26-gauge stainless steel guide cannula (Plastics One, Roanoake, VA) projecting into

the third cerebral ventricle was implanted into each mouse. Correct cannula placement and the

patency were verified 1 wk after surgery by intracerebroventricular (i.c.v.) injection of 250 ng of

angiotensin II in 0.5 μl 0.9% saline (19). Angiotensin II induces a drinking a response and only

mice with positive drinking responses were included in subsequent analyses. After allowing the

mice to recover for 7-14 days, they were handled daily for at least one week before beginning the

42

experiments. All procedures and manipulations were conducted in accordance with the protocols

and guidelines approved by the University of Toronto Animal Care Committee.

2.9 Intracerebroventricular microinjections for feeding study

Exendin-4 was obtained from American Peptide, USA. For the animal treatment,

exendin-4 was dissolved in 0.9% saline on the day of treatment and microinjections were

administered one hour prior to onset of the dark phase. The dose of exendin-4 used to induce

anorexia was based on a previous report (221). Each ad libitum-fed mouse received either

exendin-4 or 0.9% saline in a total volume of 2 μl by slow infusion over 10-15 minutes into the

third ventricle using a 30-gauge injector attached by polyethylene tubing to a 2 μl glass syringe

(Hamilton, Reno, NV). The mice were returned to their home cages with free access to a

premeasured amount of chow and water, and the effect of i.c.v. exendin-4 on feeding was

determined (4 mice per treatment group). Changes in food and water intake were measured at 2,

4, 18 and 24 h post-treatment, and change in animal body weight was recorded at 24 h post-

treatment.

The degradation-resistant human analog of GLP-2, h(Gly2)GLP-2, was purchased from

American Peptide Company, USA. For the animal treatment, h(Gly2)GLP-2 was dissolved in

0.9% saline on the day of treatment, and microinjections were administered 1 h prior to onset of

the dark phase. In order to determine the effective dose of h(Gly2)GLP-2 to study mouse

hypothalamic neuronal activation, a peptide dose-response feeding study using a range of doses

of h(Gly2)GLP-2 (100 ng, 1µg, 5 µg per mouse) was conducted. Each ad libitum-fed mouse

received either different doses of h(Gly2)GLP-2 or 0.9% saline in a total volume of 2 μl by slow

infusion into the third ventricle (4 mice per treatment group). Upon returning to their home

cages, the mice were allowed to freely access premeasured amounts of chow and water. The

effect of i.c.v. h(Gly2)GLP-2 on feeding was determined by measuring changes in food and

water intake and animal weight at 1 and 2 h post-treatment.

43

2.10 Assessment of neuronal activation by c-Fos immunohisto-

chemistry

To study the effect of acute stimulation of central GLP-1R by exendin-4 and GLP-2R by

h(Gly2)GLP-2 on the activation of hypothalamic neuropeptidergic neurons, ad libitum fed mice

(4 mice per treatment group) were treated with either i.c.v. exendin-4, h(Gly2)GLP-2 or 0.9%

saline as described above for the feeding study. 2h following i.c.v. injections, mice were

anesthetized using CO2 and perfused transcardially with ice-cold 0.1 M PBS followed by freshly

prepared 4% paraformaldehyde solution. Brains were removed immediately, post-fixed in 4%

paraformaldehyde, serially cryoprotected in 15% and 30% sucrose solution, snap-frozen in a pre-

chilled dry ice-isopentane bath, and stored at -80°C. For subsequent immunohistochemical

analysis, frozen brains were cut with a cryostat (Leica CM1510S, Leica Microsystems, Canada)

in a rostral to caudal direction in the coronal plane into 20 μm sections, and serial sections were

stored at -20°C in a cryoprotectant solution (19).

The number of c-Fos-, α-MSH/POMC-, NPY-, neurotensin-, or ghrelin-ir neurons in

specific hypothalamic regions were quantitatively assessed. For the immunohistological analysis,

every other section at 20 μm intervals through the hypothalamus was selected throughout the

hypothalamic nuclei. The sections were allocated rostral to caudal to visualize the distribution of

neuropeptide- or c-Fos-ir neurons on each hemisphere throughout these nuclei. Evenly spaced

sections covering the region -0.70 mm to -1.94 mm from bregma were defined according to the

Mouse Brain Atlas of Paxinos and Franklin (331).

The specific primary rabbit anti-mouse antibodies and their concentration used for

detection of immunoreactivity are as follows: anti-c-Fos (1:25,000; Calbiochem, Canada), anti-

neurotensin (1:5000; Immunostar, USA), anti-ghrelin (1:1000; Phoenix Pharmaceuticals, Catalog

No. H-031-31; anti-ghrelin antibody to n-octanoyl ghrelin), anti-α-MSH (1:200; Phoenix

Pharmaceuticals, Catalog No. H-043-01), and anti-NPY (1:1000; Phoenix Pharmaceuticals,

Catalog No. H-049-03). The detection of c-Fos, α-MSH, NPY, neurotensin or ghrelin

immunoreactivity was performed by conventional avidin-biotin-immunoperoxidase method

using biotinylated goat anti-rabbit IgG secondary antibody and diaminobenzidine (DAB) as

chromogen (Vectastain ABC Elite Kit; Vector Laboratories, Canada) (301). The brain sections

44

were processed for immunohistochemical staining by Tyramide Signal Amplification method

(TSA; PerkinElmer). The IHC for c-Fos was performed with DAB to yield a brown nuclear

reaction product, whereas the IHC for α-MSH/POMC, NPY, neurotensin and ghrelin was

performed using DAB with Metal Enhancer (cobalt chloride), yielding a more intense blue/black

cytoplasmic reaction product (Sigma, Catalog No. D0426).

Immunostained sections were examined under a Zeiss Axioplan 2 microscope outfitted

with an AxioCam HRc camera and AXIOVISION 4.2 imaging software. For the quantification

of cells, every other section throughout the ARC, PVN, DMH and VMH was taken to visualize

the distribution of α-MSH/POMC, NPY, neurotensin or ghrelin neurons throughout these nuclei

(total 3-4 sections/mouse). For each of the ARC, PVN, DMH and VMH from both sides, an

image was captured in a single plane of focus at X40 magnification and a 0.2 mm2 box was

placed in the center of the selected hypothalamic regions. Cells with brown nuclear staining were

considered c-Fos-ir, whereas cells with dark blue/black cytoplasmic staining were considered α-

MSH/POMC-, NPY-, neurotensin- or ghrelin-positive. The immunoreactive neuronal cells from

both hemispheres of 4 to 6 sections per each animal were counted in a blind manner, and in each

group the mean value of the cell counts per section of an individual animal was used for

statistical analysis. The results were expressed as the ratio of cells co-expressing c-Fos with

either α-MSH/POMC, NPY, neurotensin or ghrelin to the total number of respective

neuropeptide-ir cells per 0.2 mm2 area of the ARC, PVN, DMH, VMH, PeV or the internuclear

regions.

2.11 Experimental normalization

For the analysis of the protein expression, phospho-Akt expression was normalized to

total Akt, phospho-STAT3 expression was normalized to total STAT3, and phospho-CREB and

phospho-ATF-1 expression was normalized to total CREB (300); only c-Fos was normalized to

Gβ. Primary antibody against Gβ was used as an internal loading control for all protein

expression assays. For the real time and semi-quantitative qRT-PCR experiments, the sample

mRNA expression values were divided by the expression values of the housekeeping gene -

actin to obtain sample/housekeeping ratio. These values represented relative mRNA expression

of the sample gene and were utilized for statistical analysis.

45

2.12 Statistical analysis

Data are presented as the mean ± the standard error of the mean (SEM) from at least three

independent experiments. Data were analyzed using GraphPad Prism software (GraphPad

Software, Inc., USA) or SigmaPlot (Systat Software Inc., USA). Statistical analysis was

performed using one-way or two-way analysis of variance (ANOVA), and statistical significance

was determined by post hoc analysis using Bonferroni test or Student’s t-test with P < 0.05.

46

Chapter 3

Regulation of the Proglucagon mRNA Levels by

Insulin and Leptin in Embryonic versus Adult

Hypothalamic Neurons

47

Publication:

Prasad S. Dalvi, Frederick D. Erbiceanu, David M. Irwin, and Denise D. Belsham. Direct

Regulation of the Proglucagon Gene by Insulin, Leptin and cAMP in Embryonic versus Adult

Hypothalamic Neurons. Molecular Endocrinology (2012) 26:1339-1355

F.D.E. was a graduate student under the supervision of D.M.I. and performed the study on the

cAMP regulation of proglucagon gene expression included in this manuscript for his M.Sc.

thesis research. All other experiments included in the manuscript and presented in this thesis

were designed by P.S.D. and D.D.B. P.S.D. executed all the experiments included in this thesis,

analyzed all data, designed and created all figures, wrote and revised the manuscript under the

supervision of D.D.B. D.M.I. provided the human and rat proglucagon promoters and revised

the manuscript.

Published figures:

Figure 7. Characterization of the expression profile of the proglucagon-expressing hypothalamic cell

lines.

Figure 8. Insulin activates Akt and leptin activates STAT3 in the hypothalamic neuronal cells.

Figure 9. Insulin and leptin regulate proglucagon mRNA expression in the hypothalamic neuronal cells.

Figure 10. Regulation of proglucagon mRNA expression by insulin via activation of the PI3K/Akt

pathway.

Figure 11. Regulation of proglucagon mRNA expression by leptin via activation of the JAK2/STAT3

pathway.

Figure 12. Insulin and leptin do not affect the transcription of proglucagon promoter constructs, but

regulate mRNA stability.

Figure 13. In silico analysis of murine proglucagon mRNA sequence for miRNA binding sites and

RNA-binding protein sites.

Permissions were obtained to reproduce the copyrighted material.

48

3 Regulation of the Proglucagon mRNA Levels by Insulin and

Leptin in Embryonic versus Adult Hypothalamic Neurons

3.1 Abstract

The proglucagon gene is expressed not only in the pancreas and intestine, but also in the

hypothalamus. PGDPs have emerged as potential regulators of energy homeostasis. Whether

insulin or leptin activation controls proglucagon gene expression in the hypothalamus is not

known. A key reason for this has been the inaccessibility of hypothalamic proglucagon-

expressing neurons and the lack of suitable neuronal cell lines. Herein, the mechanisms involved

in the direct regulation of the proglucagon gene by insulin and leptin in hypothalamic cell

models are described. Insulin, through an Akt-dependent manner, significantly increased

proglucagon mRNA levels by 70% in adult-derived mHypoA-2/10 neurons, while significantly

suppressing them by 45% in embryonic-derived mHypoE-39 neurons. Leptin, via the

JAK2/STAT3 pathway, caused an initial increase by 66% and 43% at 1 h followed by a decrease

by 45% and 34% at 12 h in mHypoA-2/10 and mHypoE-39 cells, respectively. Transient

transfection analysis determined that human or rat proglucagon 5' flanking promoter regions

were not regulated by insulin and leptin in the mouse embryonic cell lines, whereas RNA

stability assay demonstrated that insulin and leptin increased proglucagon mRNA stability in the

mouse adult cell line at 4 h and 1 h post-treatment. These findings suggest that insulin and leptin

act directly, on specific hypothalamic neurons to regulate proglucagon mRNA expression. Since

PGDPs are potential regulators of energy homeostasis, an understanding of regulation of

hypothalamic proglucagon neurons is important to further expand our knowledge of alternative

feeding circuits.

3.2 Introduction

Among numerous appetite-regulating neuropeptides, the PGDPs, including glucagon,

GLP-1, GLP-2, and oxyntomodulin have emerged as potential regulators of feeding behavior (6).

Proglucagon is encoded by a single proglucagon gene and is expressed in pancreatic α-cells,

intestinal endocrine L-cells, brain stem neurons and the hypothalamus (172). PGDPs are

49

synthesized by post-translational processing of precursor proglucagon in cell-specific manner by

PC1/3 and PC2 (174). GLP-1, GLP-2, glicentin, and oxyntomodulin are synthesized in the

intestinal endocrine L-cells which are located mainly in the distal ileum and colon, whereas

glucagon is predominantly produced and secreted by pancreatic α-cells (174, 332). In the CNS,

proglucagon is expressed mainly in the caudal brainstem and in selective hypothalamic neurons,

and the processing of proglucagon in these neurons appears to mirror that of the intestine

yielding GLP-1, GLP-2, glicentin, and oxyntomodulin as major products (172, 183, 333).

Receptors for both GLPs have been found in several areas of the brain that regulate appetite and

energy homeostasis (332, 333).

The two key regulators of food intake and energy balance, insulin and leptin, are secreted

in proportion to body fat mass. Both, insulin and leptin, cross the BBB and interact with key

neurons in the hypothalamus that express both IR and ObRb (43, 248). Insulin is the main

metabolic hormone that regulates glucose homeostasis, and is secreted by pancreatic β-cells.

Peripheral actions of insulin are anabolic as it increases energy storage by increasing fatty acid

synthesis, whereas central actions are catabolic as it reduces food intake and body weight (334).

Neuron-specific insulin receptor knockout mice display an obese phenotype, indicating the

importance of the central actions of insulin (12). Leptin is secreted by adipocytes and was cloned

from the obese (ob) gene. Knockout mice with mutations in the ob gene or leptin receptor db

gene are morbidly obese, and administration of leptin to ob/ob mice normalizes body weight and

neuroendocrine status (39, 267), indicating the importance of leptin in energy homeostasis. It is

well established that IR and ObRb are located in the hypothalamus (43, 248, 249), and central

administration of insulin or leptin potently reduces food intake and body weight (250, 282). It is

known that insulin and leptin regulate feeding and energy balance through interaction with

complex neural circuits comprised of appetite repressing anorexigenic and appetite stimulating

orexigenic neuropeptides. It is now well established that insulin and leptin act in the

hypothalamus to regulate orexigenic NPY/AgRP and anorexigenic POMC expression (36, 335-

337).

In contrast to the well-studied hypothalamic NPY/AgRP and POMC regulation, the

regulation of hypothalamic proglucagon by insulin and leptin remains unknown. Few studies

have examined changes in the local regulation of proglucagon and production of the GLPs in the

50

brain, mainly due to a very small number of proglucagon-expressing neurons existing in the

brain. Further, the complexity of the in vivo architecture of the hypothalamus renders these

studies very challenging to perform in intact brain or whole animal models. Therefore, using

embryonic- and adult-derived immortalized, clonal, hypothalamic cell models generated in our

laboratory, whether insulin and leptin regulate hypothalamic proglucagon gene expression was

investigated. Also the signal transduction and transcriptional mechanisms involved in any such

regulation were studied.

3.3 Results

3.3.1 Characterization of the expression profile of the hypothalamic cell

lines

A series of hypothalamic cell lines has recently been developed which display a variety

of hypothalamic phenotypes, and a few of them were reported to express the proglucagon mRNA

(18, 19). To confirm these reports, RT-PCR was conducted to show that mHypoA-2/10,

mHypoE-39, and mHypoE-20/2 express both proglucagon mRNA (Figure 7A and B) and the

insulin and leptin receptors (Figure 7C). Another cell model mHypoE36/1 does not express the

proglucagon gene. Proglucagon-positive pancreatic α-TC cells and intestinal GLUTag L-cells

served as positive controls, and fibroblasts 3T3 cells served as a negative control. Further, the

expression of other hypothalamic neuropeptides involved in appetite regulation and hormone

receptors in these cell lines were analyzed (Figure 7C). All cell models express neurotensin, IR

and ObRb, whereas only mHypoE36/1 cell model expresses endogenous GLP-1R and GLP-2R.

Currently, there are no hypothalamic cell models reported with endogenous proglucagon

expression and receptors for insulin or leptin; therefore, based on their unique neuropeptide and

receptor profile, mHypoA-2/10 and mHypoE-39 cell models were selected to study insulin- and

leptin-mediated regulation of proglucagon mRNA.

51

Figure 7. Characterization of the expression profile of the proglucagon-expressing

hypothalamic cell lines. Expression of proglucagon mRNA transcripts in pancreatic α-TC cells,

intestinal GLUTag L-cells, hypothalamic neuroendocrine cell lines mHypoA-2/10, mHypoE-39,

mHypoE-20/2, mhypoE-36/1 and fibroblasts 3T3 cells. (A) RT-PCR using specific primers for

mouse proglucagon and -actin genes. Total RNA was isolated from the indicated cell lines and

used as template for RT-PCR using the One-Step RT-PCR kit. M, markers; NTC, non-template

control; +/− RT. (B) Graphical representation of relative proglucagon mRNA transcripts levels

quantified by densitometry. Proglucagon mRNA values were normalized to -actin levels. (C)

RT-PCR analysis results for the mRNA expression of neuropeptides and receptors in the

indicated hypothalamic cells. ‘+’ indicates that the gene is expressed; ‘−’ indicates that the gene

is weakly expressed or not expressed.

52

3.3.2 Activation of signaling pathways by insulin and leptin in the

hypothalamic neuronal cells

The key signaling pathway that insulin activates is the PI3K/Akt pathway. Thus, it was

investigated if this signaling pathway is activated by insulin in the hypothalamic adult and

embryonic cells. The neuronal cells were treated with 10 nM insulin, and activation of Akt was

analyzed over 60 minutes. By Western blot analysis, it was found that insulin induced

phosphorylation of Akt at Ser473 in both adult mHypoA-2/10 and embryonic mHypoE-39 cell

lines during the entire period of 60 minutes post-treatment (Figure 8). The maximum significant

increase in Akt phosphorylation by 179% was observed at the 5 min time point in the mHypoA-

2/10 cell model [pAkt/total Akt (5 min): vehicle (0.66 ± 0.05) versus 10 nM insulin (1.83 ±

0.21), P < 0.001]. Similarly, in the mHypoE-39 cells, the maximum activation of Akt was by

390% at 5 min post-treatment [pAkt/total Akt (5 min): vehicle (0.56 ± 0.06) versus 10 nM

insulin (2.74 ± 0.45), P < 0.001] (Figure 8A and B).

Further, leptin was found to significantly increase phosphorylation of STAT3 at Tyr705

by 60% at 15 min in the adult neuronal cells [pSTAT3/total STAT3 (15 min): vehicle (0.71 ±

0.06) versus 10 nM leptin (1.13 ± 0. 19), P < 0.05] (Figure 8C). Similarly, leptin significantly

increased phosphorylation of STAT3 at 5 min in the embryonic cells by 42% [pSTAT3/total

STAT3 (5 min): vehicle (0.72 ± 0.03) versus 10 nM leptin (0.99 ± 0. 08), P < 0.01] (Figure 8D).

Interestingly, these findings indicate a temporal difference in the activation of the two pathways,

as Akt is strongly and continuously activated by insulin over 60 min, while STAT3 is only

activated by leptin for 5 to 15 minutes.

Akt can be activated by its phosphorylation in a PI3K-dependent or -independent manner.

Activated Akt can then activate or deactivate its downstream substrates via its kinase activity

such as mammalian target of rapamycin or other signaling pathways to regulate effector genes.

Activated STAT3 forms homo- or heterodimers that translocate to the cell nucleus, where they

act as transcription activators and bind to STAT elements within the promoter region of

downstream target genes to regulate their expression. Overall, activation of Akt and STAT3 in

the neuronal cell models suggests that insulin and leptin could potentially regulate downstream

signaling proteins and target genes endogenously expressed in these neuronal cells.

53

Figure 8. Insulin activates Akt and leptin activates STAT3 in the hypothalamic neuronal

cells. The mHypoA-2/10 and mHypoE-39 neurons were serum starved overnight and then

treated with either insulin (10 nM), leptin (10 nM) or vehicle. Protein was isolated over 60 min at

the indicated time points, resolved on 10% SDS-PAGE, transferred to polyvinyl difluoride

membrane, and immunoblotted with antisera for phospho-Akt, total Akt, phospho-STAT3, total

STAT3 and Gβ (G protein β subunit). Insulin increased phosphorylation of Akt in the mHypoA-

2/10 cells (A) and mHypoE-39 cell line (B). Leptin induced phosphorylation of STAT3 in both

mHypoA-2/10 (C) and mHypoE-39 (D) cell lines. Phosphorylation of Akt and STAT3 was

normalized to total Akt and total STAT3, respectively. Gβ was used as a loading control.

Representative Western blots are shown. All results shown in the bar graphs are expressed as

mean ± SEM (n = 4-6 independent experiments; *P < 0.05, **P < 0.01, ***P < 0.001 vs. vehicle

control). Statistical significance was calculated by two-way ANOVA.

54

3.3.3 Regulation of proglucagon mRNA transcript levels by insulin and

leptin

Next, direct regulation of proglucagon mRNA levels by insulin and leptin was

investigated. The adult and embryonic hypothalamic neuronal cells were exposed to 10 nM

insulin or leptin over a 24 h time course. Using real-time qRT-PCR, it was found that in the adult

mHypoA-2/10 neurons proglucagon mRNA levels were significantly upregulated by 70% by

insulin at 4 h post-treatment [vehicle (0.66 ± 0.07) versus 10 nM insulin (1.12 ± 0.11), P < 0.05]

(Figure 9A), whereas in the embryonic mHypoE-39 neurons insulin significantly suppressed the

mRNA expression by 45% and 43% at 4 h and 8 h post-treatment, respectively [4 h, vehicle

(1.04 ± 0.10) versus 10 nM insulin (0.57 ± 0.08), P < 0.05; 8 h, vehicle (1.00 ± 0.16) versus 10

nM insulin (0.57 ± 0.05), P < 0.05] (Figure 9B). With respect to the leptin action on proglucagon

mRNA expression in the adult mHypoA-2/10 cell line, it was found that leptin caused significant

upregulation of mRNA levels by 66% at 1 h post-treatment followed by downregulation by 45%

at 12 h [1 h, vehicle (1.00 ± 0.07) versus 10 nM leptin (1.66 ± 0.22), P < 0.05; 12 h, vehicle

(1.47 ± 0.15) versus 10 nM leptin (0.81 ± 0.07), P < 0.05] (Figure 9C). Leptin induced a similar

biphasic effect in the mHypoE-39 cell line by initially upregulating proglucagon mRNA by 43%

at 1 h, followed by its suppression by 34% at 12 h post-treatment [1 h, vehicle (1.15 ± 0.10)

versus 10 nM leptin (1.64 ± 0.22), P < 0.05; 12 h, vehicle (1.38 ± 0.10) versus 10 nM leptin

(0.91 ± 0.05), P < 0.05] (Figure 9D). These results clearly indicate that the proglucagon mRNA

levels are regulated by insulin and leptin in the hypothalamic neuronal cell models.

3.3.4 Reversal of insulin-mediated regulation of proglucagon mRNA levels

by PI3K inhibitors

To investigate involvement of Akt, a PI3K-dependent protein kinase, in the regulation of

proglucagon mRNA expression by insulin, pharmacological inhibitors of PI3K (LY294002 at 25

µM and wortmannin at 1 µM concentrations) were used (32). First, the efficacy and specificity

of each inhibitor was determined. The adult mHypoA-2/10 and embryonic mHypoE-39 cells

were pretreated for 1 h with either of the inhibitors before exposure to 10 nM insulin. Total

protein was isolated at 15 minutes after insulin treatment for the analysis using phospho-specific

antibody for Akt. DMSO was used as a vehicle control, as the inhibitors were dissolved in it.

55

Figure 9. Insulin and leptin regulate proglucagon mRNA levels in the hypothalamic

neuronal cells. The mHypoA-2/10 (A, C) and mHypoE-39 (B, D) cells were exposed to either

insulin (10 nM) (A, B), leptin (10 nM) (C, D) or vehicle over a 24 h time course; total RNA was

extracted at the indicated time points and used as a template for real-time RT-PCR with primers

specifically designed to amplify proglucagon mRNA. Proglucagon mRNA levels from both cell

lines were quantified using the Ct method and normalized to the internal control (γ-actin). All

results shown are relative to corresponding vehicle-treated mRNA levels at each time point and

are expressed as mean ± SEM (n = 4-6 independent experiments; *P < 0.05 vs. vehicle control).

Statistical significance was calculated by two-way ANOVA.

56

Figure 10. Regulation of proglucagon mRNA levels by insulin via activation of the

PI3K/Akt pathway. mHypoA-2/10 (A, C) and mHypoE-39 (B, D) neuronal cells were pre-

treated with a PI3K inhibitor, either LY294002 (25 μM) or wortmannin (1 μM), for 1 h followed

by either vehicle (water or DMSO) or insulin (10 nM) treatment. To determine the efficacy and

specificity of the inhibitors, total protein was isolated at 15 min following insulin treatment and

analyzed using Western blot analysis with phospho-specific antibodies against Akt (A, B).

Phosphorylation of Akt was normalized to total Akt. Gβ was used as a loading control.

Representative Western blots are shown. To investigate involvement of PI3K/Akt pathway in

proglucagon mRNA regulation by insulin, cells were harvested for total RNA isolation at 4 h

following insulin exposure, and proglucagon mRNA levels were determined by real-time RT-

PCR (C, D). mRNA levels are shown relative to water only-treated mRNA levels (set to 1.0). All

results are expressed as mean ± SEM (n = 4 independent experiments; *P < 0.05). Statistical

significance was calculated by two-way ANOVA. White bars = control (water or DMSO with or

without PI3K inhibitor); black bars = treatment (insulin with or without PI3K inhibitor).

57

Using Western blot analysis, it was found that wortmannin and LY294002 inhibited the

phosphorylation of Akt by insulin in both cell models [mHypoA-2/10 cell line (15 min): DMSO

(0.55 ± 0.13) versus DMSO + 10 nM insulin (1.90 ± 0.31), P < 0.05; LY294002 (0.38 ± 0.18)

versus LY294002 + 10 nM insulin (0.34 ± 0.20), P > 0.05; wortmannin (0.36 ± 0.19) versus

wortmannin + 10 nM insulin (0.34 ± 0.20), P > 0.05; mHypoE-39 cell line (15 min): DMSO

(0.64 ± 0.27) versus DMSO + 10 nM insulin (2.16 ± 0.33), P < 0.05; LY294002 (0.37 ± 0.24)

versus LY294002 + 10 nM insulin (0.48 ± 0.23), P > 0.05; wortmannin (0.41 ± 0.26) versus

wortmannin + 10 nM insulin (0.53 ± 0.17), P > 0.05] (Figure 10A and B).

The next step was to analyze whether the PI3K/Akt pathway is involved in the regulation

of proglucagon by insulin. The adult mHypoA-2/10 and embryonic mHypoE-39 cells were pre-

treated with the inhibitors for 1 h, followed by insulin treatment. Total RNA was isolated at 4 h

after insulin treatment and analyzed using real-time qRT-PCR. As compared to the vehicle

treatment, insulin treatment alone significantly induced an increase in the proglucagon mRNA

levels in mHypoA-2/10 neuronal cells, whereas pretreatment with wortmannin or LY294002

attenuated insulin-induced increase in proglucagon mRNA expression [mHypoA-2/10 cell line (4

h): water (0.60 ± 0.13) versus 10 nM insulin (1.10 ± 0.15), P < 0.05; DMSO (0.60 ± 0.04) versus

DMSO + 10 nM insulin (1.02 ± 0.12), P < 0.05; LY294002 (1.10 ± 0.16) versus LY294002 + 10

nM insulin (0.87 ± 0.08), P > 0.05; wortmannin (1.30 ± 0.23) versus wortmannin + 10 nM

insulin (1.09 ± 0.22), P > 0.05] (Figure 10C). In the mHypoE-39 cells, pretreatment with PI3K

inhibitors attenuated insulin-mediated repression of proglucagon mRNA expression [mHypoE-

39 cell line (4 h): water (1.15 ± 0.14) versus 10 nM insulin (0.63 ± 0.11), P < 0.05; DMSO (1.02

± 0.07) versus DMSO + 10 nM insulin (0.50 ± 0.08), P < 0.05; LY294002 (1.33 ± 0.07) versus

LY294002 + 10 nM insulin (1.29 ± 0.16), P > 0.05; wortmannin (0.98 ± 0.16) versus

wortmannin + 10 nM insulin (0.91 ± 0.15), P > 0.05] (Figure 10D). These findings indicated that

the PI3K/Akt pathway is involved in the regulation of hypothalamic proglucagon mRNA levels

by insulin. It was also found that PI3K inhibitors increase basal proglucagon mRNA levels in the

adult cells, suggesting that the PI3K pathway may regulate basal proglucagon mRNA turnover,

probably by increasing the basal rate of mRNA degradation (Figure 10C).

58

3.3.5 Reversal of leptin-mediated regulation of proglucagon mRNA levels

by JAK2/STAT3 inhibitors

It was further investigated whether activation of STAT3 by leptin is involved in the

regulation of proglucagon mRNA levels, for which pharmacological inhibitors of JAK2/STAT3

(cucurbitacin I at 5 µM and SD1008 at 10 µM concentrations) were used (338, 339). In order to

determine the efficacy and specificity of each inhibitor, the adult mHypoA-2/10 and embryonic

mHypoE-39 cells were pretreated for 1 h with either of the inhibitors before treatment with 10

nM leptin. Total protein was isolated at 15 minutes after the leptin treatment and analyzed using

a phospho-specific antibody for STAT3. As the inhibitors were dissolved in DMSO, it was used

as a vehicle control. It was found that cucurbitacin I and SD1008 inhibited the phosphorylation

of STAT3 by leptin in both cell models [mHypoA-2/10 cell line (15 min): DMSO (0.19 ± 0.02)

versus DMSO + 10 nM leptin (0.38 ± 0.04), P < 0.05; cucurbitacin I (0.14 ± 0.06) versus

cucurbitacin I + 10 nM leptin (0.13 ± 0.06), P > 0.05; SD1008 (0.11 ± 0.08) versus SD1008 + 10

nM leptin (0.08 ± 0.05), P > 0.05; mHypoE-39 cell line (15 min): DMSO (0.17 ± 0.01) versus

DMSO + 10 nM leptin (0.29 ± 0.01), P < 0.05; cucurbitacin I (0.15 ± 0.03) versus cucurbitacin I

+ 10 nM leptin (0.18 ± 0.01), P > 0.05; SD1008 (0.11 ± 0.05) versus SD1008 + 10 nM leptin

(0.10 ± 0.04), P > 0.05] (Figure 11A and B).

Next, involvement of the JAK2/STAT3 pathway in the regulation of proglucagon mRNA

levels by leptin was investigated. The adult mHypoA-2/10 and embryonic mHypoE-39 cells

were pre-treated with the inhibitors for 1 h, followed by leptin treatment and total RNA was

isolated at 1 h and 12 h. Using real-time qRT-PCR, it was found that as compared to the vehicle

treatment, leptin treatment alone significantly upregulated proglucagon mRNA levels in both

mHypoA-2/10 and mHypoE-39 neuronal cells, whereas pretreatment with cucurbitacin I or

SD1008 attenuated the leptin-induced increase in proglucagon mRNA levels [mHypoA-2/10 cell

line (1 h): PBS (1.26 ± 0.10) versus 10 nM leptin (2.15 ± 0.07), P < 0.05; DMSO (0.88 ± 0.28)

versus DMSO + 10 nM leptin (1.83 ± 0.25), P < 0.05; cucurbitacin I (0.47 ± 0.10) versus

cucurbitacin I + 10 nM leptin (0.39 ± 0.11), P > 0.05; SD1008 (0.75 ± 0.28) versus SD1008 + 10

nM leptin (0.67 ± 0.23), P > 0.05; mHypoE-39 cell line (1 h): PBS (1.27 ± 0.15) versus 10 nM

leptin (1.88 ± 0.36), P < 0.05; DMSO (0.91 ± 0.17) versus DMSO + 10 nM leptin (1.61 ±

0.05), P < 0.05;

59

60

Figure 11. Regulation of proglucagon mRNA levels by leptin via activation of the

JAK2/STAT3 pathway. mHypoA-2/10 (A, C, E) and mHypoE-39 (B, D, F) neuronal cells were

pre-treated with a JAK2/STAT3 inhibitor, either cucurbitacin I (5 μM) or SD1008 (10 μM), for 1

h followed by either vehicle (PBS or DMSO) or leptin (10 nM) treatment. To determine the

efficacy and specificity of the inhibitors, total protein was isolated at 15 min following leptin

treatment and analyzed using Western blot analysis with phospho-specific antibodies against

STAT3 (A, B). Phosphorylation of STAT3 was normalized to total STAT3. Gβ was used as a

loading control. Representative Western blots are shown. To investigate involvement of

JAK2/STAT3 pathway in proglucagon mRNA regulation by leptin, cells were harvested for total

RNA isolation at 1 h (C, D) and 12 h (E, F) following leptin treatment, and proglucagon mRNA

expression was determined by real-time RT-PCR (C-F). mRNA levels are shown relative to PBS

only-treated mRNA levels (set to 1.0). All results are expressed as mean ± SEM (n = 4

independent experiments; *P < 0.05). Statistical significance was calculated by two-way

ANOVA. White bars = control (PBS or DMSO with or without JAK2/STAT3 inhibitor); black

bars = treatment (leptin with or without JAK2/STAT3 inhibitor).

61

cucurbitacin I (0.57 ± 0.08) versus cucurbitacin I + 10 nM leptin (0.79 ± 0.26), P > 0.05; SD1008

(0.72 ± 0.11) versus SD1008 + 10 nM leptin (0.80 ± 0.21), P > 0.05] (Figure 11C and D).

At 12 h post-treatment, it was found that the JAK2/STAT3 inhibitors reversed the

downregulation of proglucagon mRNA levels caused by leptin [mHypoA-2/10 cell line (12 h):

PBS (1.30 ± 0.27) versus 10 nM leptin (0.71 ± 0.11), P < 0.05; DMSO (0.88 ± 0.03) versus

DMSO + 10 nM leptin (0.47 ± 0.04), P < 0.05; cucurbitacin I (1.12 ± 0.15) versus cucurbitacin I

+ 10 nM leptin (0.87 ± 0.08), P > 0.05; SD1008 (0.84 ± 0.14) versus SD1008 + 10 nM leptin

(1.06 ± 0.32), P > 0.05; mHypoE-39 cell line (12 h): PBS (1.27 ± 0.19) versus 10 nM leptin

(0.68 ± 0.05), P < 0.05; DMSO (1.27 ± 0.13) versus DMSO + 10 nM leptin (0.84 ± 0.05), P <

0.05; cucurbitacin I (0.51 ± 0.09) versus cucurbitacin I + 10 nM leptin (0.51 ± 0.06), P > 0.05;

SD1008 (1.28 ± 0.10) versus SD1008 + 10 nM leptin (1.50 ± 0.16), P > 0.05] (Figure 11E and

F). These findings indicate that the JAK2/STAT3 pathway is involved in the regulation of

hypothalamic proglucagon mRNA levels by leptin. Further, it was found that JAK2/STAT3

inhibitors decrease basal proglucagon mRNA levels suggesting that JAK2/STAT3 pathway may

regulate basal proglucagon mRNA turnover, probably by decreasing the basal rate of mRNA

degradation (Figure 11F).

3.3.6 Regulation of human or rat proglucagon 5’ flanking promoter

constructs by insulin and leptin

In order to determine if insulin and leptin regulate proglucagon at the level of gene transcription,

proglucagon promoter reporter plasmids were transiently transfected into the mHypoA-2/10 and

mHypoE-39 cells. The human proglucagon constructs consisted of three lengths of human

proglucagon 5’ flanking region, −332 to +58, −602 to +58, and −829 to +58, inserted into

luciferase reporter vectors. The rat proglucagon constructs consisted of two lengths of rat

proglucagon 5’ flanking region, −312 to +58, and −476 to +58, inserted into luciferase reporter

vectors. The transfection efficiency in the adult hypothalamic cell model mHypoA-2/10 was less

than 10% in contrast to more than 70% observed in the mHypoE-39 cell line. Therefore, the

embryonic cell model was selected for the transient transfection analysis. 24 h after transfection,

the cells were treated with 10 nM of insulin or leptin and harvested at 12 h post-treatment as the

positive control forskolin/IBMX treatment was observed to increase proglucagon promoter

62

reporter activity at that time point. All three human proglucagon promoter constructs had basal

luciferase expression 22.51-fold, 18.15-fold, and 24.35-fold higher than the promoterless pGL2

plasmid (Figure 12A). However, the basal activity of rat proglucagon promoter constructs was

much lower than the basal activity of the human proglucagon plasmids as it was only 2.74-fold

and 2.66-fold higher than the promoterless pGL2 plasmid (Figure 12B).

Forskolin/IBMX treatment was used as a positive control as cAMP activation has been

demonstrated to modulate expression of proglucagon expression in islet and intestinal cells

(185). It was found that forskolin/IBMX treatment affected transcription of only one plasmid that

contained 829 bases of the human proglucagon promoter [human −332/+58: PBS (22.51 ± 5.42),

10 µM forskolin/IBMX (24.10 ± 0.43), P > 0.05; human −602/+58: PBS (18.15 ± 0.33), 10 µM

forskolin/IBMX (23.34 ± 1.71), P > 0.05; human −829/+58: PBS (24.36 ± 1.63), 10 µM

forskolin/IBMX (35.35 ± 2.49), P < 0.05] (Fig. 12A). Interestingly, insulin or leptin treatments

did not affect the human plasmid transcription [human −332/+58: PBS (22.51 ± 5.42), 10 nM

insulin (23.79 ± 0.49), 10 nM leptin (19.94 ± 1.43), P > 0.05 (PBS versus leptin or insulin);

human −602/+58: PBS (18.15 ± 0.33), 10 nM insulin (22.34 ± 1.32), 10 nM leptin (17.80 ±

0.41), P > 0.05 (PBS versus leptin or insulin); human −829/+58: PBS (24.36 ± 1.64), 10 nM

insulin (26.43 ± 2.28), 10 nM leptin (22.87 ± 1.02), P > 0.05 (PBS versus leptin or insulin)]

(Figure 12A), or the rat plasmid transcription [rat −312/+58: PBS (2.74 ± 0.33), 10 nM insulin

(2.20 ± 0.27), 10 nM leptin (2.42 ± 0.27), P > 0.05 (PBS versus leptin or insulin); rat −476/+58:

PBS (2.66 ± 0.21), 10 nM insulin (1.97 ± 0.19), 10 nM leptin (2.06 ± 0.16), P > 0.05 (PBS

versus leptin or insulin)] (Figure 12B). These results indicate that insulin and leptin do not affect

the transcription of the human or rat proglucagon 5’ flanking gene promoter regions in the mouse

mHypoE-39 cell line.

3.3.7 Regulation of mRNA stability by insulin and leptin in mHypoA-2/10

and mHypoE-39 neuronal cells

Due to the lack of changes in human proglucagon promoter activity in the mouse cell

lines, the regulation of proglucagon mRNA stability in the presence of insulin or leptin was

investigated using RNA polymerase II gene transcription inhibitors (actinomycin D at 10

µg/ml and DRB at 60 µM concentrations). The adult mHypoA-2/10 and embryonic mHypoE-39

63

64

Figure 12. Insulin and leptin do not affect the transcription of proglucagon promoter

constructs, but regulate mRNA stability. mHypoE-39 cells were transfected with proglucagon

5’ flanking plasmids or promoterless control plasmid pGL2 (A, B), incubated for 24 h, then

treated with either vehicle, insulin (10 nM), leptin (10 nM), or forskolin/IBMX (10 µM). Cells

were harvested 12 h after treatment and a luciferase assay was performed. Data were normalized

to protein concentration. To investigate regulation of mRNA stability by insulin (C, F) and leptin

(D, E, G, H), mHypoA-2/10 cells (C-E) and mHypoE-39 (F-H) were serum-starved overnight,

pre-treated with a RNA polymerase II gene transcription inhibitor, either actinomycin D (Act D)

(10 µg/ml) or DRB (60 µM), for 1 h followed by either vehicle (DMSO), insulin (10 nM) or

leptin (10 nM) treatment. Total RNA was isolated at 1 h (D, G), 4 h (C, F) and 12 h (E, H) post-

treatment, and proglucagon mRNA expression was determined by real-time RT-PCR. mRNA

expression data are shown relative to DMSO with or without RNA polymerase II gene

transcription inhibitor-treated mRNA levels (set to 1.0). All results are expressed as mean ± SEM

(n = 4 independent experiments; *P < 0.05). Statistical significance was calculated by two-way

ANOVA. (C-H) White bars = control (DMSO with or without RNA polymerase II gene

transcription inhibitor); black bars = treatment (insulin or leptin with or without RNA

polymerase II gene transcription inhibitor).

65

cells were pre-treated with the inhibitors for 1 h, followed by insulin or leptin treatment. It was

observed that actinomycin D was not tolerated well by embryonic mHypoE-39 cells, as this

RNA polymerase II gene transcription inhibitor caused significant cell death at the selected dose

within 12 h. Therefore, the embryonic mHypoE-39 cells were treated with DRB only, as it was

well tolerated by these cells up to 24 h. The mechanisms utilized by actinomycin D and DRB to

inhibit transcription are different and unlike DRB that induces slow and reversible inhibition of

transcription, actinomycin D induces fast and irreversible global transcription inhibition (340),

therefore, it seems that actinomycin D induces “transcriptional stress response” in the embryonic

cells that may result in their death post-treatment (341).

The mHypoA2/10 and mHypoE-39 cells were harvested for RNA isolation at 4 h after

insulin treatment and at 1 h and 12 h following leptin treatment. Total RNA was isolated and

analyzed using real-time qRT-PCR. As compared to the vehicle DMSO treatment, insulin

treatment alone significantly induced an increase in the proglucagon mRNA levels in mHypoA-

2/10 neuronal cells that was not attenuated by the pretreatment with either actinomycin D or

DRB [mHypoA-2/10 cell line (4 h): DMSO (0.23 ± 0.04) versus DMSO + 10 nM insulin (0.39 ±

0.06), P < 0.05; actinomycin D (0.19 ± 0.04) versus actinomycin D + 10 nM insulin (0.37 ±

0.06), P < 0.05; DRB (1.88 ± 0.13) versus DRB + 10 nM insulin (2.95 ± 0.12), P < 0.05] (Figure

12C). Similar effects were observed at 1 h post-treatment with leptin, as the pretreatment with

either RNA polymerase II gene transcription inhibitor did not suppress the leptin-induced

upregulation of proglucagon mRNA expression; however, the effect was significant only with

DRB pre-treatment [mHypoA-2/10 cell line (1 h): DMSO (0.23 ± 0.05) versus DMSO + 10 nM

leptin (0.60 ± 0.17), P < 0.05; actinomycin D (0.33 ± 0.07) versus actinomycin D + 10 nM leptin

(0.53 ± 0.09), P > 0.05; DRB (1.45 ± 0.17) versus DRB + 10 nM leptin (2.86 ± 0.26), P < 0.05]

(Figure 12D). In contrast, at 12 h post-treatment with leptin, the pretreatment with the RNA

polymerase II gene transcription inhibitors reversed the leptin-induced suppression of

proglucagon mRNA expression [mHypoA-2/10 cell line (12 h): DMSO (0.55 ± 0.13) versus

DMSO + 10 nM leptin (0.23 ± 0.06), P < 0.05; actinomycin D (0.24 ± 0.05) versus actinomycin

D + 10 nM leptin (0.28 ± 0.09), P > 0.05; DRB (2.39 ± 0.17) versus DRB + 10 nM leptin (2.32 ±

0.21), P > 0.05] (Figure 12E). These findings suggest that leptin and insulin increase

proglucagon mRNA stability in the adult cell line at 1 h and 4 h post-treatment, and that the

mRNA stability remains unaffected at a later period following leptin treatment. The raw data

66

indicates that DRB, but not actinomycin D, increases basal proglucagon mRNA levels in the

adult cells, suggesting that DRB may inhibit transcription of proglucagon gene suppressor

factors that could regulate basal proglucagon mRNA levels.

In the embryonic mHypoE-39 neuronal cells, as compared to the vehicle DMSO

treatment, insulin treatment alone significantly decreased proglucagon mRNA levels that was not

reversed by the pretreatment with DRB [mHypoE-39 cell line (4 h): DMSO (0.26 ± 0.04) versus

DMSO + 10 nM insulin (0.10 ± 0.03), P < 0.05; DRB (1.32 ± 0..47) versus DRB + 10 nM

insulin (2.32 ± 0.51), P > 0.05] (Figure 12F). At 1 h post-treatment with leptin, the pretreatment

with DRB suppressed the leptin-induced upregulation of proglucagon mRNA expression

[mHypoE-39 cell line (1 h): DMSO (0.49 ± 0.14) versus DMSO + 10 nM leptin (1.31 ± 0.31), P

< 0.05; DRB (1.38 ± 0.36) versus DRB + 10 nM leptin (0.82 ± 0.27), P > 0.05] (Figure 12G). At

12 h post-treatment with leptin, the pretreatment with DRB reversed the leptin-induced

suppression of proglucagon mRNA expression [mHypoE-39 cell line (12 h): DMSO (0.15 ±

0.04) versus DMSO + 10 nM leptin (0.04 ± 0.01), P < 0.05; DRB (1.82 ± 0.13) versus DRB + 10

nM leptin (1.99 ± 0.15), P > 0.05] (Figure 12H). These findings suggest that insulin and leptin do

not regulate proglucagon mRNA stability in the embryonic cell line.

3.3.8 In silico analysis of murine proglucagon mRNA sequence for

microRNA binding sites and RNA-binding protein sites

Since leptin and insulin affect mRNA stability in the adult hypothalamic cell model, we

decided to analyze proglucagon mRNA sequence for putative microRNA (miRNA) binding sites

and RNA-binding proteins. Putative miRNA binding sites on proglucagon mRNA template were

searched using a web tool MicroInspector that analyzed proglucagon mRNA sequence for the

occurrence of binding sites for known and registered mouse-specific miRNAs (342) (Figure 13A

and B). Next, using the RNA-binding protein database (RBPDB), proglucagon mRNA sequence

was scanned for putative mRNA-binding protein sites that may be involved in regulation of

proglucagon mRNA stability (343) (Figure 13C). The in silico analysis results indicate that there

are binding sites for several miRNAs and RNA-binding proteins in proglucagon mRNA

sequence, suggesting that stability of proglucagon mRNA and thereby increase or decrease in its

levels can be regulated via activation or inactivation of miRNAs or RNA-binding proteins.

67

Whether insulin or leptin utilize this mechanism to regulate proglucagon mRNA stability needs

to be investigated further, as it is likely that the regulation of miRNAs or RNA-binding proteins

by insulin and leptin may be involved in the rapid proglucagon mRNA stabilization observed in

the adult hypothalamic neuronal cell line (Figure 12C and D).

3.4 Discussion

It is unknown if insulin and leptin have any direct action on the hypothalamic

proglucagon neurons. The lack of knowledge in this area is mostly due to inaccessibility to the

hypothalamic proglucagon-expressing neurons. The previous studies on the regulation of

hypothalamic PGDPs were conducted using fetal rat hypothalamic primary cell cultures (183);

however, these cultures are quite challenging to generate and maintain. In order to circumvent

this issue, the novel cell lines generated from the embryonic and adult mouse hypothalamii that

endogenously express proglucagon mRNA, insulin and leptin receptors were used (18, 19).

Insulin-mediated proglucagon gene regulation in the pancreas and gut has been

extensively studied (185). This is the first study to demonstrate a direct role for insulin in the

control of hypothalamic proglucagon gene expression. It was demonstrated that insulin

upregulated proglucagon mRNA expression in the adult hypothalamic cell model in contrast to

its downregulating action in the embryonic cell model. This variable regulation of proglucagon

gene expression can be explained by the age-dependent differences in the expression profile of

PGDPs that are apparently inherent to the cell models generated and immortalized from adult

and embryonic hypothalamii. The developmental studies on the hypothalamic PGDPs found that

fetal rat hypothalamus predominantly expresses glicentin, oxyntomodulin and glucagon, whereas

adult rat hypothalamus expresses glicentin and oxyntomodulin in greater amounts than glucagon

and GLP-1; this suggests that processing of proglucagon in the hypothalamus is age-dependent

and changes with development (183, 184). Thus, the differential profile of PGDPs may underlie

the differential control of proglucagon gene expression by insulin in these neuronal cells and

further investigations are required to confirm this hypothesis. Furthermore, tissue-specific

variation in the regulation of proglucagon is quite possible, as insulin was demonstrated to

inhibit islet proglucagon gene expression via regulation of gene transcription (200), whereas in

GLUTag L-cells, a murine enteroendocrine cell line, insulin has been reported to stimulate

68

A Proglucagon mRNA-binding micro-RNAs

Name of miRNA Name of miRNA

mmu-miR-494* mmu-miR-3089-5p

mmu-miR-3070a mmu-miR-3080-5p

mmu-miR-5109 mmu-miR-3100-5p

mmu-miR-2182 mmu-miR-423-3p

mmu-miR-770-5p mmu-miR-5110

mmu-miR-3960 mmu-miR-221

mmu-miR-615-5p mmu-miR-1964-3p

mmu-miR-185* mmu-miR-3070b-3p

mmu-miR-770-3p mmu-miR-1224

mmu-miR-1956 mmu-miR-762

mmu-miR-329* mmu-miR-23b*

mmu-miR-669c mmu-miR-2861

mmu-miR-27a* mmu-miR-465c-5p

mmu-miR-3103* mmu-miR-574-5p

mmu-miR-187 mmu-miR-1934

mmu-miR-667* mmu-miR-146b*

mmu-miR-1198-5p mmu-miR-346

mmu-miR-465b-5p mmu-miR-344c*

mmu-miR-212-5p mmu-miR-547

mmu-miR-743a* mmu-miR-1188

mmu-miR-671-5p mmu-miR-5125

mmu-miR-705 mmu-miR-149*

mmu-miR-19b-2* mmu-miR-877

mmu-miR-328* mmu-miR-1946b

mmu-miR-678 mmu-miR-323-5p

mmu-miR-1932 mmu-miR-23a*

mmu-miR-341 mmu-miR-34a

mmu-miR-3067 mmu-miR-743b-5p

mmu-miR-365-1* mmu-miR-3104-5p

mmu-miR-1892 mmu-miR-128-2*

mmu-miR-1901 mmu-miR-3090*

mmu-miR-696 mmu-miR-19b-1*

mmu-miR-128-1* mmu-miR-342-5p

mmu-miR-1199 mmu-miR-3063*

mmu-miR-3102* mmu-miR-3092

C Proglucagon mRNA-binding proteins

Figure 13. In silico analysis of murine proglucagon mRNA sequence for miRNA binding

sites and RNA-binding protein sites. (A, B) Nucleotide sequences for the mouse proglucagon

mRNA were obtained from NCBI-GeneBank and analyzed using a web tool MicroInspector that

analyzed proglucagon mRNA sequence for the occurrence of binding sites for known and

registered mouse-specific (mmu) miRNAs. The diagram illustrates putative miRNA binding sites

on mouse proglucagon mRNA template (A). The chart lists potential miRNAs that may bind to

mouse proglucagon mRNA (B). (C) The mouse proglucagon mRNA-binding proteins were

obtained using the RNA-Binding Protein DataBase.

69

proglucagon gene expression (199). In the pancreatic α-cells, insulin represses proglucagon gene

expression via activation of Akt (262), however, in the intestinal L-cells, it is likely that insulin

utilizes the Wnt signaling pathway, but not the Akt pathway, to stimulate proglucagon promoter

activity (199). Recently, it was reported that the cAMP activator forskolin stimulated

proglucagon gene expression and hormone production in pancreatic and intestinal endocrine

cells via activation of Epac, suggesting a PKA-independent regulation of proglucagon (264,

265). Similarly, although we found that Akt activation is required for insulin to variably regulate

proglucagon mRNA levels in the hypothalamic adult and embryonic cell models, we do not rule

out involvement of other signaling pathways in the differential regulation of adult versus

embryonic hypothalamic proglucagon mRNA by insulin.

Among multiple signalling pathways activated by insulin, the PI3K-Akt pathway remains

the main focus in many studies on the CNS. Several investigations have demonstrated that

insulin activates PI3K in neurons (259, 260), and that PI3K inhibitors can block the ability of

insulin to regulate food intake (260). In the present study, it was confirmed that in the

proglucagon-expressing selective neuronal cell models insulin activates Akt, as evidenced by an

increase in the phosphorylation of Akt. Interestingly, treatment with the PI3K inhibitors reversed

the insulin-induced changes in proglucagon mRNA levels in both neuronal cell models. This

indicates that similar to peripheral tissues, insulin mediates its action via Akt activation in the

selective hypothalamic proglucagon neurons.

A number of reports suggest that leptin interacts with proglucagon-expressing neurons in

mice and increases hypothalamic GLP-1 content as well as proglucagon mRNA levels in

brainstem neurons through the STAT signaling pathway (302, 303, 305-307). Similar to these

findings, it was detected that leptin regulated proglucagon mRNA in a STAT3-dependent

manner in the hypothalamic neuronal cell models. It was found that leptin induced a rapid

increase at 1 h in the proglucagon mRNA followed by a decrease at later time points. The leptin-

mediated early proglucagon mRNA upregulation is consistent with the action of CNTF, a

cytokine that induced proglucagon mRNA in mHypoA-2/10 cells while simultaneously

activating the JAK2/STAT3 pathway (19). In contrast, leptin-mediated delayed downregulation

is in accordance with the suppressive effect of other pro-inflammatory cytokines on proglucagon

gene expression (344). The effect of leptin detected in this study is, however, difficult to interpret

70

in the physiological context, due to the biphasic regulation of proglucagon mRNA. This action

also warrants further investigation on the regulation of synthesis of PGDPs in the hypothalamic

neuronal cells that seems to be challenging at present, because the quantity of PGDPs,

particularly GLP-1 and GLP-2, synthesized or secreted by the hypothalamic neuronal cells

appears to be quite low, most probably in picomolar or femtomolar range, and currently available

detection assays, such as EIA, ELISA, RIA or Western blot analysis, are not sensitive enough for

these studies. Notwithstanding, changes observed at the proglucagon mRNA level are reflected

at the translational level (345); and therefore, the mRNA changes can be used as a reliable

indicator of changes in PGDP production. Thus, it can be hypothesized that leptin may increase

biosynthesis of PGDPs at early time points followed by their decrease at later time points as a

compensatory response in the hypothalamic neuronal cells, and future studies to examine this

possibility will be required.

The level of mRNA in a cell is determined by either changes in mRNA stability or the

direct regulation of gene transcription at the 5’ regulatory region. The regulation of mRNA half-

life, and the cis-regulatory sequences and trans-acting factors necessary for the expression and

regulation of the proglucagon gene have been determined in islet α-cells and intestinal L-cells

(185, 194, 332, 345-348). Most studies on the proglucagon promoter have focused on the rat

gene, although some studies with promoters from other species, including a few studies using the

human gene promoter have been conducted (191, 330, 349). To complement studies with cell

lines, transgenic mice have been generated which have largely been concordant with results

based on the cell lines (188, 191, 330). Yet, little is known about the sequences or factors

required for the proglucagon gene expression in select neurons of the hypothalamus (332, 346,

347). A key reason for this limited knowledge has been the lack of a suitable neuronal cell line.

Since the embryonic and adult hypothalamic cell lines express endogenous proglucagon gene,

the regulation of proglucagon gene expression in these neuronal models was studied by

transfecting available rat and human proglucagon promoter reporter genes due to the current

unavailability of mouse proglucagon promoter reporter plasmids. It was found that the

transfection efficiency in the adult hypothalamic cell model mHypoA-2/10 was very low when

compared to more than 70% in the mHypoE-39 cell line; thus the transient transfection analysis

was continued using the embryonic cell model.

71

Previous transfection of the reporter constructs containing 5’ flanking regions of the

human proglucagon gene demonstrated that a short 312-base rat proglucagon promoter region is

sufficient for activation by cAMP, CREB and other transcription factors (193, 350, 351). By

deletion analysis performed in this project, a novel area between bases -602 and -829 in human

proglucagon promoter was identified, which is required for the promoter-driven induction by

forskolin in the hypothalamic cell lines. Unlike forskolin, insulin and leptin did not directly

affect human or rat proglucagon gene transcription at the level of the promoter in the mouse

hypothalamic cell lines. There could be two possible reasons for the lack of an effect. First, there

could be low sequence similarity between the human, rat and mouse proglucagon 5’ regulatory

regions. However, it is unlikely as human proglucagon promoter transgenic mice show that the

human proglucagon promoters are expressed in mouse neurons (191, 330). Further, the proximal

promoter elements are highly conserved in the rodent and human proximal proglucagon

promoter sequences (191). The homology between human and rat proglucagon promoters shows

that they are highly related, however, there can be several differences in the nucleotide sequence

(191). The G1 promoter is highly conserved as compared to the less well conserved G2-G4

enhancer sequences. These elements have binding sites for several transcription factors including

isl-1, cdx-2/3, Brn-4, pax-6, CREB and AP-1 (185, 191). Although using transcription factor

search programs, STAT3 binding sequences within 4000 bp upstream of the transcription

initiation site were not found, further search and investigation into distal promoter regions is

required to detect STAT binding sequences and also other cis-elements for downstream

transcription factors activated by insulin and leptin.

A 300 bp sequence that contains G2, G3 and G4 enhancer elements of rat proglucagon

promoter is sufficient for proglucagon gene expression in the mouse pancreatic islet cells,

therefore, two reporter plasmids containing -476 to +58 and -312 to +58 from rat proglucagon

promoter were generated (191, 330). Previously, these promoter constructs were used to study

proglucagon gene expression in hamster islet cell line InR1-G9 and mouse islet cell line αTC-1

(330). As is evident from Figure 6A, B, transfection of embryonic hypothalamic cells with these

two rat proglucagon promoter constructs containing -476 to +58 and -312 to +58 bp sequences

from transcription initiation site resulted in a much lower basal activity of these plasmids than

the basal activity of the human proglucagon plasmids, and no significant changes were detected

in the activity of rat proglucagon promoter plasmids by insulin or leptin treatment. Again, the

72

low basal transcription activity of rat promoters in the mouse cell line could due to low sequence

similarity between the rat and mouse proglucagon 5’ regulatory regions. Second, the 5’ flanking

sequences used for the proglucagon constructs were only 312 and 476 bp long, and thus cis-

acting responsive elements in the neuronal proglucagon gene promoter could be located further

upstream of these regions. Previously, it was found that unlike the rat proximal proglucagon

promoter elements, the equivalent human proglucagon proximal promoter sequences containing

G2, G3 and G4 enhancer elements did not induce transcriptional activity in transfected mouse

islet or intestinal cells; therefore, plasmids containing additional elements up to -602 bp, -829 bp

or larger segments of human proglucagon gene promoter were generated (191). These larger

human reporter plasmids supported transcriptional activity in the mouse pancreatic islet or

intestinal cells, and also in selected neurons of transgenic mice that contained 602 bases of

human proglucagon 5’ flanking sequence (191, 330). Thus, the use of non-species specific

promoters can confound the analysis, and therefore, the lack of insulin or leptin transcriptional

effects using the rat or human promoter in the mouse hypothalamic cell lines does not

conclusively exclude transcriptional regulation of mouse hypothalamic proglucagon gene.

Generation of mouse proglucagon promoter plasmids is required to conduct further studies.

Because of the lack of changes in the activity of proglucagon reporter gene constructs,

the stability of the proglucagon mRNA in the presence of insulin and leptin was analyzed. Using

RNA polymerase II gene transcription inhibitors actinomycin D and DRB in the hypothalamic

neuronal cells, it was detected that the changes in mRNA expression at early time points induced

by both insulin and leptin were due to increased mRNA stability, whereas leptin-induced

changes in the mRNA expression at the later time point do not seem to be controlled by changes

in the rate of mRNA decay in the adult neuronal cells. The decrease in mRNA expression caused

by leptin treatment at 12 h may occur due to suppressed transcription, for which further

investigation of distal promoter regions for putative cis- and trans-elements is required.

Mechanisms involved in proglucagon mRNA stability in islet α-cells and intestinal L-cells are

known to some extent (194, 345), however, currently, no information is available upon

regulatory mechanisms for proglucagon mRNA stability in the hypothalamus.

Regulation of gene expression at the post-transcriptional level by mRNA transcript

stability is widespread in eukaryotes (352). Among several factors that regulate mRNA stability,

73

miRNA and mRNA-binding proteins play an important role is altered turnover of mRNA. For

example, for the rapid decay of TNF-α mRNA, which contains an adenylate-uridylate-rich

elements. miR16 is required alongwith RNA-binding proteins in HeLa cells (353). miRNAs are

small (22 nucleotide), noncoding RNAs that act as negative regulators of gene expression at a

post-transcriptional level (354). They regulate either the degradation of their specific target

mRNAs or the inhibition of mRNA translation (355, 356). In this thesis, as it was found that

insulin and leptin affected hypothalamic proglucagon mRNA stability, putative miRNA binding

sites on proglucagon mRNA template were searched. It was detected that there are binding sites

for several miRNAs including miRNA128, a brain-enriched microRNA, that downregulates

doublecortin gene expression by inducing degradation of doublecortin mRNA in neuroblastoma

cells (357), and miRNA494 that downregulates PTEN in the bronchial epithelial cells (358).

Because insulin and leptin have been shown to regulate miRNA expression (359, 360), it is

possible that they may regulate proglucagon mRNA stability as well via miRNA activation or

suppression. Next, using the RBPDB program, the proglucagon mRNA sequence was scanned

for putative mRNA-binding protein sites that may also regulate proglucagon mRNA stability.

Several RNA-binding proteins were detected, including nervous system-specific RNA-binding

protein ELAVL2 or QKI that have been found to regulate gene expression in the nervous system

(361-363). Further studies to determine role of insulin, leptin and RNA-binding proteins in

hypothalamic proglucagon mRNA regulation are warranted.

Taken together, these experiments indicate that insulin and leptin can act directly upon

proglucagon neuronal cells to regulate proglucagon mRNA levels. Insulin, through an Akt-

dependent mechanism, and leptin, through JAK2/STAT3 activation, regulate proglucagon

mRNA levels, although transfections with human proglucagon promoter reporter gene constructs

indicate that insulin or leptin may not act directly at the level of transcription, but may instead act

to increase the stability of proglucagon mRNA. The PGDPs are key regulators of feeding

behavior, thus a better understanding of the mechanisms through which insulin and leptin

regulate hypothalamic proglucagon neurons is important to further expand and challenge our

current knowledge of feeding circuits.

74

Chapter 4

Glucagon-like Peptide-1 Receptor Agonist, Exendin-4,

Regulates Feeding-associated Neuropeptides in

Hypothalamic Neurons in vivo and in vitro

75

Publication:

Prasad S. Dalvi, Anaies Nazarians-Armavil, Matthew J. Purser, and Denise D. Belsham.

Glucagon-like peptide-1 receptor agonist, Exendin-4, regulates feeding-associated neuropeptides

in hypothalamic neurons in vivo and in vitro. Endocrinology (2012) 153:2208-2222

A.N.A. was a graduate student in the D.D.B. lab and assisted in the treatment of cells and RNA

isolation in the inhibitor study included in the publication and presented in this thesis. M.J.P.

was a summer research and work-study student in the D.D.B. lab working directly under

supervision of P.S.D. and performed the immunohistological study on the hypothalamic sections

of ghrelin knockout mice included in the original publication, but not presented in this thesis. All

other experiments included in the publication and presented in this thesis were designed by

P.S.D. and D.D.B. P.S.D. executed all the experiments in this thesis, analyzed all data, designed

and created all figures, wrote and revised the manuscript under the supervision of D.D.B.

Published figures:

Figure 14. Intracerebroventricular injection of exendin-4 was effective to reduce food and water

intake, and body weight in mice.

Figure 15. Exendin-4 activates hypothalamic neurons.

Figure 16. Acute exendin-4 treatment increases the number of hypothalamic neurons co-

expressing c-Fos-ir with α-MSH-, NPY-, neurotensin- or ghrelin-ir.

Figure 17. Graphical representation showing the number of neurons expressing c-Fos and

neuropeptide-ir in the ARC, VMH, DMH, PVN, LH, PeV, and the internuclear space between

the VMH and DMH of the saline- or exendin-4-treated mouse hypothalamus.

Figure 18. Expression profile of GLP-1R and appetite-regulating neuropeptides in the

hypothalamic neuronal cell lines.

Figure 19. Exendin-4 induces c-Fos activation and CREB/ATF-1 phosphorylation in the

hypothalamic neuronal cell lines.

Figure 20. Regulation of neurotensin and ghrelin mRNA expression by exendin-4 in the

mHypoA-2/30 and mHypoE36/1 neuronal cell lines.

Figure 21. Regulation of neurotensin and ghrelin mRNA expression via activation of the PKA

pathway.

Permissions were obtained to reproduce the copyrighted material.

76

4 Glucagon-like Peptide-1 Receptor Agonist, Exendin-4,

Regulates Feeding-associated Neuropeptides in Hypothalamic

Neurons in vivo and in vitro

4.1 Abstract

Exendin-4, a long-acting GLP-1R agonist, is a potential regulator of feeding behavior

through its ability to inhibit gastric emptying, reduce food intake and induce satiety. GLP-1R

activation by exendin-4 induces anorexia; however, the specific populations of neuropeptidergic

neurons activated by exendin-4 within the hypothalamus, the central regulator of energy

homeostasis, remain unclear. This study determines whether exendin-4 regulates hypothalamic

neuropeptide expression and explores the signaling mechanisms involved. The distribution and

quantity of exendin-4-induced c-Fos immunoreactivity were evaluated to determine activation of

α-MSH/POMC, NPY, neurotensin and ghrelin neurons in hypothalamic nuclei during exendin-4-

induced anorexia in mice. Additionally, exendin-4 action on neurotensin and ghrelin transcript

regulation was examined in immortalized hypothalamic neurons. With anorexia induced by

intracerebroventricular exendin-4, α-MSH/POMC and NPY neurons were activated in the ARC,

with simultaneous activation of neurotensin-expressing neurons in the PVN, and ghrelin-

expressing neurons in the ARC, PVN and periventricular hypothalamus, suggesting that neurons

in one or more of these areas mediate the anorexic action of exendin-4. In the hypothalamic

neuronal cell models, exendin-4 increased cAMP, CREB/ATF-1 and c-Fos activation, and via a

PKA-dependent mechanism regulated neurotensin and ghrelin mRNA expression, indicating that

these neuropeptides may serve as downstream mediators of exendin-4 action. These findings

provide a previously unrecognized link between central GLP-1R activation by exendin-4 and the

regulation of hypothalamic neurotensin and ghrelin. Further understanding of this central GLP-

1R activation may lead to safe and effective therapeutics for the treatment of metabolic

disorders.

77

4.2 Introduction

The hypothalamus integrates peripheral and central signals to regulate appetite and

energy homeostasis. Recently, GLP-1 has emerged as a potential negative regulator of feeding

behavior (6). The GLP-1R is widely expressed in the hypothalamus (217). GLP-1 plays multiple

roles in metabolic homeostasis due to its glucoregulatory function and inhibitory action on food

and water intake (26, 364). Endogenous GLP-1 has a short plasma half-life because of its rapid

degradation by DPP-4, therefore longer-acting GLP-1 analogs, GLP-1R agonists or DPP-4

inhibitors are under investigation (365). Exendin-4 is a long-acting GLP-1R agonist peptide that

was originally isolated from the salivary gland of the Gila monster, Heloderma suspectum.

Currently, exendin-4 and several DPP-4 inhibitors that increase the concentration of intact

endogenous GLP-1 by preventing its degradation are used to treat type 2 diabetes. Exendin-4 has

been shown to induce weight loss in overweight patients, suggesting that it can regulate food

intake and body weight by either directly interacting with peripheral GLP-1R-expressing vagal

afferents to affect gastro-intestinal motility, or readily crossing the BBB to directly interact with

hypothalamic appetite-regulating centers (219, 366, 367). However, other clinical trials indicate

that DPP-4 inhibitors are weight neutral (368). These findings suggest that the hypothalamic

GLP-1R activation mediated by exendin-4 may differ from that mediated by increased

concentrations of intact endogenous GLP-1 achieved through DPP-4 inhibition. That is, while

both GLP-1 and exendin-4 act on the hypothalamus to modulate feeding behavior, key

differences exist between the anorectic actions of endogenous GLP-1 and exendin-4 (369).

Considering the widespread clinical use of exendin-4, further research is warranted to study the

mechanisms underlying the anorectic actions of exendin-4 within the hypothalamus (7).

The hypothalamus is subdivided into several nuclei consisting of groups of neurons with

specific functions. The ARC, PVN, VMH, DMH, LH, and PeV play an important role in the

regulation of energy intake and expenditure by integrating central and peripheral orexigenic and

anorexigenic signals. It is well established that there are two distinct neuronal populations in the

ARC: NPY and α-MSH/POMC neurons. Additionally, anorexigenic neurotensin is expressed

within the ARC (105, 106, 156), PVN, DMH and LH (56), whereas orexigenic ghrelin neurons

are present in the ARC, PVN, the internuclear spaces between hypothalamic nuclei, the

perifornical region, and the ependymal layer of the third ventricle (105).

78

Exendin-4 has been shown to activate neurons in the PVN via GLP-1R-dependent

networks (221). Compelling evidence demonstrates that the GLP-1Rs located in the PVN, but

not in the ARC, reduce food intake (131), implying that GLP-1Rs in specific hypothalamic

regions have discrete effects on appetite regulation. Furthermore, the finding that GLP-1Rs

expressed on the ARC POMC neurons do not induce anorexia raises the possibility that exendin-

4 may stimulate or inhibit other, yet to be identified downstream effectors within the

hypothalamus to mediate its anorectic effect (131), such as neurotensin and ghrelin neurons,

localized to similar regions as the GLP-1R (216). Importantly, neurotensin and ghrelin have been

shown to play a significant role in feeding and energy homeostasis (6). Whether exendin-4 exerts

its effect directly on neurotensin or ghrelin neurons via GLP-1R activation remains unknown.

The GLP-1R is a G-protein-coupled receptor acting through adenylate cyclase, cAMP

and PKA activation (370). Recently, exendin-4 was demonstrated to suppress appetite and

reduce body weight by PKA activation in brainstem GLP-1R-expressing neurons (327).

However, in the hypothalamus the exact neurons activated by exendin-4 to induce anorexia are

unknown, and whether appetite-regulating neuropeptide gene expression is altered within GLP-

1R-expressing hypothalamic neurons remains to be determined. The present study proposes that

hypothalamic GLP-1R activation by exendin-4 regulates feeding-related neurons, and modulates

neuropeptide expression via cAMP/PKA activation. Using in vivo and in vitro models, the

hypothalamic neuronal activation by exendin-4 was mapped, and the modulation of neurotensin

and ghrelin gene expression following exendin-4 treatment and the signaling mechanisms

involved were studied.

4.3 Results

4.3.1 Effect of i.c.v. exendin-4 on food and water intake, and animal

weight

In order to test the efficacy of the i.c.v. injections, a short-term analysis of exendin-4

effects on food and water intake was performed. The dose of exendin-4 (100 ng/mouse) used to

induce anorexia was based on a previous report (221). It was found that i.c.v. exendin-4

significantly reduced food in ad libitum-fed wild type mice, and this effect lasted for at least 24 h

79

Figure 14. Intracerebroventricular (i.c.v.) injection of exendin-4 was effective to reduce

food and water intake, and body weight in mice. To determine the efficacy of i.c.v. exendin-4

on food intake, ad libitum-fed mice received injection of 100 ng of exendin-4 dissolved in 2 μl of

0.9% normal saline 1 h before the onset of the dark cycle. Mice were returned to their home

cages with pre-weighed amount of chow and water. Change in food (A) and water (B) intake was

measured at 2, 4, 8, and 24 h post-injection, and the body weight change (C) was recorded over

24 h. 0.9% normal saline solution was used as control treatment. All results are expressed as

mean ± SEM (n = 3-4 mice/group); *P < 0.01, **P < 0.001).

80

[cumulative food intake (g): 1 h, saline (0.67 ± 0.14) versus exendin-4 (0.13 ± 0.03), P < 0.01; 4

h, saline (1.67 ± 0.07) versus exendin-4 (0.50 ± 0.05), P < 0.01; 8 h, saline (4.63 ± 0.17) versus

exendin-4 (1.23 ± 0.07), P < 0.001; 24 h, saline (5.00 ± 0.25) versus exendin-4 (1.73 ± 0.11), P <

0.001] (Figure 14A). Similarly, water intake was found to be suppressed over 24 h time period

[cumulative water intake (ml): 1 h, saline (2.67 ± 0.27) versus exendin-4 (0.08 ± 0.07), P <

0.001; 4 h, saline (4.00 ± 0.47) versus exendin-4 (0.08 ± 0.07), P < 0.001; 8 h, saline (8.67 ±

0.27) versus exendin-4 (0.33 ± 0.27), P < 0.001; 24 h, saline (10.33 ± 0.27) versus exendin-4

(0.67 ± 0.54), P < 0.001] (Figure 14B). Furthermore, significant weight loss at 24 h in exendin-4-

treated mice compared with the saline-treated control mice was observed [change in body weight

(g): 24 h, saline 1.00 ± 0.17 versus exendin-4 -3.33 ± 0.79, P < 0.01] (Figure 14C). The dose of

exendin-4 used to induce anorexia was based on a previous report (221). Using an equivalent

central dose, Baggio et al. showed that the anorectic action of central exendin-4 was not simply a

nonspecific response, as anorexia was absent in GLP-1R knock-out mice, which lack functional

GLP-1R (371). This suggests that the anorexia induced by central exendin-4 may be mediated

via activation of GLP-1Rs expressed in hypothalamic appetite-regulating nuclei located in the

vicinity of the third ventricle. Nevertheless, as peptides delivered into the third cerebral ventricle

can access GLP-1Rs not only in the hypothalamus but also in the brainstem and the central

nucleus of the amygdala, any indirect activation of hypothalamic nuclei through activation of

these extra-hypothalamic regions that can also induce suppression of food and water intake by

their own mechanisms cannot be ruled out (209).

4.3.2 Effect of i.c.v. exendin-4 on activation of hypothalamic nuclei and

neuropeptidergic neurons

As assessed by IHC for c-Fos, a protein encoded by the immediate-early gene c-fos, a

distinctive pattern of neuronal activation was noted after the central injection of exendin-4 into

the third cerebral ventricle compared with the saline controls (Figure 15B, C, E, and F).

Significant increases in the number of c-Fos-positive neurons were detected in the hypothalamic

ARC (increase by 315%), DMH (increase by 214%), PVN (increase by 467%), and PeV

(increase by 1927%), but not in the LH and VMH [c-Fos-positive nuclei: ARC, saline (12.77 ±

1.55) versus exendin-4 (53.00 ± 5.35), P < 0.001; DMH, saline (13.62 ± 1.39) versus exendin-4

81

Figure 15. Exendin-4 activates hypothalamic neurons. Immunohistochemistry was performed

to assess neuronal activation by c-Fos-immunoreactivity (ir) in wild-type mice treated with saline

or exendin-4. A-F: Representative photomicrographs showing expression of c-Fos-ir in the

hypothalamic ARC (A-C), VMH (A-C), DMH (A-C), PVN (D-F), LH (D-F) and PeV (A-F)

regions in coronal sections of the mouse hypothalamii (as indicated on the images). Scale bar: 1

mm. Inset in each image represents a higher magnification of the boxed area. Scale bar: 100 μm.

A and D represent negative control images for the c-Fos antibody. DAB staining: nuclear brown

(c-Fos). 3V, third ventricle. G: Bar graph showing the number of c-Fos-ir neurons in the

hypothalamic regions at 2 h post-treatment. Data in the bar graph are expressed as mean ± SEM

(n = 3-4 animals/group; *P < 0.05, **P < 0.01, ***P < 0.001).

82

(42.78 ± 3.93), P < 0.001; PVN, saline (12.06 ± 1.97) versus exendin-4 (68.37 ± 13.37), P =

0.002; PeV, saline (2.37 ± 0.42) versus exendin-4 (48.03 ± 11.52), P = 0.017; LH, saline (4.25 ±

1.48) versus exendin-4 (9.52 ± 2.30), P = 0.082; VMH, saline (1.71 ± 0.45) versus exendin-4

(3.08 ± 0.94), P = 0.227] (Figure 15G). Activation of these hypothalamic regions was expected,

as these regions widely express GLP-1Rs (216). In this study, the potential activation of other

CNS areas known to express the GLP-1R, such as certain brainstem nuclei and the central

nucleus of the amygdala, was not studied as the main aim of this study was to investigate the

activation of hypothalamic nuclei involved in feeding regulation.

The next goal was to determine the neuropeptidergic neurons activated in these regions

by performing double-staining IHC for c-Fos and neuropeptide co-expression (Figs. 16 and 17).

It was found that exendin-4 significantly activated α-MSH/POMC-ir neurons (by 129%) in the

hypothalamic ARC [saline (12.51 ± 1.98) versus exendin-4 (28.69 ± 3.33), P = 0.002] (Figure

17B), and also robustly activated neurotensin neurons (by 488%) in the PVN [saline (6.62 ±

0.98) versus exendin-4 (38.92 ± 1.42), P < 0.001] (Figure 17F). Both α-MSH/POMC and

neurotensin neurons, being anorexigenic, are implicated in the inhibition of appetite.

Surprisingly, it was found that orexigenic NPY-ir neurons were significantly increased in the

ARC [saline (10.75 ± 2.35) versus exendin-4 (42.45 ± 5.91), P = 0.004] (Figure 17D), whereas

ghrelin-ir neurons were activated in the ARC, PVN and PeV of exendin-4-treated mice [ARC,

saline (6.15 ± 2.14) versus exendin-4 (22.48 ± 1.16), P = 0.005; PVN, saline (7.16 ± 0.90) versus

exendin-4 (22.52 ± 2.67), P = 0.011; PeV, saline (3.33 ± 0.53) versus exendin-4 (34.11 ± 8.51),

P = 0.042] (Figure 17H). Importantly, no changes were noticed in the number of neurons

expressing only neuropeptide-immunoreactivity in the hypothalamic regions of saline- or

exendin-4-treated animals (Figure 17A, C, E, and G). Since there is still some controversy

regarding the synthesis of ghrelin in the hypothalamus, the specificity of the ghrelin antibody

was confirmed by using the same antibody in ghrelin knockout mouse in another control

experiment (372). The antibody did not detect ghrelin-ir staining in the PVN of the ghrelin

knockout sections versus the litter-matched controls (372). These findings suggest that the

induced anorexia may be a function of complex interactions occurring between appetite-

regulating anorexigenic and orexigenic neurons in the hypothalamus.

83

Figure 16. Acute exendin-4 treatment increases the number of hypothalamic neurons co-

expressing c-Fos-immunoreactivity (ir) with α-MSH-, NPY-, neurotensin- or ghrelin-ir. A-

P: Bright-field photomicrographs showing neurons that co-express c-Fos-ir (brown nuclei) and

neuropeptide-ir (blue-black cytoplasm) in the coronal sections of the hypothalamic regions from

wild-type mice at 2 h after intracerebroventricular administration of saline (A, C, E, G, I, K, M,

O) or exendin-4 (B, D, F, H, J, L, N, P) (n = 3-4 animals/group). A and B: Co-expression of c-

Fos-ir with α-MSH-ir in the ARC. C and D: Co-expression of c-Fos-ir with NPY-ir in the ARC.

E-H: Co-expression of c-Fos-ir with neurotensin-ir in the ARC (E, F) and PVN (G, H). I-P: Co-

expression of c-Fos-ir with ghrelin-ir in the ARC (I, J), PVN (K, L), PeV (M, N), and the

internuclear space between DMH and VMH (O, P). Insets in each image represent a higher

magnification of the adjacent boxed areas. Black arrowheads represent neurons expressing only

nuclear c-Fos-ir, yellow arrowheads represent neurons expressing only cytoplasmic perinuclear

neuropeptide-ir, and red arrowheads represent double-labeled neurons with co-expression of c-

Fos-ir and neuropeptide-ir. Scale bars: 100 μm (A-P) and 10 μm (Insets).

84

Figure 17. Graphical representation showing the number of neurons expressing c-Fos and

neuropeptide-immunoreactivity (ir) in the ARC, VMH, DMH, PVN, LH, PeV, and the

internuclear space between the VMH and DMH of the saline- or exendin-4-treated mouse

hypothalamus. Double-labeled immunohistochemistry for c-Fos-ir and α-MSH-ir (B), c-Fos-ir

and NPY-ir (D), c-Fos-ir and neurotensin-ir (F), or c-Fos-ir and ghrelin-ir (H) indicates that

intracerebroventricular exendin-4 activates hypothalamic neuropeptidergic neurons. Note that

there was no change in the number of neurons expressing only α-MSH-ir (A), NPY-ir (C),

Neurotensin-ir (E) or ghrelin-ir (G) in the hypothalamic regions of saline- or exendin-4-treated

animals. Data are represented as mean ± SEM (n = 3-4 mice/group); *P < 0.05, **P < 0.01,

***P < 0.001 vs. saline treatment. Statistical significance was calculated by two-tailed, unpaired

t-test.

85

4.3.3 Expression of GLP-1R and appetite-regulating neuropeptides in

adult mHypoA-2/30 and embryonic mHypoE-36/1 neurons

Based on the in vivo findings that exendin-4 activates neurotensin and ghrelin neurons in

the hypothalamus, direct regulation of hypothalamic neurotensin and ghrelin mRNA expression

by exendin-4 was investigated using hypothalamic cell models. Our lab has previously reported

the generation of embryonic and adult mouse hypothalamic cell lines (18, 19). Using RT-PCR,

the presence of GLP-1R mRNA in both adult mHypoA-2/30 and embryonic mHypoE-36/1

neuronal cell models was confirmed (Figure 18A). GLP-1R-positive mouse jejunal tissue served

as a positive control. Further, the expression of hypothalamic neuropeptides involved in appetite

regulation in these cell lines was analyzed. Both cell models express ghrelin and neurotensin, and

the adult mHypoA-2/30 cells express POMC as well (Figure 18B). Currently, there are no

hypothalamic cell models reported with a functional endogenous GLP-1R. Therefore, we sought

to determine if the GLP-1R was functionally active in the adult and embryonic neurons by

measuring GLP-1R-mediated cAMP activation following exendin-4 treatment. Forskolin was

used as a positive control. Using cAMP-RIA, it was found that exendin-4 increased cAMP levels

in mHypoA-2/30 neuronal cell line [cAMP content (pmol/µg protein): vehicle, 0.49 ± 0.08; 1

µM forskolin, 2.35 ± 0.17; 10 nM exendin-4, 0.79 ± 0.16; 50 nM exendin-4, 0.99 ± 0.07;

forskolin versus saline, P < 0.01; exendin-4 (10 nM and 50 nM) versus saline, P < 0.05] (Figure

18C). Also in mHypoE-36/1 neuronal cells exendin-4 significantly increased cAMP content

compared to vehicle control [cAMP content (pmol/µg protein): vehicle (0.54 ± 0.03), 1 µM

forskolin (2.02 ± 0.39), 10 nM exendin-4 (0. 90 ± 0.06), 50 nM exendin-4 (0.81 ± 0.06);

forskolin versus saline, P < 0.01; exendin-4 (10 nM and 50 nM) versus saline, P < 0.05, (Figure

18D)]. These data indicate that the GLP-1Rs in these cells are responsive to exendin-4. Further,

using exendin-9-39, a GLP-1R antagonist, we studied whether exendin-4 directly activates GLP-

1R to stimulate cAMP production [16]. It was found that 1 µM exendin-9-39 alone did not

stimulate cAMP production in both cell models, but completely attenuated the stimulatory effect

of exendin-4 on cAMP levels, suggesting that exendin-4 stimulates cAMP production via GLP-

1R activation in both cell lines [cAMP content (pmol/µg protein): mHypoA-2/30, vehicle, 0.49 ±

0.08; 10 µM forskolin, 10.29 ± 0.79; 10 nM exendin-4, 0.89 ± 0.14; 1 µM exendin-9-39, 0.41 ±

0.02; 1 µM exendin-9-39 + 10 nM exendin-4, 0.55 ± 0.02; mHypoE-36/1, vehicle, 0.38 ± 0.13;

86

87

Figure 18. Expression profile of GLP-1 receptor (GLP-1R) and appetite-regulating

neuropeptides in the hypothalamic neuronal cell lines. A: Expression of GLP-1R mRNA

transcript in jejunum and the mHypoA-2/30 and mHypoE-36/1 neuronal cell lines by RT-PCR

using specific primers for mouse GLP-1R gene. Total RNA, isolated from mouse jejunum and

the indicated cell lines, was used as template for RT-PCR using One-Step RT-PCR kit. Jejunal

RNA was used as a positive control for GLP-1R expression. M, markers; NTC, non-template

control. B: RT-PCR analysis results for the mRNA expression of neuropeptides in mHypoA-2/30

and mHypoE-36/1 cells. ‘+’ indicates that the gene is expressed; ‘-’ indicates that the gene is not

expressed. C and D: Expression of functional GLP-1R by cAMP-RIA. E and F: Exendin-4 (Ex-

4) stimulates cAMP production via the GLP-1R activation. cAMP is activated by Ex-4 in

mHypoA-2/30 (C, E) and mHypoE-36/1 (D, F) neuronal cells. The cells were pretreated for 5

minutes with 1 µM exendin (9-39) [Ex(9-39)] or vehicle alone prior to a 10-minute treatment

with vehicle, forskolin (1 or 10 µM) or Ex-4 (10 or 50 nM). The amount of intracellular cAMP

was determined in triplicate by RIA. Forskolin was used as a positive control. All results are

expressed as mean ± SEM (n = 4-6; *P < 0.05, **P < 0.01 vs. vehicle control).

88

10 µM forskolin, 8.69 ± 0.71; 10 nM exendin-4, 0.89 ± 0.14; 1 µM exendin-9-39, 0.35 ± 0.01; 1

µM exendin-9-39 + 10 nM exendin-4, 0.41 ± 0.01; forskolin versus saline, P < 0.01; exendin-4

(10 nM) versus vehicle, P < 0.05] (Fig 18E and F).

4.3.4 Activation of c-Fos and CREB/ATF-1 by exendin-4 in the

hypothalamic neuronal cells

The key signaling pathway that exendin-4 activates is the cAMP/PKA pathway.

Therefore, it was next determined if the downstream effectors of this pathway, such as

CREB/ATF-1 and c-Fos, are activated by exendin-4 in the hypothalamic adult and embryonic

cells. The neuronal cells were treated with 10 nM exendin-4, and activation of CREB/ATF-1 and

c-Fos was analyzed over 6 h. By Western blot analysis, exendin-4 significantly induced c-Fos

activation in the adult mHypoA-2/30 cell line from 30 min through 4 h post-treatment [c-Fos/Gβ

(30 min): vehicle (0.73 ± 0.10) versus 10 nM exendin-4 (1.14 ± 0.05), P < 0.01] (Figure 19A).

Further, exendin-4 significantly increased phosphorylation of CREB at Ser 133 at 15 and 30 min

in the adult neuronal cells, by 53% and 58%, respectively, [pCREB/total CREB: 15 min, vehicle

(0.74 ± 0.08) versus 10 nM exendin-4 (1.13 ± 0.08), P < 0.01; 30 min, vehicle (0.83 ± 0.05)

versus 10 nM exendin-4 (1.31 ± 0.22), P < 0.01] (Figure 19B). Similarly, exendin-4

significantly increased phosphorylation of CREB at 5min in the embryonic cells by 42%,

[pCREB/total CREB: 5 min, vehicle (0.74 ± 0.07) versus 10 nM exendin-4 (1.05 ± 0.15), P <

0.05] (Figure 19D). Simultaneously, exendin-4 induced phosphorylation of ATF-1 from 15 to 60

min in the adult neuronal cells with the maximal phosphorylation at 30 min [pATF-1/total

CREB: 30 min, vehicle (0.87 ± 0.08) versus 10 nM exendin-4 (1.34 ± 0.25), P < 0.01] (Figure

19C). In the embryonic cells, exendin-4 induced significant increase in ATF-1 phosphorylation

at 5 and 30 min [pATF-1/total CREB: 5 min, vehicle (1.05 ± 0.14) versus 10 nM exendin-4 (1.83

± 0.16), P < 0.05; 30 min, vehicle (1.84 ± 0.13) versus 10 nM exendin-4 (2.55 ± 0.11), P < 0.05]

(Figure 19E).

Activated CREB or ATF-1 bind to a cAMP response element (CRE) within the promoter

region of cAMP/CREB/ATF-1 downstream target genes. This leads to the recruitment of CREB

binding protein (CBP/p300), and possibly other nuclear co-activators, to enhance transcription of

89

90

Figure 19. Exendin-4 induces c-Fos activation and CREB/ATF-1 phosphorylation in the

hypothalamic neuronal cell lines. The mHypoA-2/30 and mHypoE-36/1 cells were serum

starved overnight and then treated with exendin-4 (10 nM) or vehicle. Protein was isolated over 6

h at the indicated time points, resolved on 10% SDS-PAGE, transferred to nitrocellulose

membrane, and immunoblotted with antisera for c-Fos, phospho-CREB/ATF-1, total CREB and

Gβ (G-protein β subunit). Exendin-4 significantly increased c-Fos expression in the mHypoA-

2/30 cells (A), and induced phosphorylation of CREB (B, D) and ATF-1 (C, E) in both

mHypoA-2/30 (B, C) and mHypoE-36/1 (D, E) cell lines. c-Fos expression was normalized to

Gβ, and phosphorylation of CREB and ATF-1 was normalized to total CREB. Representative

Western blots are shown. All results shown in the bar graphs are expressed as mean ± SEM (n =

4 independent experiments; *P < 0.05, **P < 0.01 vs. vehicle control). Statistical significance

was calculated by two-way ANOVA.

91

immediate-early or late-response genes. Thus, phosphorylation of CREB may directly, via CRE,

activate c-Fos, which then heterodimerizes with members of the c-Jun, ATF and JDP families to

form activator protein-1 (AP-1), a transcription factor that in turn regulates target gene

expression. Overall, activation of CREB/ATF-1 and c-Fos in the neuronal cells suggests that

exendin-4 could potentially regulate downstream target genes that express cAMP-response

element or AP-1 consensus cis-elements in their promoter region.

4.3.5 Regulation of neurotensin and ghrelin mRNA transcript expression

by exendin-4

Next, regulation of neurotensin and ghrelin mRNA transcript levels by exendin-4 was

investigated. The adult and embryonic hypothalamic neuronal cells were exposed to 10 nM

exendin-4 over a 24 h time course. Using real-time qRT-PCR, it was found that in the adult

mHypoA-2/30 neurons neurotensin mRNA levels were significantly up-regulated by 52% at 4 h

post-treatment [vehicle (0.79 ± 0.07) versus 10 nM exendin-4 (1.20 ± 0.13), P < 0.05] (Figure

20A), whereas in the embryonic mHypoE-36/1 neurons the effect was more prominent as the

mRNA expression was increased by 64% at 6 h and remained significantly higher by 47% until

24 h [6 h, vehicle (0.80 ± 0.02) versus 10 nM exendin-4 (1.31 ± 0.06), P < 0.05; 24 h, vehicle

(0.79 ± 0.07) versus 10 nM exendin-4 (1.16 ± 0.07), P < 0.05] (Figure 20C). With respect to the

ghrelin mRNA expression, it was found that exendin-4 caused significant attenuation of mRNA

levels by 34% at 1 h post-treatment in the adult mHypoA-2/30 cell line [1 h, vehicle (1.56 ±

0.16) versus 10 nM exendin-4 (1.03 ± 0.08), P < 0.05] (Figure 20B). Similarly, exendin-4

suppressed ghrelin mRNA transcript levels by 29% at 4 h post-treatment in the embryonic

mHypoE-36/1 cell line [4 h, vehicle (1.12 ± 0.10) versus 10 nM exendin-4 (0.79 ± 0.08), P <

0.05] (Figure 20D). However, in the adult mHypoA-2/30 neurons, ghrelin mRNA levels were

significantly increased by 78% at the 24 h time point [vehicle (0.82 ± 0.20) versus 10 nM

exendin-4 (1.46 ± 0.31), P < 0.05] (Figure 20B). These results clearly indicate that neurotensin

and ghrelin mRNA levels are regulated by GLP-1R activation in the hypothalamic neuronal

models, and complement the in vivo findings. The regulation of POMC expression in the adult

mHypoA-2/30 neurons was not investigated due to failure of SYBR or TaqMan primers and

probe to detect the message by real-time qRT-PCR.

92

Figure 20. Regulation of neurotensin (NT) (A, C) and ghrelin (B, D) mRNA expression by

exendin-4 in the mHypoA-2/30 (A, B) and mHypoE36/1 (C, D) neuronal cell lines. The cells

were exposed to exendin-4 (10 nM) or vehicle over a 24 h time course; total RNA was extracted

at the indicated time points and used as a template for real-time RT-PCR with primers

specifically designed to amplify neuropeptide mRNA. NT (A, C) and ghrelin (B, D) mRNA

levels were quantified using the CT method, normalized to the internal control (γ-actin) and are

shown relative to vehicle only-treated mRNA levels set to 1.0 at 1 h time point. All results

shown are relative to corresponding control mRNA levels at each time point and are expressed as

mean ± SEM (n = 4 independent experiments; *P < 0.05 vs. vehicle control). Statistical

significance was calculated by two-way ANOVA.

93

4.3.6 Reversal of exendin-4 regulation of neurotensin and ghrelin by PKA

inhibitors

The next step was to investigate whether PKA, a cAMP-dependent protein kinase, was

involved in the regulation of neurotensin and ghrelin mRNA expression by exendin-4.

Pharmacological inhibitors of PKA (H-89 at 5 µM and PKI (14-22) amide at 1 µM

concentration) were used to pre-treat the adult and embryonic cells for 1 h before exposure to 10

nM exendin-4. PBS and DMSO were used as vehicle controls, as exendin-4 was dissolved in

PBS and PKA inhibitors were dissolved in DMSO. Total RNA was isolated at 1, 4 and 24 h after

the exendin-4 treatment and analyzed using real-time qRT-PCR. As compared to the vehicle

treatments, exendin-4 treatment significantly induced the neurotensin mRNA levels in mHypoA-

2/30 neuronal cells [4 h: PBS (0.66 ± 0.08) versus PBS + exendin-4 (1.22 ± 0.06), P < 0.05;

DMSO (0.85 ± 0.06) versus DMSO + exendin-4 (1.42 ± 0.15), P < 0.05] as well as in mHypoE-

36/1 neuronal cells [24 h: PBS (1.02 ± 0.08) versus PBS + exendin-4 (1.61 ± 0.09), P < 0.05;

DMSO (0.91 ± 0.04) versus DMSO + exendin-4 (1.27 ± 0.09), P < 0.05] (Figure 21A and C). In

contrast, exendin-4 treatment significantly decreased ghrelin mRNA levels in both cell lines

[mHypoA-2/30 cell line (1 h): PBS (1.52 ± 0.09) versus PBS + 10 nM exendin-4 (0.97 ± 0.13), P

< 0.05; DMSO (1.12 ± 0.05) versus DMSO + 10 nM exendin-4 (0.68 ± 0.10), P < 0.05;

mHypoE-36/1 cell line (4 h): PBS (1.13 ± 0.05) versus PBS + 10 nM exendin-4 (0.85 ± 0.07), P

< 0.05; DMSO (1.15 ± 0.16) versus DMSO + 10 nM exendin-4 (0.76 ± 0.05), P < 0.05] (Figure

21B and D). Further, it was found that both PKA inhibitors attenuated the exendin-4-induced

increase in neurotensin mRNA expression in both neuronal cells models [mHypoA-2/30 cell line

(4 h): PKI (0.99 ± 0.10) versus PKI + 10 exendin-4 (0.96 ± 0.08), P < 0.05; H89 (0.98 ± 0.09)

versus H89 + 10 nM exendin-4 (1.06 ± 0.01), P < 0.05; mHypoE-36/1 cell line (24 h): PKI (0.99

± 0.08) versus PKI + 10 nM exendin-4 (0.74 ± 0.07), P < 0.05; H89 (0.80 ± 0.04) versus H89 +

10 exendin-4 (0.66 ± 0.06), P > 0.05] (Figure 21A and C). On the other hand, PKA inhibition

reversed the attenuation of ghrelin mRNA transcripts caused by exendin-4 [mHypoA-2/30 cell

line (1 h): PKI (0.89 ± 0.06) versus PKI + 10 exendin-4 (0.91 ± 0.02), P < 0.05; H89 (0.98 ±

0.11) versus H89 + 10 nM exendin-4 (1.19 ± 0.07), P > 0.05; mHypoE-36/1 cell line (4 h):

94

Figure 21. Regulation of neurotensin (NT) (A, C) and ghrelin (B, D) mRNA expression via

activation of the protein kinase A (PKA) pathway. Following overnight serum starvation

mHypoA-2/30 (A, B) and mHypoE-36/1 (C, D) neuronal cells were pre-treated with PKA

inhibitors PKI 14-22 amide (1 μM) or H89 (5 μM) for 1 h followed by either vehicle (PBS or

DMSO) or exendin-4 (10 nM) treatment. At 1, 4 and 24 h following exendin-4 exposure, cells

were harvested for total RNA isolation, and NT (A, C) and ghrelin (B, D) mRNA expression was

determined by real-time RT-PCR. All results are relative to PBS only-treated mRNA levels (set

to 1.0) and are expressed as mean ± SEM (n = 4 independent experiments; *P < 0.05). Statistical

significance was calculated by two-way ANOVA. White bars = control (PBS or DMSO with or

without PKA inhibitor); black bars = treatment (exendin-4 with or without PKA inhibitor).

95

PKI (0.84 ± 0.17) versus PKI + 10 nM exendin-4 (0.90 ± 0.09), P > 0.05; H89 (0.98 ± 0.09)

versus H89 + 10 exendin-4 (1.39 ± 0.11), P < 0.05] (Figure 21B and D). It was also observed

that DMSO appears to regulate basal levels of ghrelin mRNA in the mHypoA-2/30 line, as

DMSO treatment alone resulted in attenuation of ghrelin mRNA levels as compared to PBS

treatment [PBS (1.52 ± 0.09) versus DMSO (1.12 ± 0.05), P < 0.05] (Figure 21B). Overall, these

results indicate that PKA activation is involved with the regulation of neurotensin and ghrelin

mRNA by exendin-4.

4.4 Discussion

Exendin-4 has been shown to regulate food intake and energy expenditure via GLP-1R-

dependent mechanisms (371). Using an equivalent central dose, Baggio et al. showed that the

anorectic action of central exendin-4 was not simply a nonspecific response, as anorexia was

absent in GLP-1R knock-out mice, which lack functional GLP-1R (371). In the present study,

using single-label IHC with c-Fos, a marker of neuronal activation, it was found that exendin-4,

at a dose that induced anorexia, activated hypothalamic ARC, PVN, DMH and PeV, regions that

widely express GLP-1Rs along with several neuropeptides involved in energy metabolism.

Furthermore, using double-label IHC, it was detected that central exendin-4 significantly

activated α-MSH/POMC and NPY neurons in the ARC, neurotensin-expressing neurons in the

PVN, and ghrelin-expressing neurons in the ARC, PVN and PeV. Finally, using the adult

mHypoA-2/30 and embryonic mHypoE-36/1 hypothalamic neuronal cell models, it was found

that exendin-4, in a PKA-dependent manner, increased neurotensin mRNA expression while

attenuating ghrelin mRNA levels. Overall, the in vivo findings suggest that complex interactions

may occur between satiety- and hunger-related neuropeptides in one or more hypothalamic

nuclei to mediate the overall anorexic action of exendin-4. The in vitro findings suggest that

regulation of hypothalamic neurotensin and ghrelin may lie downstream of exendin-4-mediated

GLP-1R activation. However, the present results must be interpreted with caution as these data

may not be relevant to endogenous GLP-1 action due to the distinct mechanisms of receptor

activation by endogenous GLP-1 and exendin-4, as exendin-4 is speculated to activate an

unknown receptor apart from GLP-1R activation (163, 369, 373). In support, both central and

peripheral administration of exendin-4 suppressed ghrelin levels, but the suppression of ghrelin

was neither mimicked by GLP-1 nor blocked by the GLP-1R antagonist exendin-9-39,

96

suggesting that ghrelin inhibition appears to correspond to an exendin-specific mechanism,

which might be independent of GLP-1R activation (163). Several other studies have provided

evidence that the actions of exendin-4 cannot be adequately explained by activation of the GLP-

1R, and therefore, presence of another unknown receptor is suggested. Indeed, exendin-4, but not

GLP-1, increases insulin sensitivity and glucose transport in cultured L6 myotubules in vitro

(374), and the effects of exendin-4 on cAMP levels in 3T3-L1 adipocytes are not mediated by

the GLP-1R (375). Significantly, unlike exendin-4, GLP-1 activation was unable to modify

energy expenditure in mice (221). Moreover, GLP-1 could not reproduce other endocrine effects

elicited by exendin-4, such as the reduction of thyrotropin secretion in rats (376). Thus, because

exendin-4 exerts some of its effects through a mechanism that may not involve the activation of

the GLP-1R, further investigations using GLP-1R inhibitors or GLP-1R knockout mice are

required to investigate endogenous GLP-1 action and distinguish GLP-1R-dependent and -

independent actions of exendin-4 on neuropeptidergic neuronal activation.

Recently, it was found that GLP-1R activation in the ARC does not induce anorexia via

α-MSH/POMC or NPY neurons (131), suggesting that other mediators, such as anorexigenic

neurotensin and orexigenic ghrelin neurons, may be involved. Despite their potential role in the

hypothalamic regulation of energy homeostasis (6, 139, 141-143, 153, 155), neurotensin and

ghrelin have received relatively little attention with respect to appetite regulation. While it is well

established that neurotensin-expressing neurons are expressed in the hypothalamus and exert an

anorexigenic action (56, 143), the existence of central ghrelin-expressing neurons is still

controversial and a general consensus does not exist concerning the source of central ghrelin

(377). Although ghrelin-positive neurons are located within the intact hypothalamus (105, 148,

378, 379), and display synaptic interactions with POMC, NPY and other ghrelin-containing

neurons (156, 380), at present the spectrum of functions and physiological role of ghrelin-

expressing neurons in the hypothalamus are yet to be fully elucidated. The present study found

that exendin-4 induced almost a 6-fold increase in neurotensin neuronal activation in the PVN,

compared to only less than a 2-fold increase in the ARC. This might explain why the PVN GLP-

1R system, but not that of the ARC, is predominant in mediating anorexia, as reported by

Sandoval et al. (131).

97

Although ghrelin neurons are orexigenic, it was observed that exendin-4 significantly

activated these neurons in the ARC, PVN and PeV. However, whether this neuronal activation

translates into stimulation or inhibition is not completely clear. It is possible that the activation of

hypothalamic ghrelin neurons may be complementary or compensatory to exert a function other

than appetite regulation. Exendin-4 has been shown to induce anti-inflammatory action in the

peripheral tissues (381). It also exerts neuroprotective effects on dopaminergic neurons by

inhibiting pro-inflammatory mediators (382). In this context, as ghrelin is found to exert anti-

inflammatory and neuroprotective effects (383), the ghrelin neuronal activation observed in this

study may be involved in mediating anti-inflammatory effects of exendin-4 (381). Interestingly,

both orexigenic ghrelin and anorexigenic neurotensin share a common characteristic: to stimulate

release of CRH (an anorexigenic and anxiogenic factor in the PVN) resulting in hypothalamic-

pituitary-adrenal axis (HPA) activation and stimulation of glucocorticoid release (384, 385). As

exendin-4 has been demonstrated to stimulate the HPA axis in rodents and humans (386), it is

important to explore the role of hypothalamic ghrelin and neurotensin in the HPA regulation by

GLP-1R agonists. This could elucidate possible reasons for why the GLP-1R knockout mice

exhibit paradoxically increased corticosterone responses to stress, even though they appear to

have a normal HPA axis (387). Activation of ARC and PeV neurons (particularly of ghrelin

neurons) suggests a possible involvement of ghrelin in glucoregulation and neurogenesis induced

by GLP-1R activation (19, 106, 107).

Additionally, significant activation of α-MSH/POMC and NPY neurons in the ARC was

detected. The finding that ARC α-MSH/POMC neurons are activated by exendin-4 agrees with

that reported by Sandoval et al. (131). However, it is not clear whether exendin-4-activated α-

MSH/POMC neurons actually induce anorexia, as measuring the changes in the α-MSH

synthesis and secretion or POMC mRNA levels in the ARC in vivo was beyond the scope of this

study. Nevertheless, given the fact that GLP-1R-expressing ARC POMC neurons are not

involved in appetite suppression, as reported by Sandoval et al. (131), it would be appropriate to

further investigate if these neurons play any role in mediating the central glucoregulatory action

of exendin-4 or native GLP-1 (131, 369).

Given that ARC NPY neurons do not express GLP-1Rs, the observed significant

activation of these neurons is quite intriguing. Absence of GLP-1R expression on ARC NPY

98

neurons excludes any direct action of exendin-4, suggesting their indirect activation by other

neurons, such as POMC neurons or -aminobutyric acid (GABA) cells in the ARC (388). As

neurotensin was demonstrated to inhibit feeding via Ntsr expressed in the hypothalamic ARC

(134), it is possible that the activated PVN neurotensin neurons regulate ARC NPY neurons to

ultimately inhibit them. However, whether ARC NPY neurons express Ntsr is not known.

Another possibility is that the activated ghrelin-immunopositive neurons in the ARC could

influence NPY neurons via synaptic transmission (156). Given that the ARC GLP-1R system

regulates glucose homeostasis (131), and that ARC NPY neurons are glucose-sensitive neurons

directly regulated by exogenous ghrelin (106, 107), the possible interaction between ARC NPY

and ghrelin neurons in exendin-4-mediated glucose regulation needs to be analyzed further.

A striking finding in this study was the absence of c-Fos activation in the VMH, the

nucleus involved in the control of satiety (25), and in the LH, the region known to control hunger

(24). However, the lack of c-Fos activation in the hypothalamic VMH was in accordance with

very low GLP-1R expression in this region (216). The lack of c-Fos expression in the LH

suggests that the LH is not activated by exendin-4, despite moderate expression of GLP-1R in

this region (216). This result may also indicate that orexigenic MCH and orexin/hypocretin

neurons in the LH are not activated by exendin-4. It has been demonstrated that leptin inhibits

LH indirectly via the activation of ARC POMC neurons, and that when leptin-mediated

suppression of LH neurons was reversed via ablation of the ARC, c-Fos expression was

increased in LH neurons (389). This suggests that the lack of c-Fos activation in the LH neurons

observed in this study could be due to inhibitory quiescence via exendin-4-activated ARC

POMC neurons. However, caution must be exercised in interpreting these results based on the

absence of c-Fos activation in these regions, for the c-Fos activation itself does not indicate

neuronal stimulation or inhibition unless advanced techniques, such as electrophysiological

methods or magnetic resonance imaging, are simultaneously employed.

The in vitro part of this study characterized the signaling pathways and changes in

neurotensin and ghrelin gene expression that occur as a direct result of exendin-4 action on the

hypothalamic adult mHypoA-2/30 and embryonic mHypoE-36/1 neuronal cell models (18, 19).

These cell models were used in the present study because it is speculated that they originate from

the PVN (given the co-expression of galanin and arginine-vasopressin) (18, 316, 390, 391).

99

Using both adult and embryonic neuronal cell lines, it was observed that mature neurons do not

differ from neurons of embryonic origin in the signaling mechanisms triggered by GLP-1R

activation.

The intracellular signaling pathways regulating hypothalamic neurotensin and ghrelin

gene expression following GLP-1R activation have not been previously established. In

pancreatic β-cells, GLP-1R activation stimulates adenylyl cyclase, thereby increasing cAMP, and

subsequently activating PKA. Similarly, GLP-1 is shown to activate adenylyl cyclase in rodent

brain (392), and in hypothalamic POMC neurons, GLP-1R activation stimulates cAMP/PKA

pathway to induce anorectic activity (393). In the present study, hypothalamic adult mHypoA-

2/30 and embryonic mHypoE-36/1 neuronal cell models were used because they are the only

hypothalamic cell models that express functional GLP-1R as well as neurotensin and ghrelin (18,

19). It was found that GLP-1R activation lead to up-regulation of intracellular cAMP in these

neuronal cell models, and that transcription factors CREB and ATF-1 were activated upon

exendin-4 stimulation. Furthermore, activation of c-Fos in the mHypoA-2/30 line suggests that

the in vivo activation of c-Fos found in the present study could be mediated through direct

activation of GLP-1R-expressing hypothalamic neurons. Further studies are required to

determine whether exendin-4 regulates gene expression at the promoter region of these genes.

The importance of cAMP/PKA activation in appetite regulation has been demonstrated in

several studies. Recently, hindbrain GLP-1R activation by exendin-4 was shown to induce

anorexia in a PKA-dependent manner (327). Hypothalamic activation of cAMP/PKA mediates

central regulation of satiety by inhibiting NPY-induced feeding (336). Furthermore, PKA and

CREB signaling have been shown to negatively regulate NPY gene expression (337), whereas

transcription of CART, a potent appetite-suppressing peptide, is upregulated by

cAMP/PKA/CREB activation (394). Overall, it appears that stimulation of cAMP/PKA activity

inhibits feeding behavior by increasing anorexigenic and decreasing orexigenic neuropeptide

expression. These studies are therefore consistent with the findings that exendin-4 induced

neurotensin and suppressed ghrelin expression via PKA activation. However, regulation of

hypothalamic neurotensin and ghrelin gene expression was shown to occur through other

pathways as well. Neurotensin gene expression was augmented by leptin via activation of

STAT3 and MAPK ERK1/2 (300), whereas ghrelin was repressed by insulin via both PI3K/Akt

100

and MAPK ERK1/2 pathways (395). The MAPK ERK1/2 plays an important role in food intake

control, and exendin-4 has recently been shown to induce anorexia partially by activation of

MAPK ERK1/2 in hindbrain GLP-1R neurons (327). Activation of MAPK ERK1/2 by exendin-4

in hypothalamic neurons has yet to be studied.

The induction of either PKA or CREB signaling normally serves to activate gene

transcription. Therefore, the upregulation of neurotensin mRNA expression by exendin-4 may

reflect stimulation of gene transcription at the promoter level. Neurotensin gene promoter

analysis indicates that synergistic activation of AP-1 and CRE elements is important in hormonal

regulation of neurotensin gene expression (396). Furthermore, it was demonstrated that leptin

upregulated neurotensin gene expression by activation of ATF-1 and c-Fos (300). However, it

cannot be ruled out that the increase in neurotensin mRNA was due to an increase in mRNA

stability, similar to GLP-1-induced increase in insulin mRNA stability to augment insulin mRNA

levels in β-cells (397). The suppression in ghrelin mRNA levels observed following GLP-1R

activation does not seem to be due to a decrease in transcription at the promoter level, at least in

the adult cell line, as the mRNA transcript levels decrease rapidly within 1 h post-treatment. This

down-regulation likely occurred due to destabilization of the ghrelin mRNA by activated PKA,

similar to the PKA-mediated increased decay of angiotensin II type 1 receptor mRNA (398). In

another study it was proposed that insulin-triggered changes in mRNA half-life could potentially

suppress orexigenic NPY and AgRP gene expression (399). Overall, further studies are required

to determine whether exendin-4 regulates neurotensin and ghrelin mRNA expression through the

AP-1 or CRE consensus sites, or through other cis-elements within the promoter or enhancer

region of these genes, or via specific molecular components induced by PKA that mediate

mRNA decay.

Recently, the ability of miRNAs to regulate mRNA has been addressed as a powerful

mechanism to control neuronal functions. In this context, it was demonstrated that CREB-

regulated miRNA132 regulates neuronal morphogenesis by decreasing mRNA transcript levels

of the GTPase-activating protein p250GAP (400). The induction phase of the miRNA132 is

rapid and parallels c-Fos induction (400), which raises the possibility for CREB-induced

miRNA132 to target ghrelin mRNA and rapidly suppress its expression, as seen in this study.

Whether the presence of a miR132 recognition element in ghrelin mRNA or lack of it in

101

neurotensin mRNA is responsible for differential regulation of these neuropeptide transcripts by

exendin-4 needs to be studied further.

In summary, the data presented here demonstrate that exendin-4 activates multiple

hypothalamic sites where complex interactions may occur between appetite-regulating

neuropeptidergic neurons to induce anorexia. The findings from the in vitro model of direct

action of exendin-4 on neurotensin and ghrelin mRNA regulation complement the in vivo

findings that neurotensin and ghrelin neurons, in addition to α-MSH/POMC and NPY neurons,

can be activated and regulated by central exendin-4. Ultimately, this study provides a previously

unrecognized link between exendin-4 action and neurotensin and ghrelin regulation.

102

Chapter 5

Glucagon-Like Peptide-2 Directly Regulates

Hypothalamic Neurons Expressing Neuropeptides

Linked to Appetite Control in vivo and in vitro

103

Publication:

Prasad S. Dalvi and Denise D. Belsham. Glucagon-like peptide-2 regulates hypothalamic

neurons expressing neuropeptides linked to appetite control in vivo and in vitro. Endocrinology

(2012) 153:2385-2397

All the experiments included in the publication and in this thesis were designed by P.S.D.

and D.D.B. P.S.D. executed all the experiments in this thesis, analyzed all data, designed and

created all figures, wrote and revised the manuscript under the supervision of D.D.B.

Published figures:

Figure 22. Intracerebroventricular (i.c.v.) injection of h(Gly2)GLP-2 inhibits food and water

intake, and induces weight loss in a dose-dependent manner in wild-type mice.

Figure 23. Acute h(Gly2)GLP-2 treatment activates hypothalamic appetite-regulating nuclei.

Figure 24. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity (ir) in the

hypothalamic neurons expressing α-MSH-, NPY-, neurotensin- or ghrelin-ir.

Figure 25. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity (ir) in the

hypothalamic neurons expressing NPY-, neurotensin- or ghrelin-ir.

Figure 26. Graphical representation showing the number of neurons expressing c-Fos- and

neuropeptide-immunoreactivity (ir) in the ARC, VMH, DMH, PVN, LH and internuclear space

between the DMH and LH of the saline- or h(Gly2)GLP-2-treated mouse hypothalamus.

Figure 27. Expression profile of GLP-2 receptor (GLP-2R) and appetite-regulating neuropeptides

in the hypothalamic neuronal cell lines.

Figure 28. Acute h(Gly2)GLP-2 treatment induces c-Fos activation and CREB/ATF-1

phosphorylation in the hypothalamic GLP-2R-positive mHypoA-2/30 neuronal cells.

Figure 29. Regulation of neurotensin (A, C) and ghrelin (B, D) mRNA expression by

h(Gly2)GLP-2 in the mHypoA-2/30 neuronal cell line in protein kinase A (PKA)-dependent

manner.

Permissions were obtained to reproduce the copyrighted material.

104

5 Glucagon-Like Peptide-2 Directly Regulates Hypothalamic

Neurons Expressing Neuropeptides Linked to Appetite Control

in vivo and in vitro

5.1 Abstract

GLP-2 has been postulated to affect appetite at the level of the hypothalamus. To gain

better insight into this process, a degradation-resistant GLP-2 analog, human (Gly2)GLP-2(1-33)

[h(Gly2)GLP-2] was intracerebroventricularly injected into mice to examine its action on food

and water intake and also activation of hypothalamic anorexigenic α-MSH/POMC, neurotensin,

and orexigenic NPY, and ghrelin neurons. Central h(Gly2)GLP-2 administration significantly

suppressed food and water intake with acute weight loss at 2 h. Further, central h(Gly2)GLP-2

robustly induced c-Fos activation in the hypothalamic ARC, DMH, VMH, PVN, and LH nuclei.

Differential colocalization of neuropeptides with c-Fos in specific regions of the hypothalamus

was observed. To assess whether hypothalamic neuropeptides are directly regulated by GLP-2 in

vitro, an adult-derived clonal, immortalized hypothalamic cell line, mHypoA-2/30, that

endogenously expresses functional GLP-2R and two of the feeding-related neuropeptides linked

to GLP-2R activation in vivo: neurotensin and ghrelin was used in the present study. Treatment

with h(Gly2)GLP-2 stimulated c-Fos expression and phosphorylation of CREB/ATF-1 in this

neuronal cell model. In addition, treatment with h(Gly2)GLP-2 significantly increased

neurotensin and ghrelin mRNA transcript levels by 50 and 95%, respectively, at 24 h after

treatment in a PKA-dependent manner. Taken together, these findings implicate the PKA

pathway as the means by which GLP-2 can upregulate hypothalamic neuropeptide mRNA levels

and provide an evidence for a link between central GLP-2R activation and specific hypothalamic

neuropeptides involved in appetite regulation.

5.2 Introduction

The glucagon gene encodes two homologous peptides, GLP-1 and GLP-2. These peptides

are generated by post-translational processing mainly in the enteroendocrine L cells of the

intestine, and specific neurons of the caudal brainstem (171, 172, 215, 401). Both GLPs play

105

important roles in energy homeostasis by regulating nutrient ingestion and disposal. GLP-2 is a

hormone that mainly stimulates crypt cell proliferation and inhibits apoptosis of the intestinal

epithelium to exert trophic effects (402-404). GLP-2 also regulates energy homeostasis by

regulating intestinal motility, permeability, nutrient ingestion and absorption (204, 405, 406). In

the CNS, GLP-1 and GLP-2 are synthesized in the caudal brainstem and hypothalamus (172,

215, 333, 401). The GLP-1/GLP-2-expressing preproglucagon neurons of the NTS send

extensive projections to the hypothalamic PVN and DMH nuclei, both regions known to be

involved in the regulation of feeding behavior (213). It has been speculated that GLP-2 also acts

as a neurotransmitter and exerts similar actions like GLP-1 in the CNS (214, 333).

The GLP-1 and GLP-2 mediate their actions via unique G protein-coupled receptors. The

GLP-1R is widely expressed in the hypothalamus as compared with the restricted expression of

the GLP-2R in the VMH and DMH nuclei (214, 216-218, 231). Thus, they may play distinct

roles in appetite regulation despite being synthesized in the same neurons (213). However, little

is known of GLP-2-mediated signaling in appetite-regulation. Tang-Christensen et al. discovered

that central administration of GLP-2 is involved in regulation of food intake in rats (214).

Similarly, pharmacological doses of central long-acting GLP-2 inhibited short-term food intake

in mice (218). Therefore, the mechanisms of central GLP-2R action to induce reduction in food

intake require further characterization.

The GLP-2R is predominantly expressed in the gastrointestinal tract and the CNS, with

limited expression in lung, cervix, vagal afferents and other peripheral organs (229). The

signaling mechanisms of the GLP-2R activation are not completely understood due to the lack of

cell models that endogenously express GLP-2R. Data collected from heterologous cell lines

transfected with GLP-2R have demonstrated that cAMP/PKA- and AP-1-dependent pathways

are activated by GLP-2 (231). Also fetal rat intestinal cell cultures treated with h[Gly2]-GLP-2

were shown to have increased cAMP concentration (407). Further, the PI3K/Akt pathway has

been shown to be implicated in the stimulatory effects of GLP-2 to enhance intestinal IGF-I

mRNA transcript levels in the intestinal subepithelial myofibroblasts (408). However, few

studies have addressed central GLP-2R activation and the target hypothalamic neurons largely

remain elusive. Thus, clonal, hypothalamic cell lines that endogenously express functional GLP-

2R were used. The present study investigated activation of hypothalamic α-MSH/POMC-,

106

neurotensin-, NPY-, and ghrelin-expressing neurons following GLP-2 exposure in vivo, and

determined potential signaling pathways involved in vitro.

5.3 Results

5.3.1 Effect of i.c.v. h(Gly2)GLP-2 on food/water intake, and animal weight

To establish the efficacy of the h(Gly2)GLP-2, the effect on food and water intake over a

short-term exposure was studied. For the dose-response study, the anorexigenic doses of

h(Gly2)GLP-2 (100 ng, 1 µg and 5 µg) were selected based on previously conducted studies in

rodents (214, 218). It was found that the cumulative food intake was significantly reduced by

66.67% by i.c.v. dose of 5 µg of h(Gly2)GLP-2, as compared with saline at 1 h post-treatment in

ad libitum-fed wild type mice [cumulative food intake 1 h (g): saline, 0.15 ± 0.03; 100 ng

h(Gly2)GLP-2, 0.15 ± 0.03; 1 µg h(Gly

2)GLP-2, 0.13 ± 0.03; 5 µg h(Gly

2)GLP-2, 0.05 ± 0.03; 5

µg h(Gly2)GLP-2, versus saline, P < 0.05] (Figure 22A). At 2 h post-treatment, no additional

food intake suppression was detected. On the contrary, the suppression of cumulative food intake

was attenuated and it did not differ across all treatment groups [food intake 2 h (g): saline, 0.37 ±

0.03; 100 ng h(Gly2)GLP-2, 0.33 ± 0.03; 1 µg h(Gly

2)GLP-2, 0.33 ± 0.03; 5 µg h(Gly

2)GLP-2,

0.28 ± 0.08] (Figure 22A). Further, it was found that different doses of h(Gly2)GLP-2 did not

have any significant effect on cumulative water intake at 1 h post-treatment [cumulative water

intake 1 h (ml): saline, 1.75 ± 0.25; 100 ng h(Gly2)GLP-2, 1.67 ± 0.29; 1 µg h(Gly

2)GLP-2, 1.33

± 0.29; 5 µg h(Gly2)GLP-2, 1.00 ± 0.58] (Figure 22B). In contrast, at 2h post-treatment,

cumulative water intake was significantly reduced by 41.75% and 43.75% by both i.c.v. doses of

1 and 5 µg of h(Gly2)GLP-2, respectively, as compared with saline treatment [cumulative water

intake 2 h (ml): saline, 4.00 ± 0.41; 100 ng h(Gly2)GLP-2, 3.00 ± 0.87; 1 µg h(Gly

2)GLP-2, 2.33

± 0.76; 5 µg h(Gly2)GLP-2, 2.25 ± 0.48; 1 and 5 µg h(Gly

2)GLP-2 versus saline, P < 0.05]

(Figure 22B). Based on the anorexia induced, the dose of 5 µg of h(Gly2)GLP-2 was selected to

determine its effect on body weight. It was found that the weight changes following single i.c.v.

dose of h(Gly2)GLP-2 were remarkable because the weight loss was significant within 1 and 2 h

[change in weight (g): saline (1h), -0.15 ± 0.01 versus 5 µg h(Gly2)GLP-2 (1h), -0.75 ± 0.18, P

< 0.05; saline (2h), -0.20 ± 0.16 versus 5 µg h(Gly2)GLP-2 (2h), -0.75 ± 0.06, P < 0.05] (Figure

22C). Further duration of these rapid weight losses and the recovery period were not studied.

107

Figure 22. Intracerebroventricular (i.c.v.) injection of h(Gly2)GLP-2 inhibits food and

water intake, and induces weight loss in a dose-dependent manner in wild-type mice. To

determine the efficacy of i.c.v. h(Gly2)GLP-2 to induce anorexia, ad libitum-fed mice received

injection of either 100 ng, 1 µg or 5 µg of h(Gly2)GLP-2 dissolved in 2 μl of 0.9% normal saline

1 h prior to the onset of the dark cycle. Mice were returned to their home cages with pre-weighed

amount of chow and water. Changes in food (A) and water (B) intake, and animal weight (C)

were measured at 1 and 2 h post-injection. 0.9% normal saline solution was used as control

treatment. All results are expressed as mean ± SEM (n = 3-4 mice/group); *P < 0.05 vs. saline).

108

5.3.2 Effect of h(Gly2)GLP-2 on activation of hypothalamic nuclei and

neuropeptidergic neurons

As assessed by IHC for c-Fos expression, a distinctive pattern of neuronal activation was

noted after the central injection of h(Gly2)GLP-2 compared with the saline controls (Figure 23

A-F). Significant increases in the number of c-Fos-positive neurons were detected in the

hypothalamic ARC (increase by 152%), DMH (increase by 342%), VMH (increase by 99%), LH

(increase by 411%) and PVN (increase by 429%) [c-Fos-positive nuclei: ARC, saline, 14.77 ±

1.42 versus 5 µg h(Gly2)GLP-2, 37.25 ± 2.28, P < 0.001; DMH, saline, 12.89 ± 1.04 versus 5 µg

h(Gly2)GLP-2, 56.98 ± 4.83, P < 0.001; VMH, saline, 3.24 ± 0.85 versus 5 µg h(Gly

2)GLP-2,

6.45 ± 1.14, P = 0.046; LH, saline, 4.04 ± 0.81 versus 5 µg h(Gly2)GLP-2, 20.64 ± 2.61, P <

0.001; PVN, saline, 13.61 ± 1.50 versus 5 µg h(Gly2)GLP-2, 72.06 ± 10.44, P < 0.001] (Figure

23G).

The next goal was to determine the neuropeptidergic neurons activated in the

hypothalamic nuclei by performing double-staining IHC for c-Fos and neuropeptide co-

expression (Figs. 24 and 25). It was found that in the ARC, single dose of i.c.v. h(Gly2)GLP-2

significantly increased the number of c-Fos-expressing α-MSH/POMC neurons by 105% [saline,

12.76 ± 1.92 versus 5 µg h(Gly2)GLP-2, 26.22 ± 2.55, P < 0.001] (Figure 26B), NPY neurons by

140% [saline, 11.07 ± 2.53 versus 5 µg h(Gly2)GLP-2, 26.55 ± 4.69, P < 0.001] (Figure 26D),

neurotensin neurons by 332% [saline, 6.90 ± 2.59 versus 5 µg h(Gly2)GLP-2, 29.83 ± 1.67, P <

0.001] (Figure 26F) and ghrelin neurons by 87% [saline, 10.03 ± 1.72 versus 5 µg h(Gly2)GLP-

2, 18.71 ± 0.70, P = 0.003] (Figure 26H). Further, it was found that in the hypothalamic PVN,

h(Gly2)GLP-2 robustly increased the number of c-Fos-positive nuclei with neurotensin co-

expression by 504% [saline, 9.07 ± 2.85 versus 5 µg h(Gly2)GLP-2, 54.75 ± 8.67, P = 0.007]

(Figure 26F) and ghrelin co-expression by 233% [saline, 8.92 ± 1.66 versus 5 µg h(Gly2)GLP-2,

29.67 ± 0.96, P < 0.001] (Figure 26H). In the hypothalamic DMH, a significant increase in c-Fos

and neurotensin and NPY co-expressing neurons by 416% and 334%, respectively, was detected

[neurotensin: saline 9.10 ± 0.92 versus 5 µg h(Gly2)GLP-2, 46.94 ± 4.38, P= 0.003; NPY: saline

9.11 ± 1.22 versus 5 µg h(Gly2)GLP-2, 39.49 ± 7.45, P = 0.001] (Figure 26H).

109

Figure 23. Acute h(Gly2)GLP-2 treatment activates hypothalamic appetite-regulating

nuclei. Immunohistochemistry was performed to assess neuronal activation by c-Fos-

immunoreactivity (ir) in wild-type mice treated with intracerebroventricular (i.c.v.) saline or

h(Gly2)GLP-2 (hGLP-2). Representative photomicrographs showing expression of c-Fos-ir in the

hypothalamic ARC, DMH, VMH, LH, and PVN regions in coronal sections of the mouse

hypothalamii. A-F: low magnification (X50) images of the hypothalamic regions (scale bars, 1

mm). A’-F’: High magnification (X400) images representative of the regions indicated by arrows

in images A-F, respectively (scale bars, 100 μm). c-Fos–DAB-positive neurons are visible as

brown nuclei. G: Bar graph showing the number of c-Fos-ir neurons in the hypothalamic regions

at 2 h post-treatment. Data in the bar graph are expressed as mean ± SEM (n = 3-4

animals/group; *P < 0.05, **P < 0.001).

110

Figure 24. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity (ir) in the

hypothalamic neurons expressing α-MSH-, NPY-, neurotensin- or ghrelin-ir. A-L: Bright-

field photomicrographs showing neurons that coexpress c-Fos-ir (brown nuclei) and

neuropeptide-ir (blue-black cytoplasm) in the coronal sections of the hypothalamic arcuate

(ARC) and paraventricular (PVN) nuclei from wild-type mice at 2h after intracerebroventricular

administration of saline (A, C, E, G, I, K) or h(Gly2)GLP-2 (B, D, F, H, J, L). A and B:

Coexpression of c-Fos-ir with a-MSH-ir in the ARC. C and D: Coexpression of c-Fos-ir with

NPY-ir in the ARC. E-H: Coexpression of c-Fos-ir with neurotensin-ir in the ARC (E, F) and

PVN (G, H). I-L: Coexpression of c-Fos-ir with ghrelin-ir in the ARC (I, J) and PVN (K, L).

Black arrowheads represent neurons expressing only nuclear c-Fos-ir, white arrowheads

represent neurons expressing only cytoplasmic perineuclear neuropeptide-ir, and black arrows

represent double-labeled neurons with coexpression of c-Fos-ir and neuropeptide-ir. 3v, third

cerebral ventricle. Original magnification: X400, scale bar: 100 μm.

111

Figure 25. Acute h(Gly2)GLP-2 treatment induces c-Fos-immunoreactivity (ir) in the

hypothalamic neurons expressing NPY-, neurotensin- or ghrelin-ir. A-J: Bright-field

photomicrographs showing neurons that co-express c-Fos-ir (brown nuclei) and neuropeptide-ir

(blue-black cytoplasm) in the coronal sections of the hypothalamic dorsomedial (DMH),

ventromedial (VMH), lateral hypothalamic nucleus (LH) and internuclear space between DMH

and LH (INS), from wild-type mice at 2h after intracerebroventricular administration of saline

(A, C, E, G, I) or h(Gly2)GLP-2 (B, D, F, H, J). A and B: Coexpression of c-Fos-ir with NPY-ir

in the DMH. C and D: Coexpression of c-Fos-ir with NT-ir in the DMH. E and F: Coexpression

of c-Fos-ir with NT-ir in the VMH. G and H: Coexpression of c-Fos-ir with NT-ir in the LH. I

and J: Coexpression of c-Fos-ir with Ghr-ir in the INS. Black arrowheads represent neurons

expressing only nuclear c-Fos-ir, white arrowheads represent neurons expressing only

cytoplasmic perineuclear neuropeptide-ir, and black arrows represent double-labeled neurons

with coexpression of c-Fos-ir and neuropeptide-ir. f, fornix. Original magnification: X400, scale

bar: 100 μm.

112

Figure 26. Graphical representation of neurons expressing c-Fos- and neuropeptide-

immunoreactivity (ir) in the ARC, VMH, DMH, PVN, LH and internuclear space between

the DMH and LH of the saline- or h(Gly2)GLP-2-treated mouse hypothalamus. Double-

labeled immunohistochemistry for c-Fos-ir and α-MSH-ir (B), c-Fos-ir and NPY-ir (D), c-Fos-ir

and neurotensin-ir (F), or c-Fos-ir and ghrelin-ir (H) indicates that intracerebroventricular

h(Gly2)GLP-2 activates hypothalamic neuropeptidergic neurons. Note that there was no change

in the number of neurons expressing only α-MSH-ir (A), NPY-ir (C), neurotensin-ir (E) or

ghrelin-ir (G) in the hypothalamic regions of saline- or h(Gly2)GLP-2-treated animals. Data are

represented as mean ± SEM (n = 3-4 mice/group); *P < 0.05, ** P < 0.01, *** P < 0.001 vs.

saline treatment. Statistical significance was calculated by two-tailed, unpaired t-test.

113

The number of c-Fos-positive nuclei co-expressing neurotensin-immunoreactivity were

significantly increased by 404% in the VMH [saline 4.01 ± 1.67 versus 5 µg h(Gly2)GLP-2,

20.20 ± 5.02, P = 0.045], and by 535% in the LH [saline 4.01 ± 1.67 versus 5 µg h(Gly2)GLP-2,

20.20 ± 5.02, P = 0.031]. Also, the number of c-Fos-positive nuclei co-expressing ghrelin-

immunoreactivity were significantly increased by 242% in the internuclear space between DMH

and LH [saline 5.28 ± 1.76 versus 5 µg h(Gly2)GLP-2, 18.06 ± 2.09, P = 0.003]. The number of

neurons expressing only neuropeptide-immunoreactivity remained unchanged in these

hypothalamic regions of saline- or h(Gly2)GLP-2-treated animals (Figure 26A, C, E, and G).

5.3.3 Expression of GLP-2R and appetite-regulating neuropeptides in

adult mHypoA-2/30 cell line

Direct regulation of hypothalamic neurotensin and ghrelin mRNA expression by

h(Gly2)GLP-2 was studied using a clonal, immortalized hypothalamic neuronal cell model, adult

mHypoA-2/30 neuronal cell line. Prior to using this cell model, the presence of GLP-2R mRNA

in the mHypoA-2/30 neuronal cells was confirmed using RT-PCR (Figure 27A). In another cell

model, embryonic mHypoE-36/1 cell line, GLP-2R gene expression was not confirmed. GLP-

2R-positive mouse jejunal tissue mRNA was used as a positive control. Also, the expression of

hypothalamic neuropeptides involved in appetite regulation was analyzed and it was found that

neurotensin and ghrelin genes are expressed in this cell line (Figure 27B). Currently, there are no

hypothalamic cell models reported with a functional endogenous GLP-2R. Therefore, to

determine if the GLP-2R is functionally active in the adult mHypoA-2/30 neuronal cells,

activation of cAMP was measured by cAMP-RIA following GLP-2 treatment using two different

concentrations (10 and 50 nM). Forskolin (1µM) was used as a positive control for cAMP

activation. Using cAMP-RIA, it was found that both doses of GLP-2 significantly increased

cAMP levels in the neuronal cell line [cAMP content (pmol/µg protein): vehicle, 0.49 ± 0.08; 1

µM forskolin, 2.35 ± 0.17; 10 nM GLP-2, 0.79 ± 0.11; 50 nM GLP-2, 0.92 ± 0.06; forskolin

versus saline, P < 0.01; GLP-2 (10 nM and 50 nM) versus vehicle, P < 0.05] (Figure 27C).

Increases in the cAMP content indicate that the GLP-2R expressed in these cells is functional

and responsive to GLP-2. Further, using GLP-2(3-33), a GLP-2R antagonist, it was studied

whether h(Gly2)GLP-2 directly activates GLP-2R to stimulate cAMP production. It was

114

Figure 27. Expression profile of GLP-2R and appetite-regulating neuropeptides in the

hypothalamic neuronal cell lines. A: Expression of GLP-2R mRNA transcript in jejunum and

the hypothalamic neuronal cell lines mHypoA-2/30 and mHypoE-36/1 by RT-PCR using

specific primers for mouse GLP-2R gene. Total RNA, isolated from mouse jejunum and the

indicated cell lines, was used as template for RT-PCR using One-Step RT-PCR kit. Mouse

jejunal RNA was used as a positive control for GLP-2R expression. M, markers; NTC, no

template control. B: RT-PCR analysis results for the mRNA expression of neuropeptides in

mHypoA-2/30 neuronal cells. ‘+’ indicates that the gene is expressed; ‘-’ indicates that the gene

is not expressed. C: Expression of functional GLP-2R by cAMP-RIA in mHypoA-2/30 neuronal

cells. D: h(Gly2)GLP-2 (hGLP-2) stimulates cAMP production via the GLP-2R activation. The

cells were pretreated for 5 minutes with 1 µM GLP-2 (3-33) or vehicle alone prior to a 10-minute

treatment with vehicle, forskolin (1 or 10 µM), GLP-2 (10 or 50 nM) or hGLP-2 (10 nM). The

amount of intracellular cAMP was determined in triplicate by RIA. Forskolin was used as a

positive control. All results are expressed as mean ± SEM (n = 4; *P < 0.05, **P < 0.01 vs.

vehicle control).

115

found that 1 µM GLP-2(3-33) alone did not stimulate cAMP production, but completely

attenuated the stimulatory effect of h(Gly2)GLP-2, suggesting that h(Gly

2)GLP-2 stimulates

cAMP production via GLP-2R activation [cAMP content (pmol/µg protein): vehicle, 0.40 ±

0.02; 10 µM forskolin, 9.79 ± 0.28; 10 nM h(Gly2)GLP-2, 0.79 ± 0.17; 1 µM GLP-2(3-33), 0.41

± 0.01; 1 µM GLP-2(3-33) + 10 nM h(Gly2)GLP-2, 0.45 ± 0.01; forskolin versus saline, P <

0.01; h(Gly2)GLP-2 (10 nM) versus vehicle, P < 0.05] (Figure 27D).

5.3.4 Activation of CREB/ATF-1 and c-Fos by h(Gly2)GLP-2 in the

hypothalamic neuronal cells

The key signaling pathway that GLP-2 activates is the cAMP/PKA pathway. Therefore,

the next step was to determine downstream effectors of cAMP/PKA pathway activated by

h(Gly2)GLP-2 in the hypothalamic adult neuronal cells. Therefore, neuronal cells were treated

with 10 nM h(Gly2)GLP-2 and activation of CREB/ATF-1 and c-Fos was analyzed over 6 h. By

western blot analysis, it was found that h(Gly2)GLP-2 significantly induced c-Fos activation by

62% in the adult mHypoA-2/30 cell line at 2 h [c-Fos/Gβ: vehicle, 0.91 ± 0.07 versus 10 nM

h(Gly2)GLP-2, 1.69 ± 0.14, P < 0.05] (Figure 28A). Similarly, h(Gly

2)GLP-2 significantly

increased phosphorylation of CREB at Ser 133 at 15 min, 1 h and 2 h by 63%, 33% and 59%,

respectively, in the adult neuronal cells (Figure 28B). Simultaneously, h(Gly2)GLP-2 induced

significant increase in phosphorylation of ATF-1 from 5 min to 6 h and the maximum increase

was by 72% at 2 h post-treatment in these neuronal cells (Figure 28C).

5.3.5 Regulation of neurotensin and ghrelin mRNA transcript levels by

h(Gly2)GLP-2

Next, neurotensin and ghrelin mRNA transcript regulation by h(Gly2)GLP-2 was

investigated. The adult hypothalamic neuronal cells were exposed to 10 nM h(Gly2)GLP-2 over

a 24 h time course. Using real-time qRT-PCR, it was found that in the adult mHypoA-2/30

neurons, neurotensin mRNA levels were significantly up-regulated by 50% at 24h post-treatment

[neurotensin mRNA expression (fold change relative to -actin): vehicle, 1.02 ± 0.11 versus 10

nM h(Gly2)GLP-2, 1.53 ± 0.17, P < 0.05] (Figure 29A). Similarly, it was observed that

h(Gly2)GLP-2 significantly increased ghrelin mRNA levels by 95% at 24 h post-treatment

116

Figure 28. Acute h(Gly2)GLP-2 treatment induces c-Fos activation and CREB/ATF-1

phosphorylation in the hypothalamic GLP-2R-positive mHypoA-2/30 neuronal cells. The

mHypoA-2/30 neuronal cells were serum starved overnight and then treated with h(Gly2)GLP-2

(10 nM) or vehicle. Protein was isolated over 6 h at the indicated time points, resolved on 10%

SDS-PAGE, transferred to PVDF membrane, and immune-blotted with antisera for c-Fos,

phospho-CREB/ATF-1, total CREB and Gβ (G-protein β subunit). h(Gly2)GLP-2 significantly

increased c-Fos expression (A), and induced phosphorylation of CREB (B) and ATF-1 (C) in the

mHypoA-2/30 cells. c-Fos expression was normalized to Gβ, and phosphorylation of CREB and

ATF-1 was normalized to total CREB. Representative Western blots are shown. All results

shown in the bar graphs are expressed as mean + SEM (n = 4 independent experiments; *P <

0.05, **P < 0.01 vs. vehicle control). Statistical significance was calculated by two-way

ANOVA.

117

Figure 29. Regulation of neurotensin (A, C) and ghrelin (B, D) mRNA levels by

h(Gly2)GLP-2 in the mHypoA-2/30 neuronal cell line in protein kinase A (PKA)-dependent

manner. Following overnight incubation with DMEM containing 0.5% FBS, the cells were

exposed to h(Gly2)GLP-2 (10 nM) or vehicle over a 24h time course and at the indicated time

points RNA was isolated (A, B). For the use of PKA inhibitors, the neuronal cells were pre-

treated with PKA inhibitor PKI 14-22 (1 mM) or H89 (5 mM) for 1h followed by either vehicle

(PBS or DMSO) or h(Gly2)GLP-2 (10 nM) treatment and the RNA was isolated at 24h time

point (C, D). Total RNA was used as a template for real-time RT-PCR with primers specifically

designed to amplify neurotensin or ghrelin mRNA. Neurotensin (A, C) and Ghrelin (B, D)

mRNA levels were quantified using the standard curve method and normalized to the internal

control (γ-actin). All results shown are relative to corresponding control mRNA levels at each

time point (A, B) or to PBS only-treated mRNA levels (set to 1.0) (C, D), and are expressed as

mean ± SEM (n = 4 independent experiments; *P < 0.05, **P < 0.01). White bars (C, D)

represent vehicle (PBS or DMSO)-treated samples (with or without PKA inhibitor); black bars

represent h(Gly2)GLP-2-treated samples (with or without PKA inhibitor).

118

in the adult mHypoA-2/30 cell line [ghrelin mRNA levels (change relative to -actin): vehicle,

0.80 ± 0.19 versus 10 nM h(Gly2)GLP-2, 1.56 ± 0.28, P < 0.05] (Figure 29B). These results

clearly indicate that neurotensin and ghrelin mRNA levels are regulated by GLP-2R activation in

the hypothalamic neuronal models, and complement the findings within the in vivo setting.

5.3.6 PKA inhibitors reverse the h(Gly2)GLP-2-induced up-regulation of

neurotensin and ghrelin mRNA transcript levels

Further, the role of PKA in the regulation of neurotensin and ghrelin mRNA transcript

levels by h(Gly2)GLP-2 was determined. Pharmacological inhibitors of PKA (H-89 at 5 µM and

PKI (14-22) amide at 1 µM concentration) were used to pre-treat the adult neuronal cells for 1 h

before exposure to 10 nM h(Gly2)GLP-2. PBS and DMSO were used as vehicle controls, as

h(Gly2)GLP-2 was dissolved in PBS and PKA inhibitors were dissolved in DMSO. Total RNA

was isolated at 24 h after the h(Gly2)GLP-2 treatment and analyzed using real-time qRT-PCR.

As compared to the vehicle treatments, at 24 h, h(Gly2)GLP-2 treatment significantly induced

increase in the neurotensin mRNA levels [PBS, 0.91 ± 0.02 versus PBS + 10 nM h(Gly2)GLP-2,

1.61 ± 0.19, P < 0.05; DMSO, 0.68 ± 0.03 versus DMSO + 10 nM h(Gly2)GLP-2, 1.12 ± 0.12, P

< 0.05] (Figure 29C). Further, it was found that both PKA inhibitors did not affect the basal

neurotensin mRNA levels, but significantly attenuated the h(Gly2)GLP-2-induced increase in

neurotensin mRNA levels in the mHypoA-2/30 neuronal cell model [PKI, 0.98 ± 0.13 versus

PKI + 10 nM h(Gly2)GLP-2, 0.69 ± 0.10, P < 0.05; H89, 0.69 ± 0.10 versus H89 + 10 nM

h(Gly2)GLP-2, 0.50 ± 0.04, P < 0.05] (Figure 29C). Next, the role of PKA in the ghrelin mRNA

upregulation caused by h(Gly2)GLP-2 was investigated. At 24 h, h(Gly

2)GLP-2 significantly

increased ghrelin mRNA levels as compared to the vehicle treatment [PBS, 0.72 ± 0.10 versus

PBS + 10 nM h(Gly2)GLP-2, 1.42 ± 0.06, P < 0.05; DMSO, 0.72 ± 0.02 versus DMSO + 10 nM

h(Gly2)GLP-2, 1.09 ± 0.07, P < 0.05] (Figure 29D). Similar to the suppression of neurotensin

mRNA levels detected at 24 h, both PKA inhibitors attenuated the h(Gly2)GLP-2-induced up-

regulation in ghrelin mRNA levels [PKI, 0.98 ± 0.12 versus PKI + 10 nM h(Gly2)GLP-2, 1.13 ±

0.15; H89, 0.82 ± 0.06 versus H89 + 10 nM h(Gly2)GLP-2, 1.05 ± 0.16] (Figure 29D). Overall,

these results indicate that PKA activation is involved with the regulation of neurotensin and

ghrelin mRNA by h(Gly2)GLP-2.

119

5.4 Discussion

The proglucagon system plays an important role in the central regulation of energy

homeostasis (7, 333). Many studies have described the hypothalamic actions of GLP-1 on energy

homeostasis. However, the potential central effects of GLP-2 still remain largely unknown. The

hypothalamus is comprised of various nuclei, a subdivision of which are known to be involved in

the regulation of feeding behavior. Besides the well-documented NPY and POMC neurons, two

other major regulators of energy homeostasis, neurotensin and ghrelin neurons are also expressed

within the ARC (56, 105, 106, 156). The neurotensin system in the hypothalamus is extensive,

and strong neurotensin immunoreactivity is found in the ARC, PVN, DMH and LH, with low

immunoreactivity in the VMH (56, 409). Ghrelin neurons are also located in the PVN, the

perifornical region and the internuclear spaces between hypothalamic nuclei (105). At present,

the neuronal phenotypes expressing GLP-2R in the DMH and VMH are not clear. It was found

that h(Gly2)GLP-2 remarkably activated main hypothalamic regions involved in regulation of

energy homeostasis. Furthermore, it was detected that central h(Gly2)GLP-2 differentially

activated α-MSH/POMC, NPY, neurotensin, and ghrelin neurons in a number of these regions.

These findings indicate that GLP-2R activation involves complex interactions between several

hypothalamic nuclei, activation of feeding-related neuropeptidergic neurons, and transcriptional

regulation of downstream neuropeptides.

Central h(Gly2)GLP-2 administration activated several hypothalamic regions with a

distinct pattern of c-Fos expression in comparison with the long acting GLP-1R agonist, exendin-

4 (372). Not only the major nuclei that are located in the vicinity of the third ventricle were

activated, but also the lateral hypothalamus was activated. As the DMH and VMH have been

demonstrated to express GLP-2R, activation of c-Fos in these areas suggests direct action of the

h(Gly2)GLP-2, however activation of ARC, PVN and LH could be due to indirect activation, as

GLP-2R expression is not found in these regions (214, 218). Because these regions function

together to regulate energy balance (410), their indirect activation by GLP-2R-expressing

neurons is quite possible. Interestingly, except for the VMH, these hypothalamic regions highly

express GLP-1R (217). Thus, there is a remote possibility of cross-reactivity of h(Gly2)GLP-2

with GLP-1R at the pharmacological dose used in the present study. However, it has been

demonstrated that h(Gly2)GLP-2 activates only the GLP-2R-expressing cells, but not those

120

expressing GLP-1R (218). Moreover, the effects of h(Gly2)GLP-2 on food intake in GLP-1R

knockout mice were not attenuated by disruption of GLP-1R signaling, but rather potentiated,

suggesting that GLP-2 does not mediate its effects on feeding through GLP-1R (218). However,

based on the activation of several hypothalamic nuclei by pharmacological h(Gly2)GLP-2,

specificity of GLP-2 action in the hypothalamus needs to be investigated further using GLP-2R

antagonists or GLP-2R knock-down strategy.

The detection of significant activation of NPY and neurotensin neurons in the DMH and

neurotensin neurons in the VMH coordinates well with the expression of GLP-2R in these areas.

However, whether NPY or neurotensin neurons express GLP-2R in vivo is not known. The

significant increase in the number c-Fos-immunopositive α-MSH/POMC, NPY, neurotensin and

ghrelin neurons in those areas that do not express GLP-2R is intriguing. c-Fos immunostaining

may have arisen from activation of projections from the GLP-2R-expressing neurons of the

DMH or VMH, or from activation of brainstem GLP-2R-expressing neurons by h(Gly2)GLP-2

carried from the third to the fourth ventricle. These findings suggest that during the central GLP-

2R-induced anorexia not only anorexigenic neurons, but also orexigenic neurons are activated in

multiple hypothalamic nuclei. Furthermore, activation of all these areas by h(Gly2)GLP-2

implies a complex interaction between these nuclei and neurons that ultimately enhances the

perception of satiety and thereby facilitates anorexia and decreased body weight.

It is not quite clear whether endogenous GLP-2 or peripherally administered

h(Gly2)GLP-2 cross the BBB. Because GLP-2 and GLP-1 have nearly 50% sequence homology

and are secreted in parallel from intestinal L-cells (411), and also because GLP-1, glucagon and

oxyntomodulin cross the BBB (219, 412, 413), it is likely that GLP-2 can also cross the BBB.

Therefore, as the long acting GLP-2, teduglutide, is currently undergoing clinical trials for short

bowel syndrome, careful consideration should be given to the novel findings on the central

action of GLP-2. It must be noted that although recent studies in humans found no effect of

intravenous GLP-2 on food intake (414), it was found that GLP-2 decreases fat mass and

increases lean mass in short-bowel patients (415). The present data show that pharmacological

dose of h(Gly2)GLP-2 decreases food and water intake, and induces body weight loss within 2 h

post-i.c.v. administration. It is possible that the body weight loss could be entirely due to

significant water intake suppression due to GLP-2 action on the central angiotensin II receptors

121

or vasopressin-expressing neurons that were not examined in the present study (416). On the

other hand, the fact that GLP-2 reduces food intake yet increases the CNS ghrelin levels suggests

a role for ghrelin signaling to decrease water intake (417). Further, it can be speculated that the

weight loss was due to the stimulation of diuresis or defecation that also needs further attention.

Apart from appetite regulation, other potential actions from the pharmacological

activation of central GLP-2R have yet to be fully delineated. GLP-2 has been shown to reduce

glutamate-induced cell death in cultured murine hippocampal and cortical cells (418). Similarly,

ghrelin is also found to exert neuroprotective effects by its anti-inflammatory action in the CNS

(383). Thus, ghrelin may potentially mediate anti-inflammatory action and thereby

neuroprotective effects of central GLP-2R action (418). Further, GLP-2 has been shown to

stimulate proliferation of cultured rat astrocytes (419). A synthetic neurotensin agonist was

found to induce an increase in the proliferation of the astrocytic cell lines (420), whereas ghrelin

was demonstrated to promote neurogenesis (421, 422). Whether neurotensin and ghrelin act as

downstream effectors of GLP-2 action in cell proliferation is not known and requires further

study. Furthermore, GLP-2-mediated regulation of murine hippocampal neurons by increasing

glucose uptake implicates an important role for GLP-2 in neurotransmitter release and hormone

secretion from GLP-2R-positive neurons and endocrine cells (423). As hypothalamic α-

MSH/POMC, NPY and ghrelin neurons are implicated in glucose regulation (106, 107, 131),

whether GLP-2-mediated activation plays any role in peripheral glucose metabolism needs to be

explored further. This study focused exclusively on the action of GLP-2 in the hypothalamus, but

extrahypothalamic action of GLP-2 also needs to be investigated.

The direct action of GLP-2 on the regulation of hypothalamic neuropeptides remains

unstudied due to the lack of endogenous GLP-2R-expressing neuronal cells. Thus, the adult

mouse-derived hypothalamic mHypoA-2/30 neuronal cell model that expresses functional GLP-

2R was used in the current study. The hypothalamic cell model represents a clonal, homogeneous

population of a single neuron, and provides a useful tool to perform gene regulation and

mechanistic studies that are quite challenging to perform in vivo (19). GLP-2 stimulates cAMP

production and activates PKA to exert its CNS actions (392, 418, 423). Therefore, the present

study focused on the cAMP/PKA pathway in the regulation of hypothalamic neurotensin and

ghrelin mRNA expression. Direct GLP-2R activation lead to an increase in intracellular cAMP,

122

and activation of transcription factors CREB and ATF-1. Furthermore, the observed increase in

c-Fos expression in these neuronal cells correlates well with the in vivo c-Fos activation found in

the GLP-2R-expressing DMH and VMH regions, suggesting that the DMH and VMH neurons

could be directly activated by central h(Gly2)GLP-2. Similarly, GLP-2 was found to up-regulate

c-Fos mRNA in BHK-GLP-2R cells as well as in astrocytes to promote cellular proliferation

(231, 419), again pointing to the involvement of GLP-2 in cell proliferation.

Neurotensin and Ntsr are widely expressed in the CNS controlling a number of

physiological functions, such as regulation of circadian rhythm, anti-psychotic action, analgesia,

thermoregulation, regulation of HPA axis, and neuromodulation of dopamine neurotransmission.

Particularly, hypothalamic neurotensin neurons are known to regulate feeding behavior (143).

Ntsr deficiency moderately increases food intake and body weight, and blocks neurotensin-

induced anorexia in mice, implicating neurotensin-Ntsr signaling pathway in feeding and body

weight regulation (139). The observed increase in neurotensin transcript levels in the present

study suggests that GLP-2 may activate neurotensin in favor of anorexigenic action quite similar

to other satiety-inducing neuropeptides and hormones.

Central ghrelin-expressing neurons are the main source of ghrelin in the CNS (105, 424).

Apart from the peripheral ghrelin, it is reported that central ghrelin is also involved in the

regulation of food intake and body weight (153, 155). In the hypothalamic neuronal cell model, a

stimulatory effect on ghrelin mRNA expression was observed. Recently, it was found that GLP-

1R activation by exendin-4 induced a decrease in the ghrelin mRNA levels in agreement with the

anorexigenic action of GLP-1R agonism (372). Also, insulin, a satiety factor and adiposity

signal, was shown to directly inhibit ghrelin expression in another hypothalamic cell model

(395). Similar to the exendin-4 or insulin action, negative regulation of ghrelin mRNA levels by

GLP-2 was expected due to its orexigenic characteristics. The positive regulation may suggest

that ghrelin could play a role in the promotion of hypothalamic neuronal survival, as ghrelin is

known to exert anti-inflammatory and neuroprotective effects in rat brain (383). Although,

ghrelin and neurotensin exert opposite effects on food intake and energy homeostasis, they exert

comparable actions in improving memory and learning, as is evident from decreased ghrelin and

neurotensin mRNA transcript expression in the brains of patients suffering from Alzheimer's

123

disease (425). Thus, it is possible that GLP-2-activated hypothalamic ghrelin and also

neurotensin neurons may potentially exert other functions unrelated to appetite regulation.

The cAMP/PKA pathway has been demonstrated to play an important role in appetite

regulation. Hypothalamic activation of cAMP/PKA mediates central regulation of satiety by

inhibiting NPY-induced feeding (336). Recently, it was shown that exendin-4 activated

hindbrain GLP-1R to induce anorexia in a PKA-dependent manner (327). Furthermore,

cAMP/PKA/CREB signaling upregulates transcription of CART, a potent appetite-suppressing

peptide (394), but PKA and CREB have been shown to negatively regulate NPY gene expression

(337). Thus, it appears that stimulation of cAMP/PKA activity inhibits feeding behavior by

increasing anorexigenic and decreasing orexigenic neuropeptide expression. This agrees with the

finding that the h(Gly2)GLP-2 induced anorexigenic neurotensin mRNA transcript levels via

PKA activation, but does not correlate with the increase in the orexigenic ghrelin mRNA

transcript levels. In fact, at present, the exact role of ghrelin expression in the hypothalamus is

not clear. In addition to mature ghrelin, the ghrelin gene also encodes an entirely different

peptide, obestatin (149). Although the physiological relevance of obestatin remains unclear,

initially it was found to exert anorectic effect that was subsequently disputed (149, 150).

Further, circulating ghrelin exists in two forms, acylated ghrelin and des-acyl ghrelin. Acylated

ghrelin acts as an orexigenic peptide to increase food intake, whereas des-acyl ghrelin, although

its function is not quite clear, has been shown to decrease food intake and gastric emptying

(426). Thus, further investigation is needed to ascertain the correlation between GLP-2-mediated

PKA-dependent stimulation of ghrelin mRNA transcript levels and post-translational processing

of the hypothalamic precursor polypeptide proghrelin.

Apart from cAMP/PKA pathway, regulation of hypothalamic neurotensin and ghrelin

gene expression can occur through alternative pathways. Neurotensin gene expression may be

augmented by leptin via activation of STAT3 and MAPK ERK1/2 (300), whereas ghrelin gene

expression can be repressed by insulin via both PI3-K/Akt and MAPK ERK1/2 pathways (395).

In a similar fashion, GLP-2 has been shown to activate other pathways to mediate intestinal cell

survival and proliferative actions i.e. Akt in the intestinal epithelium (232), and the wnt/β-catenin

signaling pathway in the intestinal crypt cells through an indirect mechanism requiring IGF-

1R/IGF-1 signaling (233). As well, h(Gly2)GLP-2 was found to stimulate IGF-1 mRNA through

124

PI3K/Akt pathway in intestinal subepithelial fibroblasts (408). Likewise, whether GLP-2R

activates MAPK ERK1/2 and PI3K/Akt pathways in hypothalamic neurons remains to be

investigated.

It is normally observed that the activation of PKA or CREB induces gene transcription.

Therefore, the h(Gly2)GLP-2-induced increase in neurotensin and ghrelin mRNA transcript

levels could suggest stimulation of 5’ regulatory promoter elements, such as CRE, to induce gene

transcription. Previous neurotensin gene promoter analysis study indicates that AP-1 site and

CRE element are involved in hormonal regulation of this gene (396). Recently, it was

demonstrated that neurotensin gene expression was induced by leptin via activation of

transcription factors ATF-1 and c-Fos (300). However, it is also possible that the increase in

neurotensin mRNA could be due to an increase in mRNA stability. Contrary to neurotensin gene

promoter region, mouse ghrelin promoter region is not well characterized; whether it contains

regulatory AP-1 site and CRE element remains unknown. Although human ghrelin promoter

does not have AP-1 site or CRE element, glucagon and its second messenger cAMP enhanced

the ghrelin promoter activity, suggesting that ghrelin gene transcription may be regulated by

some cell-specific transcription factors and cofactors to integrate cAMP stimulation to activate

ghrelin gene transcription (427). Thus, further studies are required to determine whether GLP-2

regulates hypothalamic neurotensin and ghrelin mRNA transcript levels through the AP-1 or

CRE consensus sites or through other cis-elements within the promoter or enhancer region of

these genes, or via specific molecular components induced by PKA to increase mRNA stability.

In summary, the results of the present study demonstrate that pharmacological action of

central GLP-2 results in activation of multiple hypothalamic regions to trigger complex

interactions between appetite-regulating neuropeptidergic neurons to induce transient

anorexigenic effect. Using a novel GLP-2R-positive hypothalamic cell line, it was found that

h(Gly2)GLP-2, in a PKA-dependent manner, directly regulates mRNA transcript levels of

feeding-related neuropeptides neurotensin and ghrelin. Overall, these findings suggest that

central GLP-2R may serve as a potential target for pharmacological intervention to modulate

neuropeptides involved in appetite regulation.

125

Chapter 6

Discussion

126

6 Discussion

6.1 Overall conclusions

The PGDPs are generated from a single common proglucagon precursor expressed in

pancreatic islet α-cells, gut enteroendocrine L-cells, and in selective brainstem and hypothalamic

neurons. Until now, the vast majority of research has focused only on the regulation of peripheral

proglucagon gene expression; therefore, much remains unknown about the regulation of

hypothalamic proglucagon gene expression. Specifically, the mechanisms by which two major

hormones, insulin and leptin, regulate proglucagon-expressing hypothalamic neurons have not

been elucidated. For over last two decades, a wealth of data has widened our understanding of

the role of PGDPs in the periphery and the CNS. The GLPs serve important roles in the control

of energy balance, gastro-intestinal motility, nutrient absorption and glucose homeostasis.

Particularly, GLP-1 has been well established as a central regulator of physiological functions

ranging from energy homeostasis, fluid homeostasis, memory and neuronal regeneration.

However, we are still far from having clear knowledge of potential central actions of other

PGPDs including GLP-2. Although PGDPs have emerged as potential regulators of feeding

behavior (1), the exact mechanisms of action of GLP-1R and GLP-2R stimulation in the

hypothalamus are not completely clear; particularly how GLP-1R and GLP-2R stimulation

regulates hypothalamic appetite-regulating neuropeptides has not been investigated. In this

thesis, the hypothalamic control mechanisms at the level of proglucagon-specific neurons were

defined; specifically, the molecular mechanisms utilized by insulin and leptin to regulate

expression of proglucagon in the hypothalamic neurons were studied using novel neuronal cell

models. Further, the actions of long-acting GLP-1R and GLP-2R agonists on mRNA transcript

levels of appetite-regulating hypothalamic neuropeptides that have remained elusive until now

were investigated (372, 428).

Central actions of insulin and leptin are critical in the control of appetite regulation and

maintenance of energy homeostasis. Insulin and leptin activate catabolic POMC neurons (36,

335), while inhibiting anabolic NPY/AgRP neurons in the hypothalamus (429, 430). Apart from

these neuronal circuits that have been well established as critical appetite regulators, PGDPs also

play important roles in regulation of feeding behavior (6). Similar to insulin and leptin, glucagon

127

plays an important role in energy homeostasis, and its central administration suppresses food

intake in rats and chicks (431, 432). The available data suggest that centrally administered GLP-

1 acts on the hypothalamus to regulate food intake and body weight (26, 209, 210). Although

GLP-2 has intestinotrophic activity in rodents, it also functions as a hormonal ileal brake since it

was demonstrated that GLP-2 dose-dependently inhibited centrally-induced antral motility in

pigs (204). Further, central administration of GLP-2 is involved in regulation of food intake in

rats and mice (214, 218). Acute central or peripheral administration of another PGDP,

oxyntomodulin, inhibits food intake in rodents, whereas chronic administration reduces body

weight (205, 433). Thus, it is evident that exogenous PGDPs act as central regulators of appetite,

and hence it is imperative to study the regulation of hypothalamic proglucagon-expressing

neurons and action of hypothalamic PGDPs. At present, it is not known whether central,

peripheral, or a combination of both sources of PGDPs signal hypothalamic nuclei to regulate

energy homeostasis. For example, it is possible that due to its very short plasma half-life, gut-

derived GLP-1 may not even reach the hypothalamus or other parts of the central nervous

system. Therefore, only centrally-derived GLP-1 may be functional in these central regions or

peripheral GLP-1 may act only on the circumventricular regions of the brain which have an

incomplete BBB, such as ME or area postrema. The ambiguity and lack of knowledge is mostly

due to inaccessibility to the hypothalamic proglucagon-, GLP-1R- or -2R-expressing neurons.

Previous studies on the regulation of hypothalamic PGDPs were conducted on fetal rat

hypothalamic primary cell cultures (183, 202); however, these cultures are quite challenging to

generate and maintain. In order to circumvent this technical issue, the present research used

phenotypically distinct cell lines generated from embryonic and adult mouse hypothalamii that

endogenously express proglucagon mRNA, insulin and leptin receptors, and receptor signaling

proteins (18, 19).

At present, it is unknown if insulin and leptin have any direct action on the hypothalamic

proglucagon neurons, and the regulatory mechanisms utilized by these molecules remain

unstudied. This thesis research characterized the signaling pathways and changes in proglucagon

mRNA levels that occur as a direct result of insulin and leptin administration to adult and

embryonic hypothalamic, clonal cell lines (Figure 30). A new evidence that insulin and leptin

can act directly upon proglucagon neurons to regulate proglucagon mRNA levels was

demonstrated. It was found that insulin activates an Akt-dependent pathway and leptin triggers

128

Figure 30. Summary of the mechanisms activated by insulin and leptin to regulate

proglucagon mRNA transcript levels in the mHypoA-2/10 and mHypoE-39 neuronal cells. Exposure of insulin and leptin leads to the activation of classic signal transduction pathways in

the novel hypothalamic proglucagon-expressing neurons: induction of Akt by insulin and

JAK2/STAT3 by leptin to regulate proglucagon mRNA levels. Insulin and leptin regulate

proglucagon mRNA stability through unidentified mechanisms and/or may trigger yet unknown

transcriptional mechanisms to modulate proglucagon mRNA levels in the hypothalamic neuronal

cells. TF, transcription factor; upward pointing solid green arrows represent upregulation and

downward pointing red arrows represent downregulation of mRNA levels or stability; solid blue

arrows represent interactions by known mechanisms, whereas dashed blue arrows and question

marks represent yet to be identified mechanisms.

129

activation of STAT3 to regulate hypothalamic proglucagon mRNA levels in adult and embryonic

cell models models. It was further determined that insulin and leptin do not regulate activity of

the transfected human or rat proglucagon promoter reporter constructs in the mouse embryonic

cells, but rather affect proglucagon mRNA stability. Overall, these findings suggest that insulin

and leptin can act directly on specific hypothalamic neurons to regulate proglucagon mRNA

levels. A better understanding of the mechanisms through which insulin and leptin regulate

hypothalamic proglucagon neurons will further enable us to understand roles of PGDPs in

energy homeostasis.

GLP-1R activation by exendin-4 has been shown to regulate food intake and energy

expenditure (371). The widespread clinical use of exendin-4 as an anti-diabetic drug warrants

further research on the mechanisms underlying the anorectic actions of exendin-4 within the

hypothalamus (7). The present study identified hypothalamic neuropeptidergic neurons

responsive to central GLP-1R activation in vivo, and elucidated the direct action of exendin-4 on

neurotensin and ghrelin mRNA transcript regulation in vitro (Figure 31). It was found that

exendin-4 activated hypothalamic ARC, PVN, DMH and PeV, regions that widely express GLP-

1R along with several neuropeptides involved in energy metabolism. Furthermore, it was

detected that central exendin-4 significantly activated α-MSH/POMC and NPY neurons in the

ARC, neurotensin-expressing neurons in the PVN, and ghrelin-expressing neurons in the ARC,

PVN and PeV. Overall, the in vivo findings suggest that complex interactions may occur between

satiety- and hunger-related neuropeptides in one or more hypothalamic nuclei to mediate the

overall anorexic action of exendin-4. Finally, using the hypothalamic neuronal cell models, it

was found that exendin-4, in a PKA-dependent manner, increased neurotensin mRNA levels

while attenuating ghrelin mRNA levels. These in vitro findings complement the in vivo findings

and suggest that regulation of hypothalamic neurotensin and ghrelin may lie downstream of

exendin-4 GLP-1R activation. Given that central exendin-4 induces net inhibition of both

drinking and feeding that results in significant weight loss, it is likely that hypothalamic

anorexigenic neurotensin neurons are stimulated and orexigenic ghrelin neurons are inhibited in

vivo. Ultimately, this study provides a previously unrecognized link between exendin-4 action

and neurotensin and ghrelin regulation.

130

Figure 31. Summary of the mechanisms activated by exendin-4 to regulate neurotensin and

ghrelin mRNA transcript levels in the mHypoA-2/30 and mHypoE-36/1 neuronal cells.

Exendin-4 activates classic cAMP-PKA signal transduction pathway in the novel hypothalamic

GLP-1R-expressing neuronal cells to regulate neurotensin and ghrelin mRNA levels. Exendin-4

may trigger yet unknown transcriptional mechanisms and/or modify mRNA stability through

unidentified mechanisms to regulate neurotensin and ghrelin mRNA levels in the hypothalamic

neuronal cells. NT, neurotensin; upward pointing solid green, yellow and black arrows represent

upregulation and downward pointing red arrows represent downregulation of mRNA levels or

stability; solid blue arrows represent interactions by known mechanisms, whereas dashed blue

arrows and question marks represent yet to be identified mechanisms.

131

In the periphery, following nutrient ingestion, GLP-1 and GLP-2 are secreted in

equimolar quantities from intestinal L-cells (7, 434, 435). The circulating biologically active

GLP-1 is degraded within two minutes (436), whereas GLP-2 is more stable, with plasma half-

life of about seven minutes (437, 438). Inactivation of these peptides by DDP-4 is responsible for

their relatively short plasma half-lives. In the CNS, GLP-1 and GLP-2 are synthesized in the

caudal brainstem and the hypothalamus, and the expression pattern of the central proglucagon

neurons appears to be well conserved across mammalian species (172, 215, 333, 401). The GLP-

1/GLP-2-expressing proglucagon neurons of the NTS send extensive projections to the

hypothalamic PVN and DMH, both regions known to be involved in the regulation of feeding

behavior (213). Because GLP-2 is co-localized with GLP-1 in the NTS region and is likely co-

secreted with GLP-1, it has been speculated that GLP-2 also acts as a neurotransmitter and exerts

similar actions like GLP-1 in the CNS (214, 333). Only a few studies have addressed the

signaling mechanisms triggered by central GLP-2R activation in the CNS. Unlike GLP-1, GLP-2

is not a potent appetite regulator, as peripheral GLP-2, at physiological plasma concentration,

does not contribute significantly to regulate appetite in humans (414, 439, 440). Nevertheless,

Tang-Christensen et al. discovered that central administration of high amounts of GLP-2 is

involved in regulation of food intake in rats (214). Similarly, pharmacological doses of central

long-acting GLP-2 inhibited short-term food intake in mice (218). These studies demonstrate that

pharmacological intervention with the central GLP-2R potentially results in appetite suppression.

Therefore, the mechanisms of central GLP-2R action to induce reduction in food intake require

further characterization. Using in vivo and in vitro models, the present thesis investigated

activation of hypothalamic anorexigenic α-MSH/POMC and neurotensin, and orexigenic NPY

and ghrelin neurons following central administration of long-acting GLP-2, and determined

possible signal transduction pathways involved in this regulation (Figure 32).

To evaluate the response to acute central GLP-2R stimulation, a degradation-resistant

GLP-2, h(Gly2)GLP-2, was infused into the third cerebral ventricle of adult mice to induce

transient anorexigenic effects similar to previous reports (214, 218). During this period of

transient anorexia, it was found that h(Gly2)GLP-2 remarkably activated hypothalamic ARC,

DMH, VMH, PVN and LH (as indicated by increased c-Fos-immunoreactivity in these nuclei),

the main hypothalamic regions involved in regulation of energy homeostasis. Furthermore, it was

132

Figure 32. Summary of the mechanisms activated by GLP-2 to regulate neurotensin and

ghrelin mRNA transcript expression in the mHypoE-36/1 neuronal cells. Long acting GLP-2

triggers classic cAMP-PKA signal transduction pathway to increase neurotensin and ghrelin

mRNA transcript levels in the novel hypothalamic GLP-2R-expressing neuronal cells. GLP-2

may trigger yet unknown transcriptional mechanisms and/or modify mRNA stability through

unidentified mechanisms to induce neurotensin and ghrelin mRNA levels in the hypothalamic

neuronal cells. NT, neurotensin, upward pointing solid green, yellow and black arrows represent

upregulation and downward pointing red arrow represents downregulation of mRNA transcript

levels or stability; solid blue arrows represent interactions by known mechanisms, whereas

dashed blue arrows and question marks represent yet to be identified mechanisms.

133

detected that central h(Gly2)GLP-2 significantly activated α-MSH/POMC neurons in the ARC,

NPY neurons in the ARC and DMH, neurotensin neurons in the ARC, DMH, VMH, PVN and

LH, and ghrelin neurons in the ARC, PVN and internuclear space between DMH and LH.

Ultimately, to investigate the direct effect of increased GLP-2R signaling on hypothalamic

neuropeptides, recently immortalized GLP-2R-positive adult mHypoA-2/30 hypothalamic

neuronal cell model was used. It was found that h(Gly2)GLP-2 significantly increased

neurotensin and ghrelin mRNA expression via PKA activation. Collectively, these novel findings

indicate that the transient anorexia induced by central GLP-2R activation involves complex

interactions between several hypothalamic nuclei, activation of feeding-related neuropeptidergic

neurons, and regulation of mRNA transcript levels of downstream mediators such as neurotensin

and ghrelin.

The findings from the first part of this thesis indicate that insulin and leptin regulate

proglucagon mRNA levels in both adult and embryonic hypothalamic neuronal cell models.

Indeed, this regulation may play an important role in appetite control, most probably in

mediating anorexigenic action of both hormones that act as adiposity signals in living organisms.

However, the differential regulation of proglucagon mRNA levels by insulin in adult versus

embryonic cells indicates that it may occur due to differential post-translational processing of

precursor polypeptide proglucagon that results in the generation of distinct PGDPs in adult and

embryonic hypothalamic neurons. As glucagon is one of the main PGDPs synthesized in the

embryonic hypothalamus (183, 184), the observed downregulation of proglucagon mRNA

transcript levels by insulin in the embryonic neuronal cells can be related to the counter-

regulatory actions of insulin on glucagon synthesis in embryonic hypothalamus quite similar to

the insulin-mediated downregulation of proglucagon gene expression and glucagon synthesis in

pancreatic α-cells (200, 201). Further, GLP-1, GLP-2 and oxyntomodulin are the main PGDPs

synthesized in the adult hypothalamus (172, 183, 333), therefore, the upregulation of

hypothalamic proglucagon mRNA by insulin in the adult hypothalamic neuronal cells reflects

similar regulation of proglucagon mRNA by insulin in enteroendocrine L-cells (199, 263). This

differential regulation of hypothalamic proglucagon in adult versus embryonic neurons must be

physiologically important. As the peripheral GLP-1 and GLP-2 play important roles in nutrient

absorption, metabolism, cyto- and entero-protection, the upregulation of hypothalamic

proglucagon mRNA transcripts by insulin and leptin could be involved in similar functions in the

134

CNS, such as central feeding regulation, neuroprotection and neuronal regeneration that need to

be investigated further. To partially address this issue, in the second part of this thesis, action of

GLP-1R and GLP-2R agonists on anorexia and activation of hypothalamic neuropeptidergic

neurons in mice were investigated. Both GLP-R agonists activate specific neuronal populations

in the hypothalamus during suppression of feeding. Furthermore, it was found that both GLP-R

agonists suppressed not only feeding but also water intake. In this context, it must be noted that

both GLP-R agonists activated the PVN, an important region of the hypothalamus that

synthesizes and secrets several major hormones involved in energy homeostasis such as CRH,

TRH, or vasopressin into the general circulation (50-52). Previously it was found that GLP-1

increased vasopressin and CRH levels in rats (441), however, it remains unknown whether these

PVN neurons were affected by both GLP-R agonists in mice used in the present study.

Another important finding is the differential regulation of grhelin by exendin-4 and

h(Gly2)GLP-2 during early time period post-treatment in the hypothalamic adult neurons.

Although both GLP-R agonists activate similar signaling proteins, it was found that only

exendin-4 suppressed ghrelin mRNA levels. A possible explanation for this differential action

can be the specific properties possessed by the GLP-1R. Unlike GLP-2R, the GLP-1R is

promiscuous, in that it has been shown to couple to multiple G proteins, activating multiple

signalling pathways. Although the GLP-1R is known to preferentially couple to Gαs, it can also

couple to Gαq proteins resulting in activation of phospholipase C, protein kinase C (PKC) and

mobilisation of intracellular Ca2+

(375, 442, 443). Thus, these findings indicate that

hypothalamic GLP-1R and GLP-2R activation can lead to differential regulation of hypothalamic

neuropeptides due to unique characteristics of these receptors.

It was found that at pharmacological concentrations, both GLP-1R and GLP-2R agonists

activate hypothalamic specific neuronal populations during anorexia. However, it is noteworthy

that the dose of h(Gly2)GLP-2 (5 µg/mouse) used to induce suppression of food intake in mice

was 50 times higher than the dose of exendin-4 (100 ng/mouse) to induce similar anorexigenic

action. This difference in the effectiveness of exendin-4 and h(Gly2)GLP-2 indicates that

although pharmacological intervention with the central GLP-2R potentially results in appetite

suppression, h(Gly2)GLP-2 does not appear to have potent anorexigenic properties that exendin-

4 possesses. Further, as both GLP-1 and GLP-2 are co-secreted in equimolar concentrations

135

under physiological conditions, based on our findings, it can be speculated that unlike GLP-1,

the contribution of GLP-2 to induce anorexia could be minimal or even negligible. However,

similar to the h(Gly2)GLP-2-activated hypothalamic neuronal phenotypes detected in this

research project, it is possible that endogenous GLP-2 also activates these neurons under normal

physiological conditions. Thus, it can be speculated that similar to its intestinotrophic actions in

the periphery, the central GLP-2 may exert neurotrophic functions in the CNS, such as neuronal

protection, regeneration, plasticity and memory formation via activation of neurotensin- and

ghrelin-expressing neurons, and therefore, further investigation in this area is warranted.

Currently, much remains unknown about the central mechanisms behind the anorectic

effects of GLP-1 and -2. It is known that food intake is regulated by two complementary central

drives: the homeostatic and hedonic pathways. The homeostatic pathway controls energy balance

by increasing the motivation to eat following depletion of energy reserves. In contrast, hedonic

or reward-based regulation can override the homeostatic pathway during periods of relative

energy abundance by increasing the desire to consume foods that are highly palatable. The

present research has focused primarily on the impact of GLP-1 and -2 on the homeostatic brain

circuits that include well established areas involved in metabolic control in the hypothalamus.

However, it is also possible that these receptors regulate mesolimbic reward system to control

food intake. Importantly, GLP-1R and -2R are expressed in key brain areas controlling reward

and motivated behaviors that include the ventral tegmental area (VTA) and the nucleus

accumbens (NAc) (216). The role of GLP-1R and -2R within these brain areas, however,

remains largely unstudied. Given the rapid and widespread use of the GLP-1R agonist exendin-4,

and its potential to cross the BBB and gain access to brain parenchyma (219), it is of

considerable interest to determine the function of GLP-1R as well as GLP-2R agonists in central

areas involved in hedonic feeding in addition to those involved in homeostatic and metabolic

control. Given that the CNS control of food intake involves cross communication between

classic homeostatic feeding (e.g. NTS, hypothalamus) and higher-order/hedonic nuclei (e.g.

VTA, NAc), it is possible that the intake suppression by GLP-1R and -2R agonists involves

action in homeostatic centers as well as modulation of the rewarding value of food via direct

interaction with the mesolimbic reward system.

136

Recently, it was found that the exendin-4-mediated inhibition of food reward could be

driven from VTA and NAc without inducing concurrent visceral sickness, malaise or locomotor

impairment (444). The findings that the activation of central GLP-1Rs suppresses food reward by

interacting with the mesolimbic reward system indicate an entirely novel mechanism by which

the GLP-1R stimulation affects feeding-oriented behavior (444). Further, it was found that the

brainstem GLP-1-producing proglucagon neurons send projections to both VTA and NAc, and

potentially contribute to the regulation of reward behavior (445, 446). Thus, these novel findings

suggest that central GLP-1Rs exert an impact on the mesolimbic reward system and play an

important role in reward-based food regulation. This novel mechanism by which the GLP-1R

stimulation affects feeding-oriented behavior, and also the mechanisms triggered by GLP-2R

activation in the regulation of hedonic feeding need to be investigated further.

To summarize, this thesis attempts to provide information on some of the unknown

mechanisms underlying hypothalamic proglucagon gene regulation and the action of GLP-1R

and GLP-2R activation on hypothalamic neuropeptides, such as neurotensin and ghrelin,

involved in homeostatic feeding regulation. Elucidation of the control mechanisms regulating

proglucagon gene expression and action of GLPs in the hypothalamus is critical to the

understanding of how energy homeostasis is regulated. Further, regulation of hypothalamic

proglucagon, neurotensin and ghrelin may play an important role in the regulation of reward-

based behavior and hedonic feeding that needs to be investigated further.

6.2 Limitations

Within this thesis, the mechanisms utilized by insulin and leptin to regulate hypothalamic

proglucagon, and by exendin-4 and h(Gly2)GLP-2 to regulate hypothalamic neuropeptides have

been investigated; however, the experimental models used in this thesis have several advantages

and limitations that must not be overlooked in order to avoid overstatement of conclusions that

may be drawn from these findings.

The current experimental models range widely in complexity from clonal, single cell

lines to complex animals. It is expected that the selection of experimental models should be

based on the aim of the experiment and careful consideration of the limitations of each model.

137

The use of immortalized cell lines is advantageous because they are easy to culture and maintain

indefinitely or at least for over several passages. Moreover, they are inexpensive relative to other

experimental models, and largely generate highly reproducible results. For example, ample

amounts of protein or mRNA for analysis of second messenger or gene expression can be readily

obtained, particularly when compared to primary cultures or animal models. Furthermore,

hypothalamic clonal, neuronal cell lines are derived from a single cell type, so there is no danger

of contamination by other confounding non-neuronal cell types. Another great advantage is that

the cell culture offers a controlled physiochemical environment to study specific cellular and

molecular mechanisms, ease of experimental designs and procedures, as well as shortened

experimental timescales that economizes time and other resources. In addition, smaller quantities

of reagents are needed for cell culture experiments that drastically reduce the cost of compounds

for experiments when compared to in vivo research. In the study of hypothalamic gene

regulation, clonal cell models are critical for gaining an understanding of direct hormone and

neuromodulator action on specific neuronal subpopulations and of the molecular events

underlying these actions. Despite all these advantages, cell lines have several limitations that

must be acknowledged to accurately interpret experimental results and avoid drawing premature

conclusions. There is no doubt that in vivo or ex vivo analysis is necessary in supplementation

with in vitro analysis, however, as it is inherent to any technique, in vivo or ex vivo

manipulations, such as bilateral stereotactic or electrical stimulation or lesioning, central

injections or analysis of brain slices could typically stimulate or destroy a wide range of

hypothalamic neuronal subtypes as well as activate or disrupt afferent or efferent neuronal

terminals, thereby producing erroneous results and contributing to the faulty findings. Also,

classical in vivo approaches may not be feasible to investigate any direct effect of an agent on

specific hypothalamic neuronal subtypes, or on neuropeptide gene transcription, synthesis, or

secretion, largely due to limited number of neurons belonging to a particular neuronal phenotype

and also due to the multitude of synaptic inputs received from other adjacent neurons. Although

rodent genetic models have recently been used to examine the consequences of eliminating

neuropeptides or ablating neurons that endogenously express these neuropeptides, these models

have limitations due to the challenges in creating hypothalamus-specific knockouts.

Wherever possible, we tried to use at least two cell models. We used adult and embryonic

mouse-derived hypothalamic cell models. We found that insulin exerted opposite actions in adult

138

versus embryonic neuronal cells. Compared to adult neuronal cells that exhibit the characteristics

of mature neurons, our embryonic neuronal models may not accurately represent fully-

differentiated neurons due to physiological and developmental differences. As the embryonic

cells were generated from primary hypothalamic cultures obtained from fetal mice on E15, E17

and E18, the signaling mechanisms may not be fully functional or the enzymes required for post-

translational processing may not be active. It is noteworthy that prenatal development in rodents

corresponds to the first and second trimesters of human pregnancy, whereas the first week of

neonatal life in rodents corresponds to the third trimester of human pregnancy (447, 448).

Therefore, care must be exercised when interpreting the results of studies that have used both

types of cell models. Further, to confirm our findings from the embryonic neuronal cells,

complementary investigations into a more physiologically relevant cellular context using mouse

embryos is required that, of course, remains very challenging at present.

Another critical issue is that the biological response of one neuronal cell type may differ

from that of another as functionally unique neuronal subpopulations are present in the

hypothalamus and have been identified for both NPY- and GnRH-expressing neurons (449-451).

Thus findings may differ between cell models although they may have derived from the same

organ. Furthermore, removing neurons from their native physiological environment places

obvious limitations on these cell models. Because these cells lack the synaptic afferent and

efferent connections that are critical to neuronal function in intact brain, it is also likely that they

adapt to the artificial neurochemical environment in which they are grown. These adaptations

may cause changes in their gene expression or metabolic profile, and thus overall phenotype,

especially after several passages. For this reason, experiments were conducted at a low passage

and within a narrow passage range.

Another important issue is that genomic incorporation of the SV40 T-antigen oncogene

may also alter the phenotype of immortalized cells. Utilizing short-hairpin RNA to acutely

knockdown T-antigen, preliminary work in the Belsham laboratory has demonstrated that T-

antigen enhances the basal activity of phospho-proteins, including Akt, JAK2, STAT3, and

AMPK, in several immortalized cell types. In immortalized hypothalamic neurons specifically,

an elevation in the endogenous expression of select neuropeptides, such as AgRP and oxytocin,

was observed (Belsham, unpublished data). To minimize the basal activity of these key signaling

139

proteins, a common practice in our laboratory has been to culture neurons in serum-free, low-

glucose medium for several hours prior to treatment. Nonetheless, further investigation needs to

be conducted in order to more fully elucidate the phenotypic changes induced by SV40T-antigen

incorporation and to minimize these effects for the accuracy of the data.

Within this thesis, inhibitors were used to analyze the role of specific signaling proteins

and pathways. The use of pharmacological antagonists is also associated with some important

considerations. Despite the specificity of the antagonists employed in this thesis, it is possible

that these pharmacological agents exert unintended or non-specific effects on the cells. For

example, LY294002 can potentially inhibit glycogen synthase kinase 3, polo-like kinase 1 and

casein kinase 2 at concentrations similar to those that inhibit PI3K and was recently shown to

bind a number of other ATP-binding proteins that are not protein kinases (452, 453). Thus, the

inhibitor studies do not fully rule out the involvement of other enzymes or proteins.

Nevertheless, wherever possible, at least two inhibitors with different mechanisms of action were

used in the inhibitor studies in order to minimize unintended and non-specific inhibition of other

proteins. Also use of two different inhibitors supports the final conclusions as it is less likely that

both would cause the same effect by chance. Future studies utilizing small interfering RNA

(siRNA) would assist in conclusively determining the role of the signaling proteins and pathways

studied in this thesis.

In this thesis, the embryonic hypothalamic cells were transfected with the presently

available human and rat proglucagon promoter constructs due to the current unavailability of

mouse proglucagon promoter reporter plasmids. The use of human or rat proglucagon promoter

constructs in this study can be justified by the fact that the proximal promoter elements are

highly conserved in the rodent and human proximal proglucagon promoter sequences (191). The

homology between human and rat proglucagon promoters shows that they are highly related,

however, there are several differences in the nucleotide sequence (191). The G1 promoter is

highly conserved as compared to the less well conserved G2-G4 enhancer sequences. These

elements have binding sites for several transcription factors including isl-1, cdx-2/3, Brn-4, pax-

6, CREB and AP-1, however putative STAT3 binding sites are not detected (185, 191). A 300 bp

sequence that contains G2, G3 and G4 enhancer elements of rat proglucagon promoter is

sufficient for proglucagon gene expression in the mouse pancreatic islet cells, therefore, two

140

reporter plasmids containing -476 to +58 and -312 to +58 from rat proglucagon promoter were

generated (191, 330). Previously, these promoter constructs were used to study proglucagon gene

expression in hamster islet cell line InR1-G9 and mouse islet cell line αTC-1 (330). As is evident

from Figure 12A, B, the basal activity of these plasmids was much lower than the basal activity

of the human proglucagon plasmids, and any significant activation of rat proglucagon promoter

plasmids by insulin or leptin was not detected. The low basal transcription activity of rat

promoters in the mouse cell line could be due to low sequence similarity between the rat and

mouse proglucagon 5’ regulatory regions. Second, the 5’ flanking sequences used for the

proglucagon constructs were only 312 and 476 bp long, and thus cis-acting responsive elements

in the neuronal proglucagon gene promoter could be located in the upstream GUE promoter

segment that is composed of multiple positive and negative cis-acting DNA regulatory elements

involved in regulating tissue-specific proglucagon gene transcription (190). Further, it should be

noted that although the CRE motif is 100% conserved among rodent species, in the human

proglucagon promoter this site is CAACGTCA instead of TGACGTCA (191), and it remains

unknown whether this motif in human proglucagon gene promoter is capable of interacting with

CREB. Furthermore, trans-activating elements, miRNAs and other factors necessary for gene

transcription could differ among these species or even cell types. Thus, the use of non-species-

specific promoters can confound the analysis, and the lack of insulin or leptin transcriptional

effects using the rat or human promoter in the mouse hypothalamic cell lines does not

conclusively exclude transcriptional regulation of mouse proglucagon gene. Generation of mouse

proglucagon promoter plasmids is required to conduct further studies.

In the in vivo studies of this thesis, only male mice were used for the in vivo experiments.

Further experiments using female mice must be conducted to rule out any sex-specific effects of

GLP-1R and -2R activation. Also it must be taken into consideration that there can be species-

specific differences in the activation of hypothalamic neuropeptides (454). Furthermore, the

present results must be interpreted with caution as these data may not be relevant to endogenous

GLP-1 action due to the distinct mechanisms of receptor activation by endogenous GLP-1 and

exendin-4 (163, 369, 373). The same may be true for endogenous and long-acting GLP-2

activation of the hypothalamic neurons. Finally, it should be recognized that co-expression of

activated c-Fos, neuropeptide, and GLP-1R or -2R in the hypothalamus could not be detected

due to the limitations of current in vivo methods of detection, such as multi-label

141

immunohistochemistry and in situ hybridization techniques. Another important issue is that the

detection of neuronal activity by c-Fos-ir does not provide information on neuronal stimulation

or inhibition, therefore, more advanced techniques such as magnetic resonance imaging (455), or

light-activated channels should be considered (456). Finally, this research project focused only

on the action of GLP-1R and -2R agonists on the hypothalamic neurons, however, as the post-

translational processing of a single precursor polypeptide proglucagon generates several related

peptides, similar experiments must be conducted to investigate action of other PGDPs, such as

glucagon, oxyntomodulin and glicentin, that have been shown to regulate gastric acid secretion

and gut motility (205, 457-459).

To detect neuronal activation by Exendin-4 and GLP-2 in mouse hypothalamus, we

investigated c-Fos activation as an established marker of neuronal activation. After stimulation,

c-Fos expression reaches its highest level within 60-90 min post-stimulation and persists for

about 2-4 hours (460). Thus, there is a very narrow period for the detection of activated

hypothalamic regions and neurons by using c-Fos-immunoreactivity as an indicator of neuronal

activation. This was the main reason for us to isolate the brains within 2 hours post-treatment,

but not over a longer time period.

While studying GLP-2 action on the hypothalamic neuropeptides, treatments over a 24 h

time course were performed using the GLP-2R-expressing cell model, however, significant

changes in mRNA expression at early time points were not observed as anticipated based on the

in vivo findings. Notwithstanding, many studies have found that changes or turnover in

hypothalamic neuropeptide mRNA may not necessarily be rapid and can take place over a longer

period of time. For example, leptin decreases food intake within one hour, but induces

neuropeptide mRNA changes in the rat hypothalamus at 48 h post-treatment (43). Another study

found that in gonadotropin-releasing hormone (GnRH)-expressing hypothalamic GT1-7 cells,

12-O-tetradecanoylphorbol-13-acetate, an activator of PKC, induces c-Fos mRNA within 30

minutes, stimulates robust and rapid secretion of GnRH within 2.5 minutes, and produces

changes in GnRH mRNA only by 16 h (461). Thus, there is a possibility that neuropeptide

synthesis and secretion may be affected at early time points in vivo or in vitro by GLP-2 that

needs to be investigated further by using sensitive detection methods.

142

Only a few previous studies have examined the effects of GLP-1 and -2 in the brain, and

therefore, the main aim of the present study was to map the mouse hypothalamic regions and

neuropeptidergic neurons activated by the anorexigenic dose of the long-acting GLP-1 and -2R

agonists, and accordingly to study the changes in the neuropeptide mRNA levels over a period of

24 hours in vitro using an endogenous GLP-1R and -2R expressing hypothalamic neuronal cell

models. These both in vivo and in vitro models are independent of each other and therefore there

does not seem to be a clear connection between the in vivo and in vitro experiments, however,

there may not actually be descrepencies between the findings from these two models based on

the c-Fos activation detected in both experimental models. Overall, the present study establishes

novel hypothalamic neuronal cell models and these findings will enhance further investigation of

the physiologic role of GLP-1 and -2 in the hypothalamus or other regions of the brain involved

in feeding regulation using complimentary in vivo and in vitro models.

6.3 Future directions

The first study presented in this thesis determined that insulin and leptin utilize the

PI3K/Akt and JAK2/STAT3 pathways to regulate proglucagon mRNA expression. In the same

study, the human 5’ regulatory regions containing ~ 829 bp sequences were utilized, and as

previously discussed, this may have excluded insulin or leptin regulatory regions in the distal

promoter. Utilizing luciferase reporter constructs with a larger sequence, upwards of 5000 bp of

the human, rat and mouse proglucagon 5’regulatory regions, might allow us to uncover specific

DNA binding sequences, where transcription factors activated by insulin and leptin could bind.

As it was found that insulin and leptin regulate mRNA stability, the possible role of

miRNA and mRNA-binding proteins in altered turnover of proglucagon mRNA needs to be

investigated further. Using in silico analysis, it was found that there are binding sites for several

miRNAs and mRNA-binding proteins (Figure 13). Some of the miRNAs, such as miRNA128, a

brain-enriched microRNA, and miRNA494, have been already found to act as negative

regulators of gene expression at post-transcriptional level. miRNAs regulate either the

degradation of their specific target mRNAs or the inhibition of mRNA translation (355, 356). To

confirm the role of miRNAs in insulin- and leptin-mediated regulation of mRNA stability,

miRNAs that have binding sites on proglucagon mRNA transcript must be studied to dissect

143

their role in proglucagon mRNA turnover. To this aim, it must be investigated whether miRNAs

are expressed in the hypothalamic cell models, and if expressed, whether their expression is

modulated by insulin and leptin, and also whether changes in their expression can be associated

with alteration in proglucagon mRNA levels. For example, miRNA128 or miRNA494

expression must be measured following insulin and leptin treatment of hypothalamic neuronal

cells, and also the expression of the predicted target proglucagon mRNA must be evaluated, in

search of a possible inverse correlation with miR-128 or miRNA494 expression. Ectopic

overexpression and knockdown of miRNA128 or miRNA494 also need to be performed, in order

to unequivocally assign the observed proglucagon mRNA variations solely to miRNA128 or

miRNA494 expression in hypothalamic cell models. Similar to miRNA study, it is necessary to

determine whether RNA-binding proteins are expressed in the hypothalamic cell models.

Further, using overexpression and knockdown strategy, role of RNA-binding proteins, such as

ELAVL2 or QKI that have been found to regulate gene expression in the nervous system (361-

363), can be investigated in insulin- or leptin-mediated regulation of proglucagon mRNA

stability in the hypothalamic cell models.

This study has shown that insulin regulates proglucagon through a PI3K/Akt-dependent

pathway. Using Western blot analysis with antibodies specific to signaling proteins downstream

of Akt, further study to determine insulin-activated downstream factors, such as ELK-1, c-fos, c-

Myc and Ets, that could be involved in proglucagon gene regulation needs to be conducted. Next,

the role of these insulin-induced signaling molecules in proglucagon regulation can be

determined using specific inhibitors or siRNA technology. By chromatin immunoprecipitation or

electrophoretic mobility shift assay, further research is required to identify proglucagon gene

promoter cis-regulatory elements for binding of stimulatory or inhibitory transcription factors

activated by insulin and leptin. Similar experiments can be designed for the promoter analysis

and detection of transcription factors that may be involved in the regulation of neurotensin and

ghrelin gene expression by GLP-1R and GLP-2R activation.

In the in vivo studies of this thesis, it was found that exendin-4 and h(Gly2)GLP-2

significantly activated hypothalamic neurons, particularly neurotensin- and ghrelin-expressing

neurons that may serve as potential hypothalamic targets of their anorexigenic action in vivo. As

there are a significant number of reports of the effects of exendin-4 that are apparently not GLP-

144

1R-mediated (163, 369, 373), it would be important to demonstrate that the effects of exendin-4

detected in this thesis can be blocked by a GLP-1 receptor antagonist. To address this issue, an

experiment can be designed using exendin-9-39, a GLP-1R antagonist, in wild-type mice, or

using the GLP-1R knockout mice to demonstrate that the activation of hypothalamic neurons is

exclusively via GLP-1R. Similarly, using the GLP-2R antagonist GLP-2(3-33) in vivo, or using

knock-down strategy it could demonstrated that h(Gly2)GLP-2 activates hypothalamic neurons

selectively via GLP-2R. Further, specificity of GLP-2 action via GLP-2R can be examined in

GLP-1R knockout mice or using exendin-9-39 to block GLP-1R activity.

There are some gaps between the data generated from the in vivo and in vitro experiments

presented in this thesis. Therefore these experiments need to be clearly tied together to fill in the

gaps. Specifically, although this thesis has identified hypothalamic neurons, particularly

neurotensin- and ghrelin-expressing neurons as potential hypothalamic targets of central GLP-1R

or GLP-2R activation, it is still not clear whether activation of these neuropeptide-expressing

neurons leads to significant changes in the neuropeptide synthesis and release that may be

necessary for mediating the downstream anorexigenic action of GLPs. Particularly, it remains

unknown whether activation (stimulation) of neurotensin neurons or inactivation (inhibition) of

ghrelin neurons, and thereby a possible increase in neurotensin or a decrease in ghrelin synthesis

and release, play any role in mediating the action of GLPs on feeding. To address these issues in

vivo, it is necessary to initially measure neurotensin and ghrelin mRNA levels or peptide

concentrations in the hypothalamus and cerebrospinal fluid or plasma to study how these

neuropeptides are affected by GLP-1R or GLP-2R stimulation in the intact hypothalamus. To

further examine the relative role of neurotensin or ghrelin in mediating central GLP-1/2R action,

a follow up experiment can be designed to study the effects of a neurotensin antiserum (NT-AS)

(462), Ntsr antagonist (SR48692) (463), or using a ghrelin knockout mouse model on the satiety

action of GLP-1R or GLP-2R activation (464). These experiments will confirm the role of

neurotensin and ghrelin as the potential downstream mediators of anorexigenic effect of GLP-1R

or GLP-2R stimulation that will lead to better understanding of the mechanisms of appetite

regulation.

It is possible that the anorexigenic action of centrally administered GLP-1R or GLP-2R

agonists in mice will be due to increased synthesis or secretion of hypothalamic neurotensin

145

and/or decreased synthesis or secretion of ghrelin. It is also anticipated that

immunoneutralization by NT-AS (i.c.v.) or Ntsr antagonism by SR 48692 (intraperitoneal) will

block the anorexigenic effect of the central GLP-1R or GLP-2R activation on feeding in mice.

These findings will confirm that neurotensin is involved in mediating anorexigenic action of

GLP-R activation, and further suggest that this neuropeptide is an important component of the

neural appetite-regulating circuitry directly involving central PGDPs, GLP-1R and GLP-2R.

Ghrelin may be linked to GLP-1R or GLP-2R action through the use of the ghrelin knockout

mouse model. Importantly, knockout mice often have secondary confounding characteristics that

may interfere with physiological function. The mice proposed above already have a neuronal

phenotype (involved in thermoregulation and sleep, and thus may not be optimal to use for this

study). Therefore, establishing the direct link to ghrelin may be more difficult than neurotensin,

since ghrelin mRNA expression is decreased by exendin-4. A similar strategy can be utilized to

investigate NPY/AgRP or POMC to link these neuropeptides directly to the anorexigenic action

of exendin-4 or h(Gly2)GLP-2 using inhibitors or knock-down strategy. As this thesis has

established GLP-1R- and GLP-2R-expressing neuronal cell models, similar cost-effective

experiments to quantify neuropeptide synthesis or secretion can be performed in vitro using the

hypothalamic cell models for which it is necessary to develop and establish ultra-sensitive

neuropeptide detection methods.

Since detection of c-Fos-ir is not well suited to investigate neuronal inhibition,

application of advanced techniques such as magnetic resonance imaging for direct and fast

detection of neuronal activation in the mouse brain (455), or more advanced techniques, such as

light-activated channels for studying brain function and circuitry, should be considered (456).

Together, these future experiments will confirm the downstream mediators of anorexigenic

action of central GLP-1R and GLP-2R activation. Understanding the mechanism of action of

PGDPs in the hypothalamus is important not only due to the use of long-acting GLP-1R and

GLP-2R agonists, and agents that prevent PGDP degradation as therapies for disease, but also to

more accurately target prevention and treatment strategies for obesity and ultimately prevent its

complications, such as type 2 diabetes.

146

Chapter 7

References

147

1. WHO 2012 World Health Statistics: A snapshot of global health. In: World Health

Organization

2. Haslam DW, James WP 2005 Obesity. Lancet 366:1197-1209

3. Freedman DM, Ron E, Ballard-Barbash R, Doody MM, Linet MS 2006 Body mass

index and all-cause mortality in a nationwide US cohort. Int J Obes (Lond) 30:822-829

4. Barsh GS, Farooqi IS, O'Rahilly S 2000 Genetics of body-weight regulation. Nature

404:644-651

5. Comuzzie AG 2002 The emerging pattern of the genetic contribution to human obesity.

Best Pract Res Clin Endocrinol Metab 16:611-621

6. Ahima RS, Osei SY 2001 Molecular regulation of eating behavior: new insights and

prospects for therapeutic strategies. Trends Mol Med 7:205-213

7. Drucker DJ 2002 Biological actions and therapeutic potential of the glucagon-like

peptides. Gastroenterology 122:531-544

8. Gelling RW, Morton GJ, Morrison CD, Niswender KD, Myers MG, Jr., Rhodes CJ,

Schwartz MW 2006 Insulin action in the brain contributes to glucose lowering during

insulin treatment of diabetes. Cell Metab 3:67-73

9. Morton GJ, Cummings DE, Baskin DG, Barsh GS, Schwartz MW 2006 Central

nervous system control of food intake and body weight. Nature 443:289-295

10. Schwartz MW, Woods SC, Porte D, Jr., Seeley RJ, Baskin DG 2000 Central nervous

system control of food intake. Nature 404:661-671

11. Woods SC, Lotter EC, McKay LD, Porte D, Jr. 1979 Chronic intracerebroventricular

infusion of insulin reduces food intake and body weight of baboons. Nature 282:503-505

12. Bruning JC, Gautam D, Burks DJ, Gillette J, Schubert M, Orban PC, Klein R,

Krone W, Muller-Wieland D, Kahn CR 2000 Role of brain insulin receptor in control

of body weight and reproduction. Science 289:2122-2125

13. Taguchi A, Wartschow LM, White MF 2007 Brain IRS2 signaling coordinates life

span and nutrient homeostasis. Science 317:369-372

14. Koch L, Wunderlich FT, Seibler J, Konner AC, Hampel B, Irlenbusch S, Brabant

G, Kahn CR, Schwenk F, Bruning JC 2008 Central insulin action regulates peripheral

glucose and fat metabolism in mice. J Clin Invest 118:2132-2147

15. Elmquist JK, Ahima RS, Elias CF, Flier JS, Saper CB 1998 Leptin activates distinct

projections from the dorsomedial and ventromedial hypothalamic nuclei. Proc Natl Acad

Sci U S A 95:741-746

148

16. Vaisse C, Halaas JL, Horvath CM, Darnell JE, Jr., Stoffel M, Friedman JM 1996

Leptin activation of Stat3 in the hypothalamus of wild-type and ob/ob mice but not db/db

mice. Nat Genet 14:95-97

17. Zhang Y, Proenca R, Maffei M, Barone M, Leopold L, Friedman JM 1994 Positional

cloning of the mouse obese gene and its human homologue. Nature 372:425-432

18. Belsham DD, Cai F, Cui H, Smukler SR, Salapatek AM, Shkreta L 2004 Generation

of a phenotypic array of hypothalamic neuronal cell models to study complex

neuroendocrine disorders. Endocrinology 145:393-400

19. Belsham DD, Fick LJ, Dalvi PS, Centeno ML, Chalmers JA, Lee PK, Wang Y,

Drucker DJ, Koletar MM 2009 Ciliary neurotrophic factor recruitment of glucagon-like

peptide-1 mediates neurogenesis, allowing immortalization of adult murine hypothalamic

neurons. Faseb J 23:4256-4265

20. Elmquist JK, Marcus JN 2003 Rethinking the central causes of diabetes. Nat Med

9:645-647

21. Woods SC, D'Alessio DA 2008 Central control of body weight and appetite. J Clin

Endocrinol Metab 93:S37-50

22. Kopelman PG 2000 Obesity as a medical problem. Nature 404:635-643

23. Nieuwenhuys R, Voogd J, Huijzen Cv 2008 The human central nervous system. 4th ed.

Berlin ; New York: Springer

24. Bernardis LL, Bellinger LL 1996 The lateral hypothalamic area revisited: ingestive

behavior. Neurosci Biobehav Rev 20:189-287

25. Gao Q, Horvath TL 2008 Neuronal control of energy homeostasis. FEBS Lett 582:132-

141

26. Turton MD, O'Shea D, Gunn I, Beak SA, Edwards CM, Meeran K, Choi SJ, Taylor

GM, Heath MM, Lambert PD, Wilding JP, Smith DM, Ghatei MA, Herbert J,

Bloom SR 1996 A role for glucagon-like peptide-1 in the central regulation of feeding.

Nature 379:69-72

27. Tang-Christensen M, Larsen PJ, Goke R, Fink-Jensen A, Jessop DS, Moller M,

Sheikh SP 1996 Central administration of GLP-1-(7-36) amide inhibits food and water

intake in rats. Am J Physiol 271:R848-856

28. Donahey JCK, van Dijk G, Woods SC, Seeley RJ 1998 Intraventricular GLP-1 reduces

short- but not long-term food intake or body weight in lean and obese rats. Brain

Research 779:75-83

29. Brightman MW, Broadwell RD 1976 The morphological approach to the study of

normal and abnormal brain permeability. Adv Exp Med Biol 69:41-54

149

30. Lindner D, Stichel J, Beck-Sickinger AG 2008 Molecular recognition of the NPY

hormone family by their receptors. Nutrition 24:907-917

31. Bagnol D, Lu XY, Kaelin CB, Day HE, Ollmann M, Gantz I, Akil H, Barsh GS,

Watson SJ 1999 Anatomy of an endogenous antagonist: relationship between Agouti-

related protein and proopiomelanocortin in brain. J Neurosci 19:RC26

32. Broberger C, Johansen J, Johansson C, Schalling M, Hokfelt T 1998 The

neuropeptide Y/agouti gene-related protein (AGRP) brain circuitry in normal, anorectic,

and monosodium glutamate-treated mice. Proc Natl Acad Sci U S A 95:15043-15048

33. Ollmann MM, Wilson BD, Yang YK, Kerns JA, Chen Y, Gantz I, Barsh GS 1997

Antagonism of central melanocortin receptors in vitro and in vivo by agouti-related

protein. Science 278:135-138

34. Gee CE, Chen CL, Roberts JL, Thompson R, Watson SJ 1983 Identification of

proopiomelanocortin neurones in rat hypothalamus by in situ cDNA-mRNA

hybridization. Nature 306:374-376

35. Mizuno TM, Kleopoulos SP, Bergen HT, Roberts JL, Priest CA, Mobbs CV 1998

Hypothalamic pro-opiomelanocortin mRNA is reduced by fasting in ob/ob and db/db

mice, but is stimulated by leptin. Diabetes 47:294-297

36. Schwartz MW, Seeley RJ, Woods SC, Weigle DS, Campfield LA, Burn P, Baskin

DG 1997 Leptin increases hypothalamic pro-opiomelanocortin mRNA expression in the

rostral arcuate nucleus. Diabetes 46:2119-2123

37. Thornton JE, Cheung CC, Clifton DK, Steiner RA 1997 Regulation of hypothalamic

proopiomelanocortin mRNA by leptin in ob/ob mice. Endocrinology 138:5063-5066

38. Coll AP, Farooqi IS, Challis BG, Yeo GS, O'Rahilly S 2004 Proopiomelanocortin and

energy balance: insights from human and murine genetics. J Clin Endocrinol Metab

89:2557-2562

39. Elmquist JK, Elias CF, Saper CB 1999 From lesions to leptin: hypothalamic control of

food intake and body weight. Neuron 22:221-232

40. Sawchenko PE 1998 Toward a new neurobiology of energy balance, appetite, and

obesity: the anatomists weigh in. J Comp Neurol 402:435-441

41. Kristensen P, Judge ME, Thim L, Ribel U, Christjansen KN, Wulff BS, Clausen JT,

Jensen PB, Madsen OD, Vrang N, Larsen PJ, Hastrup S 1998 Hypothalamic CART is

a new anorectic peptide regulated by leptin. Nature 393:72-76

42. van Houten M, Posner BI, Kopriwa BM, Brawer JR 1979 Insulin-binding sites in the

rat brain: in vivo localization to the circumventricular organs by quantitative

radioautography. Endocrinology 105:666-673

150

43. Schwartz MW, Seeley RJ, Campfield LA, Burn P, Baskin DG 1996 Identification of

targets of leptin action in rat hypothalamus. J Clin Invest 98:1101-1106

44. Gao Q, Wolfgang MJ, Neschen S, Morino K, Horvath TL, Shulman GI, Fu XY 2004

Disruption of neural signal transducer and activator of transcription 3 causes obesity,

diabetes, infertility, and thermal dysregulation. Proc Natl Acad Sci U S A 101:4661-4666

45. Kitamura T, Feng Y, Kitamura YI, Chua SC, Jr., Xu AW, Barsh GS, Rossetti L,

Accili D 2006 Forkhead protein FoxO1 mediates Agrp-dependent effects of leptin on

food intake. Nat Med 12:534-540

46. Belgardt BF, Husch A, Rother E, Ernst MB, Wunderlich FT, Hampel B, Klockener

T, Alessi D, Kloppenburg P, Bruning JC 2008 PDK1 deficiency in POMC-expressing

cells reveals FOXO1-dependent and -independent pathways in control of energy

homeostasis and stress response. Cell Metab 7:291-301

47. Bouret SG, Draper SJ, Simerly RB 2004 Formation of projection pathways from the

arcuate nucleus of the hypothalamus to hypothalamic regions implicated in the neural

control of feeding behavior in mice. J Neurosci 24:2797-2805

48. Sawchenko PE, Swanson LW 1983 The organization of forebrain afferents to the

paraventricular and supraoptic nuclei of the rat. J Comp Neurol 218:121-144

49. Swanson LW, Sawchenko PE 1983 Hypothalamic integration: organization of the

paraventricular and supraoptic nuclei. Annu Rev Neurosci 6:269-324

50. Guilleman R, Rosenberg, B. 1955 Humoral hypothalamic control of anterior pituitary: a

study with combined tissue cultures. Endocrinology 57:599-607

51. Saffran M, Schally, A.V. 1955 The release of corticotropin by anterior pituitary tissue in

vitro. Can J Biochem Physiol 33:408-415

52. Swanson LW 1986 Organization of mammalian neuroendocrine system. In: Bloom F ed.

Handbook of Physiology - The Nervous System. Baltimore: Waverly Press; 317-363

53. Hillebrand JJ, de Wied D, Adan RA 2002 Neuropeptides, food intake and body weight

regulation: a hypothalamic focus. Peptides 23:2283-2306

54. Leibowitz SF, Wortley KE 2004 Hypothalamic control of energy balance: different

peptides, different functions. Peptides 25:473-504

55. Olszewski PK, Levine AS 2004 Minireview: Characterization of influence of central

nociceptin/orphanin FQ on consummatory behavior. Endocrinology 145:2627-2632

56. Kahn D, Abrams GM, Zimmerman EA, Carraway R, Leeman SE 1980 Neurotensin

Neurons in the Rat Hypothalamus: An Immunocytochemical Study. Endocrinology

107:47-54

151

57. Wynne K, Stanley S, McGowan B, Bloom S 2005 Appetite control. J Endocrinol

184:291-318

58. Satoh N, Ogawa Y, Katsuura G, Hayase M, Tsuji T, Imagawa K, Yoshimasa Y,

Nishi S, Hosoda K, Nakao K 1997 The arcuate nucleus as a primary site of satiety effect

of leptin in rats. Neurosci Lett 224:149-152

59. Monnikes H, Heymann-Monnikes I, Tache Y 1992 CRF in the paraventricular nucleus

of the hypothalamus induces dose-related behavioral profile in rats. Brain Res 574:70-76

60. Legradi G, Lechan RM 1999 Agouti-Related Protein Containing Nerve Terminals

Innervate Thyrotropin-Releasing Hormone Neurons in the Hypothalamic Paraventricular

Nucleus. Endocrinology 140:3643-3652

61. Fekete C, Legradi G, Mihaly E, Huang Q-H, Tatro JB, Rand WM, Emerson CH,

Lechan RM 2000 alpha -Melanocyte-Stimulating Hormone Is Contained in Nerve

Terminals Innervating Thyrotropin-Releasing Hormone-Synthesizing Neurons in the

Hypothalamic Paraventricular Nucleus and Prevents Fasting-Induced Suppression of

Prothyrotropin-Releasing Hormone Gene Expression. J Neurosci 20:1550-1558

62. Fekete C, Sarkar S, Rand WM, Harney JW, Emerson CH, Bianco AC, Lechan RM

2002 Agouti-Related Protein (AGRP) Has a Central Inhibitory Action on the

Hypothalamic-Pituitary-Thyroid (HPT) Axis; Comparisons between the Effect of AGRP

and Neuropeptide Y on Energy Homeostasis and the HPT Axis. Endocrinology

143:3846-3853

63. Neary NM, Goldstone AP, Bloom SR 2004 Appetite regulation: from the gut to the

hypothalamus. Clin Endocrinol (Oxf) 60:153-160

64. Everitt BJ 1989 The coexistence of neuropeptide Y with other peptides and amines in

the central nervous system. In: Mutt V, Fuxe, K., Hokfelt, T., Lundberg, J. ed.

Neuropeptide Y. New York: Raven Press; 61-72

65. Finley JC, Lindstrom P, Petrusz P 1981 Immunocytochemical localization of beta-

endorphin-containing neurons in the rat brain. Neuroendocrinology 33:28-42

66. Khachaturian H, Lewis, M., E., Martin Schäfer M.,K.,H., Watson, S.,J. 1985

Anatomy of the CNS opioid systems. Trends in Neurosciences 8:111-119

67. Koylu EO, Couceyro PR, Lambert PD, Ling NC, DeSouza EB, Kuhar MJ 1997

Immunohistochemical localization of novel CART peptides in rat hypothalamus, pituitary

and adrenal gland. J Neuroendocrinol 9:823-833

68. Brobeck JR 1946 Mechanism of the development of obesity in animals with

hypothalamic lesions. Physiol Rev 26:541-559

69. Anand BK, Brobeck JR 1951 Hypothalamic control of food intake in rats and cats. Yale

J Biol Med 24:123-140

152

70. Powley TL, Opsahl, C.,H., Cox, J.,E., Weingarten, H.,P. 1980 The role of the

hypothalamus in energy homeostasis. In: Morgane PJ, Panksepp, J. ed. Handbook of the

Hypothalamus Part A: Behavioral Studies of the Hypothalamus. New York: Marcel

Dekker, Inc.; 211-298

71. Matsuda M, Liu Y, Mahankali S, Pu Y, Mahankali A, Wang J, DeFronzo RA, Fox

PT, Gao JH 1999 Altered hypothalamic function in response to glucose ingestion in

obese humans. Diabetes 48:1801-1806

72. Simpson KA, Martin NM, R. Bloom S 2009 Hypothalamic regulation of food intake

and clinical therapeutic applications. Arquivos Brasileiros de Endocrinologia &

Metabologia 53:120-128

73. Pelleymounter MA, Cullen MJ, Wellman CL 1995 Characteristics of BDNF-induced

weight loss. Experimental Neurology 131:229-238

74. Xu B, Goulding EH, Zang K, Cepoi D, Cone RD, Jones KR, Tecott LH, Reichardt

LF 2003 Brain-derived neurotrophic factor regulates energy balance downstream of

melanocortin-4 receptor. 6:736-742

75. Satoh N, Ogawa Y, Katsuura G, Tsuji T, Masuzaki H, Hiraoka J, Okazaki T,

Tamaki M, Hayase M, Yoshimasa Y, Nishi S, Hosoda K, Nakao K 1997

Pathophysiological significance of the obese gene product, leptin, in ventromedial

hypothalamus (VMH)-lesioned rats: evidence for loss of its satiety effect in VMH-

lesioned rats. Endocrinology 138:947-954

76. Jacobowitz DM, O'Donohue TL 1978 alpha-Melanocyte stimulating hormone:

immunohistochemical identification and mapping in neurons of rat brain. Proceedings of

the National Academy of Sciences of the United States of America 75:6300-6304

77. Kalra SP, Dube MG, Pu S, Xu B, Horvath TL, Kalra PS 1999 Interacting appetite-

regulating pathways in the hypothalamic regulation of body weight. Endocr Rev 20:68-

100

78. Bernardis LL, Bellinger LL 1987 The dorsomedial hypothalamic nucleus revisited:

1986 update. Brain Res 434:321-381

79. Elmquist JK, Ahima RS, Maratos-Flier E, Flier JS, Saper CB 1997 Leptin Activates

Neurons in Ventrobasal Hypothalamus and Brainstem. Endocrinology 138:839-842

80. Yokosuka M, Xu B, Pu S, Kalra PS, Kalra SP 1998 Neural substrates for leptin and

neuropeptide Y (NPY) interaction: hypothalamic sites associated with inhibition of NPY-

induced food intake. Physiology & Behavior 64:331-338

81. Kamegai J, Minami S, Sugihara H, Hasegawa O, Higuchi H, Wakabayashi I 1996

Growth hormone receptor gene is expressed in neuropeptide Y neurons in hypothalamic

arcuate nucleus of rats. Endocrinology 137:2109-2112

153

82. Chan Y, Steiner R, Clifton D 1996 Regulation of hypothalamic neuropeptide-Y neurons

by growth hormone in the rat. Endocrinology 137:1319-1325

83. Guan XM, Yu H, Trumbauer M, Frazier E, Van der Ploeg LH, Chen H 1998

Induction of neuropeptide Y expression in dorsomedial hypothalamus of diet-induced

obese mice. Neuroreport 9:3415-3419

84. Kesterson RA, Huszar D, Lynch CA, Simerly RB, Cone RD 1997 Induction of

Neuropeptide Y Gene Expression in the Dorsal Medial Hypothalamic Nucleus in Two

Models of the Agouti Obesity Syndrome. Mol Endocrinol 11:630-637

85. Li C, Chen P, Smith MS 1998 Neuropeptide Y (NPY) neurons in the arcuate nucleus

(ARH) and dorsomedial nucleus (DMH), areas activated during lactation, project to the

paraventricular nucleus of the hypothalamus (PVH). Regul Pept 75-76:93-100

86. de Lecea L, Kilduff TS, Peyron C, Gao X-B, Foye PE, Danielson PE, Fukuhara C,

Battenberg ELF, Gautvik VT, Bartlett FS, Frankel WN, van den Pol AN, Bloom FE,

Gautvik KM, Sutcliffe JG 1998 The hypocretins: Hypothalamus-specific peptides with

neuroexcitatory activity. Proceedings of the National Academy of Sciences of the United

States of America 95:322-327

87. Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams

SC, Richarson JA, Kozlowski GP, Wilson S, Arch JR, Buckingham RE, Haynes AC,

Carr SA, Annan RS, McNulty DE, Liu WS, Terrett JA, Elshourbagy NA, Bergsma

DJ, Yanagisawa M 1998 Orexins and orexin receptors: a family of hypothalamic

neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell 92:1

page following 696

88. Bittencourt JC, Presse F, Arias C, Peto C, Vaughan J, Nahon JL, Vale W,

Sawchenko PE 1992 The melanin-concentrating hormone system of the rat brain: an

immuno- and hybridization histochemical characterization. J Comp Neurol 319:218-245

89. Peyron C, Tighe DK, van den Pol AN, de Lecea L, Heller HC, Sutcliffe JG, Kilduff

TS 1998 Neurons Containing Hypocretin (Orexin) Project to Multiple Neuronal Systems.

J Neurosci 18:9996-10015

90. Chronwall BM, DiMaggio DA, Massari VJ, Pickel VM, Ruggiero DA, O'Donohue

TL 1985 The anatomy of neuropeptide-Y-containing neurons in rat brain. Neuroscience

15:1159-1181

91. Bai FL, Yamano M, Shiotani Y, Emson PC, Smith AD, Powell JF, Tohyama M 1985

An arcuato-paraventricular and -dorsomedial hypothalamic neuropeptide Y-containing

system which lacks noradrenaline in the rat. Brain Res 331:172-175

92. Herzog H, Hort YJ, Ball HJ, Hayes G, Shine J, Selbie LA 1992 Cloned human

neuropeptide Y receptor couples to two different second messenger systems. Proc Natl

Acad Sci U S A 89:5794-5798

154

93. Blomqvist AG, Herzog H 1997 Y-receptor subtypes--how many more? Trends Neurosci

20:294-298

94. Gehlert DR 1999 Role of hypothalamic neuropeptide Y in feeding and obesity.

Neuropeptides 33:329-338

95. Pesonen U, Huupponen R, Rouru J, Koulu M 1992 Hypothalamic neuropeptide

expression after food restriction in Zucker rats: evidence of persistent neuropeptide Y

gene activation. Brain Res Mol Brain Res 16:255-260

96. Malabu UH, McCarthy HD, McKibbin PE, Williams G 1992 Peripheral insulin

administration attenuates the increase in neuropeptide Y concentrations in the

hypothalamic arcuate nucleus of fasted rats. Peptides 13:1097-1102

97. Mizuno TM, Makimura H, Silverstein J, Roberts JL, Lopingco T, Mobbs CV 1999

Fasting regulates hypothalamic neuropeptide Y, agouti-related peptide, and

proopiomelanocortin in diabetic mice independent of changes in leptin or insulin.

Endocrinology 140:4551-4557

98. Beck B, Burlet A, Nicolas JP, Burlet C 1992 Unexpected regulation of hypothalamic

neuropeptide Y by food deprivation and refeeding in the Zucker rat. Life Sci 50:923-930

99. Brady LS, Smith MA, Gold PW, Herkenham M 1990 Altered expression of

hypothalamic neuropeptide mRNAs in food-restricted and food-deprived rats.

Neuroendocrinology 52:441-447

100. Sahu A, Sninsky CA, Phelps CP, Dube MG, Kalra PS, Kalra SP 1992 Neuropeptide

Y release from the paraventricular nucleus increases in association with hyperphagia in

streptozotocin-induced diabetic rats. Endocrinology 131:2979-2985

101. Kalra SP, Dube MG, Sahu A, Phelps CP, Kalra PS 1991 Neuropeptide Y secretion

increases in the paraventricular nucleus in association with increased appetite for food.

Proc Natl Acad Sci U S A 88:10931-10935

102. Jang M, Romsos DR 1998 Neuropeptide Y and corticotropin-releasing hormone

concentrations within specific hypothalamic regions of lean but not ob/ob mice respond

to food-deprivation and refeeding. J Nutr 128:2520-2525

103. Kalra SP, Kalra PS 2003 Neuropeptide Y: a physiological orexigen modulated by the

feedback action of ghrelin and leptin. Endocrine 22:49-56

104. Wang J, Leibowitz KL 1997 Central insulin inhibits hypothalamic galanin and

neuropeptide Y gene expression and peptide release in intact rats. Brain Res 777:231-236

105. Cowley MA, Smith RG, Diano S, Tschop M, Pronchuk N, Grove KL, Strasburger

CJ, Bidlingmaier M, Esterman M, Heiman ML, Garcia-Segura LM, Nillni EA,

Mendez P, Low MJ, Sotonyi P, Friedman JM, Liu H, Pinto S, Colmers WF, Cone

RD, Horvath TL 2003 The distribution and mechanism of action of ghrelin in the CNS

155

demonstrates a novel hypothalamic circuit regulating energy homeostasis. Neuron

37:649-661

106. Muroya S, Yada T, Shioda S, Takigawa M 1999 Glucose-sensitive neurons in the rat

arcuate nucleus contain neuropeptide Y. Neurosci Lett 264:113-116

107. Kohno D, Gao HZ, Muroya S, Kikuyama S, Yada T 2003 Ghrelin directly interacts

with neuropeptide-Y-containing neurons in the rat arcuate nucleus: Ca2+ signaling via

protein kinase A and N-type channel-dependent mechanisms and cross-talk with leptin

and orexin. Diabetes 52:948-956

108. Ceccatelli S, Cintra A, Hokfelt T, Fuxe K, Wikstrom AC, Gustafsson JA 1989

Coexistence of glucocorticoid receptor-like immunoreactivity with neuropeptides in the

hypothalamic paraventricular nucleus. Exp Brain Res 78:33-42

109. Larsen PJ, Jessop DS, Chowdrey HS, Lightman SL, Mikkelsen JD 1994 Chronic

administration of glucocorticoids directly upregulates prepro-neuropeptide Y and Y1-

receptor mRNA levels in the arcuate nucleus of the rat. J Neuroendocrinol 6:153-159

110. Sanacora G, Kershaw M, Finkelstein JA, White JD 1990 Increased hypothalamic

content of preproneuropeptide Y messenger ribonucleic acid in genetically obese Zucker

rats and its regulation by food deprivation. Endocrinology 127:730-737

111. Wilding JP, Gilbey SG, Bailey CJ, Batt RA, Williams G, Ghatei MA, Bloom SR

1993 Increased neuropeptide-Y messenger ribonucleic acid (mRNA) and decreased

neurotensin mRNA in the hypothalamus of the obese (ob/ob) mouse. Endocrinology

132:1939-1944

112. Luquet S, Perez FA, Hnasko TS, Palmiter RD 2005 NPY/AgRP neurons are essential

for feeding in adult mice but can be ablated in neonates. Science 310:683-685

113. Bewick GA, Gardiner JV, Dhillo WS, Kent AS, White NE, Webster Z, Ghatei MA,

Bloom SR 2005 Post-embryonic ablation of AgRP neurons in mice leads to a lean,

hypophagic phenotype. Faseb J 19:1680-1682

114. Clark JT, Kalra PS, Crowley WR, Kalra SP 1984 Neuropeptide Y and human

pancreatic polypeptide stimulate feeding behavior in rats. Endocrinology 115:427-429

115. Erickson JC, Clegg KE, Palmiter RD 1996 Sensitivity to leptin and susceptibility to

seizures of mice lacking neuropeptide Y. Nature 381:415-421

116. Erickson JC, Hollopeter G, Palmiter RD 1996 Attenuation of the obesity syndrome of

ob/ob mice by the loss of neuropeptide Y. Science 274:1704-1707

117. Coll AP, Loraine Tung YC 2009 Pro-opiomelanocortin (POMC)-derived peptides and

the regulation of energy homeostasis. Mol Cell Endocrinol 300:147-151

156

118. Ellacott KL, Cone RD 2004 The central melanocortin system and the integration of

short- and long-term regulators of energy homeostasis. Recent Prog Horm Res 59:395-

408

119. Krude H, Gruters A 2000 Implications of proopiomelanocortin (POMC) mutations in

humans: the POMC deficiency syndrome. Trends Endocrinol Metab 11:15-22

120. Krude H, Biebermann H, Luck W, Horn R, Brabant G, Gruters A 1998 Severe

early-onset obesity, adrenal insufficiency and red hair pigmentation caused by POMC

mutations in humans. Nat Genet 19:155-157

121. Krude H, Biebermann H, Schnabel D, Tansek MZ, Theunissen P, Mullis PE,

Gruters A 2003 Obesity due to proopiomelanocortin deficiency: three new cases and

treatment trials with thyroid hormone and ACTH4-10. J Clin Endocrinol Metab 88:4633-

4640

122. Yaswen L, Diehl N, Brennan MB, Hochgeschwender U 1999 Obesity in the mouse

model of pro-opiomelanocortin deficiency responds to peripheral melanocortin. Nat Med

5:1066-1070

123. Kishi T, Aschkenasi CJ, Lee CE, Mountjoy KG, Saper CB, Elmquist JK 2003

Expression of melanocortin 4 receptor mRNA in the central nervous system of the rat. J

Comp Neurol 457:213-235

124. Cone RD 2006 Studies on the physiological functions of the melanocortin system.

Endocr Rev 27:736-749

125. Tung YC, Piper SJ, Yeung D, O'Rahilly S, Coll AP 2006 A comparative study of the

central effects of specific proopiomelancortin (POMC)-derived melanocortin peptides on

food intake and body weight in pomc null mice. Endocrinology 147:5940-5947

126. Benoit SC, Schwartz MW, Lachey JL, Hagan MM, Rushing PA, Blake KA, Yagaloff

KA, Kurylko G, Franco L, Danhoo W, Seeley RJ 2000 A novel selective

melanocortin-4 receptor agonist reduces food intake in rats and mice without producing

aversive consequences. J Neurosci 20:3442-3448

127. Huszar D, Lynch CA, Fairchild-Huntress V, Dunmore JH, Fang Q, Berkemeier LR,

Gu W, Kesterson RA, Boston BA, Cone RD, Smith FJ, Campfield LA, Burn P, Lee

F 1997 Targeted disruption of the melanocortin-4 receptor results in obesity in mice. Cell

88:131-141

128. Yeo GS, Farooqi IS, Aminian S, Halsall DJ, Stanhope RG, O'Rahilly S 1998 A

frameshift mutation in MC4R associated with dominantly inherited human obesity. Nat

Genet 20:111-112

129. Farooqi IS, Keogh JM, Yeo GS, Lank EJ, Cheetham T, O'Rahilly S 2003 Clinical

spectrum of obesity and mutations in the melanocortin 4 receptor gene. N Engl J Med

348:1085-1095

157

130. Tao YX 2010 The melanocortin-4 receptor: physiology, pharmacology, and

pathophysiology. Endocr Rev 31:506-543

131. Sandoval DA, Bagnol D, Woods SC, D'Alessio DA, Seeley RJ 2008 Arcuate glucagon-

like peptide 1 receptors regulate glucose homeostasis but not food intake. Diabetes

57:2046-2054

132. Levine AS, Kneip J, Grace M, Morley JE 1983 Effect of centrally administered

neurotensin on multiple feeding paradigms. Pharmacol Biochem Behav 18:19-23

133. Luttinger D, King RA, Sheppard D, Strupp J, Nemeroff CB, Prange AJ, Jr. 1982

The effect of neurotensin on food consumption in the rat. Eur J Pharmacol 81:499-503

134. Dobner PR 2005 Multitasking with neurotensin in the central nervous system. Cell Mol

Life Sci 62:1946-1963

135. Alexander MJ, Leeman SE 1998 Widespread expression in adult rat forebrain of

mRNA encoding high-affinity neurotensin receptor. J Comp Neurol 402:475-500

136. Fassio A, Evans G, Grisshammer R, Bolam JP, Mimmack M, Emson PC 2000

Distribution of the neurotensin receptor NTS1 in the rat CNS studied using an amino-

terminal directed antibody. Neuropharmacology 39:1430-1442

137. Mazella J, Botto JM, Guillemare E, Coppola T, Sarret P, Vincent JP 1996 Structure,

functional expression, and cerebral localization of the levocabastine-sensitive

neurotensin/neuromedin N receptor from mouse brain. J Neurosci 16:5613-5620

138. Nicot A, Rostene W, Berod A 1994 Neurotensin receptor expression in the rat forebrain

and midbrain: a combined analysis by in situ hybridization and receptor autoradiography.

J Comp Neurol 341:407-419

139. Remaury A, Vita N, Gendreau S, Jung M, Arnone M, Poncelet M, Culouscou JM,

Le Fur G, Soubrie P, Caput D, Shire D, Kopf M, Ferrara P 2002 Targeted

inactivation of the neurotensin type 1 receptor reveals its role in body temperature control

and feeding behavior but not in analgesia. Brain Res 953:63-72

140. Ahima RS, Lazar MA 2008 Adipokines and the peripheral and neural control of energy

balance. Mol Endocrinol 22:1023-1031

141. Cui H, Cai F, Belsham DD 2005 Anorexigenic hormones leptin, insulin, and alpha-

melanocyte-stimulating hormone directly induce neurotensin (NT) gene expression in

novel NT-expressing cell models. J Neurosci 25:9497-9506

142. Kim ER, Leckstrom A, Mizuno TM 2008 Impaired anorectic effect of leptin in

neurotensin receptor 1-deficient mice. Behav Brain Res 194:66-71

143. Sahu A, Carraway RE, Wang YP 2001 Evidence that neurotensin mediates the central

effect of leptin on food intake in rat. Brain Res 888:343-347

158

144. Nowak A, Bojanowska E 2008 Effects of peripheral or central GLP-1 receptor blockade

on leptin-induced suppression of appetite. J Physiol Pharmacol 59:501-510

145. Young AA, Gedulin BR, Bhavsar S, Bodkin N, Jodka C, Hansen B, Denaro M 1999

Glucose-lowering and insulin-sensitizing actions of exendin-4: studies in obese diabetic

(ob/ob, db/db) mice, diabetic fatty Zucker rats, and diabetic rhesus monkeys (Macaca

mulatta). Diabetes 48:1026-1034

146. Gutierrez JA, Solenberg PJ, Perkins DR, Willency JA, Knierman MD, Jin Z,

Witcher DR, Luo S, Onyia JE, Hale JE 2008 Ghrelin octanoylation mediated by an

orphan lipid transferase. Proc Natl Acad Sci U S A 105:6320-6325

147. Yang J, Brown MS, Liang G, Grishin NV, Goldstein JL 2008 Identification of the

acyltransferase that octanoylates ghrelin, an appetite-stimulating peptide hormone. Cell

132:387-396

148. Kojima M, Hosoda H, Date Y, Nakazato M, Matsuo H, Kangawa K 1999 Ghrelin is a

growth-hormone-releasing acylated peptide from stomach. Nature 402:656-660

149. Zhang JV, Ren PG, Avsian-Kretchmer O, Luo CW, Rauch R, Klein C, Hsueh AJ

2005 Obestatin, a peptide encoded by the ghrelin gene, opposes ghrelin's effects on food

intake. Science 310:996-999

150. Nogueiras R, Pfluger P, Tovar S, Arnold M, Mitchell S, Morris A, Perez-Tilve D,

Vazquez MJ, Wiedmer P, Castaneda TR, DiMarchi R, Tschop M, Schurmann A,

Joost HG, Williams LM, Langhans W, Dieguez C 2007 Effects of obestatin on energy

balance and growth hormone secretion in rodents. Endocrinology 148:21-26

151. Holst B, Egerod KL, Schild E, Vickers SP, Cheetham S, Gerlach LO, Storjohann L,

Stidsen CE, Jones R, Beck-Sickinger AG, Schwartz TW 2007 GPR39 signaling is

stimulated by zinc ions but not by obestatin. Endocrinology 148:13-20

152. Nakazato M, Murakami N, Date Y, Kojima M, Matsuo H, Kangawa K, Matsukura

S 2001 A role for ghrelin in the central regulation of feeding. Nature 409:194-198

153. Wren AM, Small CJ, Ward HL, Murphy KG, Dakin CL, Taheri S, Kennedy AR,

Roberts GH, Morgan DG, Ghatei MA, Bloom SR 2000 The novel hypothalamic

peptide ghrelin stimulates food intake and growth hormone secretion. Endocrinology

141:4325-4328

154. Tschop M, Smiley DL, Heiman ML 2000 Ghrelin induces adiposity in rodents. Nature

407:908-913

155. Murakami N, Hayashida T, Kuroiwa T, Nakahara K, Ida T, Mondal MS, Nakazato

M, Kojima M, Kangawa K 2002 Role for central ghrelin in food intake and secretion

profile of stomach ghrelin in rats. J Endocrinol 174:283-288

159

156. Guan JL, Wang QP, Kageyama H, Takenoya F, Kita T, Matsuoka T, Funahashi H,

Shioda S 2003 Synaptic interactions between ghrelin- and neuropeptide Y-containing

neurons in the rat arcuate nucleus. Peptides 24:1921-1928

157. Sun Y, Ahmed S, Smith RG 2003 Deletion of ghrelin impairs neither growth nor

appetite. Mol Cell Biol 23:7973-7981

158. Wortley KE, Anderson KD, Garcia K, Murray JD, Malinova L, Liu R, Moncrieffe

M, Thabet K, Cox HJ, Yancopoulos GD, Wiegand SJ, Sleeman MW 2004 Genetic

deletion of ghrelin does not decrease food intake but influences metabolic fuel

preference. Proc Natl Acad Sci U S A 101:8227-8232

159. Wortley KE, del Rincon JP, Murray JD, Garcia K, Iida K, Thorner MO, Sleeman

MW 2005 Absence of ghrelin protects against early-onset obesity. J Clin Invest

115:3573-3578

160. Pfluger PT, Kirchner H, Gunnel S, Schrott B, Perez-Tilve D, Fu S, Benoit SC,

Horvath T, Joost HG, Wortley KE, Sleeman MW, Tschop MH 2008 Simultaneous

deletion of ghrelin and its receptor increases motor activity and energy expenditure. Am J

Physiol Gastrointest Liver Physiol 294:G610-618

161. Zigman JM, Nakano Y, Coppari R, Balthasar N, Marcus JN, Lee CE, Jones JE,

Deysher AE, Waxman AR, White RD, Williams TD, Lachey JL, Seeley RJ, Lowell

BB, Elmquist JK 2005 Mice lacking ghrelin receptors resist the development of diet-

induced obesity. J Clin Invest 115:3564-3572

162. Schusdziarra V, Zimmermann JP, Erdmann J, Bader U, Schick RR 2008

Differential inhibition of galanin- and ghrelin-induced food intake by i.c.v. GLP-1(7-36)-

amide. Regul Pept 147:29-32

163. Perez-Tilve D, Gonzalez-Matias L, Alvarez-Crespo M, Leiras R, Tovar S, Dieguez

C, Mallo F 2007 Exendin-4 potently decreases ghrelin levels in fasting rats. Diabetes

56:143-151

164. Soares JB, Roncon-Albuquerque R, Jr., Leite-Moreira A 2008 Ghrelin and ghrelin

receptor inhibitors: agents in the treatment of obesity. Expert Opin Ther Targets 12:1177-

1189

165. Bell GI, Sanchez-Pescador R, Laybourn PJ, Najarian RC 1983 Exon duplication and

divergence in the human preproglucagon gene. Nature 304:368-371

166. Bell GI, Santerre RF, Mullenbach GT 1983 Hamster preproglucagon contains the

sequence of glucagon and two related peptides. Nature 302:716-718

167. Lopez LC, Frazier ML, Su CJ, Kumar A, Saunders GF 1983 Mammalian pancreatic

preproglucagon contains three glucagon-related peptides. Proc Natl Acad Sci U S A

80:5485-5489

160

168. Heinrich G, Gros P, Habener JF 1984 Glucagon gene sequence. Four of six exons

encode separate functional domains of rat pre-proglucagon. J Biol Chem 259:14082-

14087

169. Irwin DM 2001 Molecular evolution of proglucagon. Regul Pept 98:1-12

170. Heinrich G, Gros P, Lund PK, Bentley RC, Habener JF 1984 Pre-proglucagon

messenger ribonucleic acid: nucleotide and encoded amino acid sequences of the rat

pancreatic complementary deoxyribonucleic acid. Endocrinology 115:2176-2181

171. Mojsov S, Heinrich G, Wilson IB, Ravazzola M, Orci L, Habener JF 1986

Preproglucagon gene expression in pancreas and intestine diversifies at the level of post-

translational processing. J Biol Chem 261:11880-11889

172. Drucker DJ, Asa S 1988 Glucagon gene expression in vertebrate brain. J Biol Chem

263:13475-13478

173. Rouille Y, Westermark G, Martin SK, Steiner DF 1994 Proglucagon is processed to

glucagon by prohormone convertase PC2 in alpha TC1-6 cells. Proc Natl Acad Sci U S A

91:3242-3246

174. Dhanvantari S, Seidah NG, Brubaker PL 1996 Role of prohormone convertases in the

tissue-specific processing of proglucagon. Mol Endocrinol 10:342-355

175. Brubaker PL, So DC, Drucker DJ 1989 Tissue-specific differences in the levels of

proglucagon-derived peptides in streptozotocin-induced diabetes. Endocrinology

124:3003-3009

176. Drucker DJ 2002 Gut adaptation and the glucagon-like peptides. Gut 50:428-435

177. Seidah NG, Chretien M 1999 Proprotein and prohormone convertases: a family of

subtilases generating diverse bioactive polypeptides. Brain Res 848:45-62

178. Canaff L, Bennett HP, Hendy GN 1999 Peptide hormone precursor processing: getting

sorted? Mol Cell Endocrinol 156:1-6

179. Van de Ven WJ, Roebroek AJ, Van Duijnhoven HL 1993 Structure and function of

eukaryotic proprotein processing enzymes of the subtilisin family of serine proteases. Crit

Rev Oncog 4:115-136

180. Rothenberg ME, Eilertson CD, Klein K, Zhou Y, Lindberg I, McDonald JK,

Mackin RB, Noe BD 1995 Processing of mouse proglucagon by recombinant

prohormone convertase 1 and immunopurified prohormone convertase 2 in vitro. J Biol

Chem 270:10136-10146

181. Rouille Y, Martin S, Steiner DF 1995 Differential processing of proglucagon by the

subtilisin-like prohormone convertases PC2 and PC3 to generate either glucagon or

glucagon-like peptide. J Biol Chem 270:26488-26496

161

182. Dhanvantari S, Brubaker PL 1998 Proglucagon processing in an islet cell line: effects

of PC1 overexpression and PC2 depletion. Endocrinology 139:1630-1637

183. Lui EY, Asa SL, Drucker DJ, Lee YC, Brubaker PL 1990 Glucagon and related

peptides in fetal rat hypothalamus in vivo and in vitro. Endocrinology 126:110-117

184. Lee YC, Brubaker PL, Drucker DJ 1990 Developmental and tissue-specific regulation

of proglucagon gene expression. Endocrinology 127:2217-2222

185. Jin T 2008 Mechanisms underlying proglucagon gene expression. J Endocrinol 198:17-

28

186. Philippe J, Drucker DJ, Chick WL, Habener JF 1987 Transcriptional regulation of

genes encoding insulin, glucagon, and angiotensinogen by sodium butyrate in a rat islet

cell line. Mol Cell Biol 7:560-563

187. Herzig S, Fuzesi L, Knepel W 2000 Heterodimeric Pbx-Prep1 homeodomain protein

binding to the glucagon gene restricting transcription in a cell type-dependent manner. J

Biol Chem 275:27989-27999

188. Efrat S, Teitelman G, Anwar M, Ruggiero D, Hanahan D 1988 Glucagon gene

regulatory region directs oncoprotein expression to neurons and pancreatic alpha cells.

Neuron 1:605-613

189. Lee YC, Asa SL, Drucker DJ 1992 Glucagon gene 5'-flanking sequences direct

expression of simian virus 40 large T antigen to the intestine, producing carcinoma of the

large bowel in transgenic mice. J Biol Chem 267:10705-10708

190. Jin T, Drucker DJ 1995 The proglucagon gene upstream enhancer contains positive and

negative domains important for tissue-specific proglucagon gene transcription. Mol

Endocrinol 9:1306-1320

191. Nian M, Drucker DJ, Irwin D 1999 Divergent regulation of human and rat proglucagon

gene promoters in vivo. Am J Physiol 277:G829-837

192. Baggio LL, Drucker DJ 2007 Biology of incretins: GLP-1 and GIP. Gastroenterology

132:2131-2157

193. Knepel W, Chafitz J, Habener JF 1990 Transcriptional activation of the rat glucagon

gene by the cyclic AMP-responsive element in pancreatic islet cells. Mol Cell Biol

10:6799-6804

194. Drucker DJ, Jin T, Asa SL, Young TA, Brubaker PL 1994 Activation of proglucagon

gene transcription by protein kinase-A in a novel mouse enteroendocrine cell line. Mol

Endocrinol 8:1646-1655

162

195. Jin T, Drucker DJ 1996 Activation of proglucagon gene transcription through a novel

promoter element by the caudal-related homeodomain protein cdx-2/3. Mol Cell Biol

16:19-28

196. Gevrey JC, Malapel M, Philippe J, Mithieux G, Chayvialle JA, Abello J, Cordier-

Bussat M 2004 Protein hydrolysates stimulate proglucagon gene transcription in

intestinal endocrine cells via two elements related to cyclic AMP response element.

Diabetologia 47:926-936

197. McKinnon CM, Ravier MA, Rutter GA 2006 FoxO1 is required for the regulation of

preproglucagon gene expression by insulin in pancreatic alphaTC1-9 cells. J Biol Chem

281:39358-39369

198. Yi F, Brubaker PL, Jin T 2005 TCF-4 mediates cell type-specific regulation of

proglucagon gene expression by beta-catenin and glycogen synthase kinase-3beta. J Biol

Chem 280:1457-1464

199. Yi F, Sun J, Lim GE, Fantus IG, Brubaker PL, Jin T 2008 Cross talk between the

insulin and Wnt signaling pathways: evidence from intestinal endocrine L cells.

Endocrinology 149:2341-2351

200. Philippe J 1989 Glucagon gene transcription is negatively regulated by insulin in a

hamster islet cell line. J Clin Invest 84:672-677

201. Philippe J 1991 Insulin regulation of the glucagon gene is mediated by an insulin-

responsive DNA element. Proc Natl Acad Sci U S A 88:7224-7227

202. Stobie-Hayes KM, Brubaker PL 1992 Control of proglucagon-derived peptide

synthesis and secretion in fetal rat hypothalamus. Neuroendocrinology 56:340-347

203. Stobie-Hayes KM, Brubaker PL 1995 Role of glutamate in regulating hypothalamic

proglucagon-derived peptide secretion in vitro. Life Sci 56:1325-1331

204. Wojdemann M, Wettergren A, Hartmann B, Holst JJ 1998 Glucagon-like peptide-2

inhibits centrally induced antral motility in pigs. Scand J Gastroenterol 33:828-832

205. Dakin CL, Gunn I, Small CJ, Edwards CM, Hay DL, Smith DM, Ghatei MA, Bloom

SR 2001 Oxyntomodulin inhibits food intake in the rat. Endocrinology 142:4244-4250

206. Druce MR, Bloom SR 2006 Oxyntomodulin : a novel potential treatment for obesity.

Treat Endocrinol 5:265-272

207. Munroe DG, Gupta AK, Kooshesh F, Vyas TB, Rizkalla G, Wang H, Demchyshyn

L, Yang ZJ, Kamboj RK, Chen H, McCallum K, Sumner-Smith M, Drucker DJ,

Crivici A 1999 Prototypic G protein-coupled receptor for the intestinotrophic factor

glucagon-like peptide 2. Proc Natl Acad Sci U S A 96:1569-1573

163

208. Eng J, Kleinman WA, Singh L, Singh G, Raufman JP 1992 Isolation and

characterization of exendin-4, an exendin-3 analogue, from Heloderma suspectum

venom. Further evidence for an exendin receptor on dispersed acini from guinea pig

pancreas. J Biol Chem 267:7402-7405

209. Kinzig KP, D'Alessio DA, Seeley RJ 2002 The diverse roles of specific GLP-1

receptors in the control of food intake and the response to visceral illness. J Neurosci

22:10470-10476

210. Meeran K, O'Shea D, Edwards CM, Turton MD, Heath MM, Gunn I, Abusnana S,

Rossi M, Small CJ, Goldstone AP, Taylor GM, Sunter D, Steere J, Choi SJ, Ghatei

MA, Bloom SR 1999 Repeated intracerebroventricular administration of glucagon-like

peptide-1-(7-36) amide or exendin-(9-39) alters body weight in the rat. Endocrinology

140:244-250

211. Brubaker PL 2006 The glucagon-like peptides: pleiotropic regulators of nutrient

homeostasis. Ann N Y Acad Sci 1070:10-26

212. Drucker DJ, Shi Q, Crivici A, Sumner-Smith M, Tavares W, Hill M, DeForest L,

Cooper S, Brubaker PL 1997 Regulation of the biological activity of glucagon-like

peptide 2 in vivo by dipeptidyl peptidase IV. Nat Biotechnol 15:673-677

213. Vrang N, Hansen M, Larsen PJ, Tang-Christensen M 2007 Characterization of

brainstem preproglucagon projections to the paraventricular and dorsomedial

hypothalamic nuclei. Brain Res 1149:118-126

214. Tang-Christensen M, Larsen PJ, Thulesen J, Romer J, Vrang N 2000 The

proglucagon-derived peptide, glucagon-like peptide-2, is a neurotransmitter involved in

the regulation of food intake. Nat Med 6:802-807

215. Larsen PJ, Tang-Christensen M, Holst JJ, Orskov C 1997 Distribution of glucagon-

like peptide-1 and other preproglucagon-derived peptides in the rat hypothalamus and

brainstem. Neuroscience 77:257-270

216. Merchenthaler I, Lane M, Shughrue P 1999 Distribution of pre-pro-glucagon and

glucagon-like peptide-1 receptor messenger RNAs in the rat central nervous system. J

Comp Neurol 403:261-280

217. Campos RV, Lee YC, Drucker DJ 1994 Divergent tissue-specific and developmental

expression of receptors for glucagon and glucagon-like peptide-1 in the mouse.

Endocrinology 134:2156-2164

218. Lovshin J, Estall J, Yusta B, Brown TJ, Drucker DJ 2001 Glucagon-like Peptide

(GLP)-2 Action in the Murine Central Nervous System Is Enhanced by Elimination of

GLP-1 Receptor Signaling. Journal of Biological Chemistry 276:21489-21499

219. Kastin AJ, Akerstrom V 2003 Entry of exendin-4 into brain is rapid but may be limited

at high doses. Int J Obes Relat Metab Disord 27:313-318

164

220. Kastin AJ, Akerstrom V, Pan W 2002 Interactions of glucagon-like peptide-1 (GLP-1)

with the blood-brain barrier. J Mol Neurosci 18:7-14

221. Baggio LL, Huang Q, Brown TJ, Drucker DJ 2004 Oxyntomodulin and glucagon-like

peptide-1 differentially regulate murine food intake and energy expenditure.

Gastroenterology 127:546-558

222. Buteau J, Roduit R, Susini S, Prentki M 1999 Glucagon-like peptide-1 promotes DNA

synthesis, activates phosphatidylinositol 3-kinase and increases transcription factor

pancreatic and duodenal homeobox gene 1 (PDX-1) DNA binding activity in beta (INS-

1)-cells. Diabetologia 42:856-864

223. Holz GG 2004 Epac: A new cAMP-binding protein in support of glucagon-like peptide-1

receptor-mediated signal transduction in the pancreatic beta-cell. Diabetes 53:5-13

224. Skoglund G, Hussain MA, Holz GG 2000 Glucagon-like peptide 1 stimulates insulin

gene promoter activity by protein kinase A-independent activation of the rat insulin I

gene cAMP response element. Diabetes 49:1156-1164

225. Bode HP, Moormann B, Dabew R, Goke B 1999 Glucagon-like peptide 1 elevates

cytosolic calcium in pancreatic beta-cells independently of protein kinase A.

Endocrinology 140:3919-3927

226. Gomez E, Pritchard C, Herbert TP 2002 cAMP-dependent protein kinase and Ca2+

influx through L-type voltage-gated calcium channels mediate Raf-independent

activation of extracellular regulated kinase in response to glucagon-like peptide-1 in

pancreatic beta-cells. J Biol Chem 277:48146-48151

227. Park S, Dong X, Fisher TL, Dunn S, Omer AK, Weir G, White MF 2006 Exendin-4

uses Irs2 signaling to mediate pancreatic beta cell growth and function. J Biol Chem

281:1159-1168

228. Maida A, Lovshin JA, Baggio LL, Drucker DJ 2008 The glucagon-like peptide-1

receptor agonist oxyntomodulin enhances beta-cell function but does not inhibit gastric

emptying in mice. Endocrinology 149:5670-5678

229. Dube PE, Brubaker PL 2007 Frontiers in glucagon-like peptide-2: multiple actions,

multiple mediators. Am J Physiol Endocrinol Metab 293:E460-465

230. Rowland KJ, Brubaker PL 2008 Life in the crypt: a role for glucagon-like peptide-2?

Mol Cell Endocrinol 288:63-70

231. Yusta B, Somwar R, Wang F, Munroe D, Grinstein S, Klip A, Drucker DJ 1999

Identification of glucagon-like peptide-2 (GLP-2)-activated signaling pathways in baby

hamster kidney fibroblasts expressing the rat GLP-2 receptor. J Biol Chem 274:30459-

30467

165

232. Burrin DG, Stoll B, Guan X, Cui L, Chang X, Hadsell D 2007 GLP-2 rapidly activates

divergent intracellular signaling pathways involved in intestinal cell survival and

proliferation in neonatal piglets. Am J Physiol Endocrinol Metab 292:E281-291

233. Dube PE, Rowland KJ, Brubaker PL 2008 Glucagon-like peptide-2 activates beta-

catenin signaling in the mouse intestinal crypt: role of insulin-like growth factor-I.

Endocrinology 149:291-301

234. Madadi G, Dalvi PS, Belsham DD 2008 Regulation of brain insulin mRNA by glucose

and glucagon-like peptide 1. Biochem Biophys Res Commun 376:694-699

235. Mayer CM, Fick LJ, Gingerich S, Belsham DD 2009 Hypothalamic cell lines to

investigate neuroendocrine control mechanisms. Front Neuroendocrinol 30:405-423

236. Banting FG, Best CH, Collip JB, Campbell WR, Fletcher AA 1991 Pancreatic

extracts in the treatment of diabetes mellitus: preliminary report. 1922. CMAJ 145:1281-

1286

237. Goodge KA, Hutton JC 2000 Translational regulation of proinsulin biosynthesis and

proinsulin conversion in the pancreatic beta-cell. Semin Cell Dev Biol 11:235-242

238. Wing SS 2008 The UPS in diabetes and obesity. BMC Biochem 9 Suppl 1:S6

239. Rorsman P, Renstrom E 2003 Insulin granule dynamics in pancreatic beta cells.

Diabetologia 46:1029-1045

240. Ullrich A, Bell JR, Chen EY, Herrera R, Petruzzelli LM, Dull TJ, Gray A, Coussens

L, Liao YC, Tsubokawa M, et al. 1985 Human insulin receptor and its relationship to

the tyrosine kinase family of oncogenes. Nature 313:756-761

241. Ebina Y, Ellis L, Jarnagin K, Edery M, Graf L, Clauser E, Ou JH, Masiarz F, Kan

YW, Goldfine ID, et al. 1985 The human insulin receptor cDNA: the structural basis for

hormone-activated transmembrane signalling. Cell 40:747-758

242. Czech MP 1985 The nature and regulation of the insulin receptor: structure and function.

Annu Rev Physiol 47:357-381

243. Ottensmeyer FP, Beniac DR, Luo RZ, Yip CC 2000 Mechanism of transmembrane

signaling: insulin binding and the insulin receptor. Biochemistry 39:12103-12112

244. Seino S, Seino M, Nishi S, Bell GI 1989 Structure of the human insulin receptor gene

and characterization of its promoter. Proc Natl Acad Sci U S A 86:114-118

245. Seino S, Bell GI 1989 Alternative splicing of human insulin receptor messenger RNA.

Biochem Biophys Res Commun 159:312-316

166

246. Pandini G, Frasca F, Mineo R, Sciacca L, Vigneri R, Belfiore A 2002 Insulin/insulin-

like growth factor I hybrid receptors have different biological characteristics depending

on the insulin receptor isoform involved. J Biol Chem 277:39684-39695

247. Belfiore A, Frasca F, Pandini G, Sciacca L, Vigneri R 2009 Insulin receptor isoforms

and insulin receptor/insulin-like growth factor receptor hybrids in physiology and

disease. Endocr Rev 30:586-623

248. Corp ES, Woods SC, Porte D, Jr., Dorsa DM, Figlewicz DP, Baskin DG 1986

Localization of 125I-insulin binding sites in the rat hypothalamus by quantitative

autoradiography. Neurosci Lett 70:17-22

249. Marks JL, Porte D, Jr., Stahl WL, Baskin DG 1990 Localization of insulin receptor

mRNA in rat brain by in situ hybridization. Endocrinology 127:3234-3236

250. Air EL, Benoit SC, Blake Smith KA, Clegg DJ, Woods SC 2002 Acute third

ventricular administration of insulin decreases food intake in two paradigms. Pharmacol

Biochem Behav 72:423-429

251. White MF, Maron R, Kahn CR 1985 Insulin rapidly stimulates tyrosine

phosphorylation of a Mr-185,000 protein in intact cells. Nature 318:183-186

252. Sun XJ, Rothenberg P, Kahn CR, Backer JM, Araki E, Wilden PA, Cahill DA,

Goldstein BJ, White MF 1991 Structure of the insulin receptor substrate IRS-1 defines a

unique signal transduction protein. Nature 352:73-77

253. Backer JM, Myers MG, Jr., Shoelson SE, Chin DJ, Sun XJ, Miralpeix M, Hu P,

Margolis B, Skolnik EY, Schlessinger J, et al. 1992 Phosphatidylinositol 3'-kinase is

activated by association with IRS-1 during insulin stimulation. EMBO J 11:3469-3479

254. Sarbassov DD, Guertin DA, Ali SM, Sabatini DM 2005 Phosphorylation and

regulation of Akt/PKB by the rictor-mTOR complex. Science 307:1098-1101

255. Iskandar K, Cao Y, Hayashi Y, Nakata M, Takano E, Yada T, Zhang C, Ogawa W,

Oki M, Chua S, Jr., Itoh H, Noda T, Kasuga M, Nakae J 2010 PDK-1/FoxO1

pathway in POMC neurons regulates Pomc expression and food intake. Am J Physiol

Endocrinol Metab 298:E787-798

256. Plum L, Belgardt BF, Bruning JC 2006 Central insulin action in energy and glucose

homeostasis. J Clin Invest 116:1761-1766

257. Pearson G, Robinson F, Beers Gibson T, Xu BE, Karandikar M, Berman K, Cobb

MH 2001 Mitogen-activated protein (MAP) kinase pathways: regulation and

physiological functions. Endocr Rev 22:153-183

258. Boulton TG, Nye SH, Robbins DJ, Ip NY, Radziejewska E, Morgenbesser SD,

DePinho RA, Panayotatos N, Cobb MH, Yancopoulos GD 1991 ERKs: a family of

167

protein-serine/threonine kinases that are activated and tyrosine phosphorylated in

response to insulin and NGF. Cell 65:663-675

259. Mirshamsi S, Laidlaw HA, Ning K, Anderson E, Burgess LA, Gray A, Sutherland

C, Ashford ML 2004 Leptin and insulin stimulation of signalling pathways in arcuate

nucleus neurones: PI3K dependent actin reorganization and KATP channel activation.

BMC Neurosci 5:54

260. Niswender KD, Morrison CD, Clegg DJ, Olson R, Baskin DG, Myers MG, Seeley

RJ, Schwartz MW 2003 Insulin Activation of Phosphatidylinositol 3-Kinase in the

Hypothalamic Arcuate Nucleus. Diabetes 52:227-231

261. Magnan C, Philippe J, Kassis N, Laury MC, Penicaud L, Gilbert M, Ktorza A 1995

In vivo effects of glucose and insulin on secretion and gene expression of glucagon in

rats. Endocrinology 136:5370-5376

262. Schinner S, Barthel A, Dellas C, Grzeskowiak R, Sharma SK, Oetjen E, Blume R,

Knepel W 2005 Protein kinase B activity is sufficient to mimic the effect of insulin on

glucagon gene transcription. J Biol Chem 280:7369-7376

263. Lim GE, Huang GJ, Flora N, LeRoith D, Rhodes CJ, Brubaker PL 2009 Insulin

regulates glucagon-like peptide-1 secretion from the enteroendocrine L cell.

Endocrinology 150:580-591

264. Islam D, Zhang N, Wang P, Li H, Brubaker PL, Gaisano HY, Wang Q, Jin T 2009

Epac is involved in cAMP-stimulated proglucagon expression and hormone production

but not hormone secretion in pancreatic alpha- and intestinal L-cell lines. Am J Physiol

Endocrinol Metab 296:E174-181

265. Lotfi S, Li Z, Sun J, Zuo Y, Lam PP, Kang Y, Rahimi M, Islam D, Wang P, Gaisano

HY, Jin T 2006 Role of the exchange protein directly activated by cyclic adenosine 5'-

monophosphate (Epac) pathway in regulating proglucagon gene expression in intestinal

endocrine L cells. Endocrinology 147:3727-3736

266. Montague CT, Farooqi IS, Whitehead JP, Soos MA, Rau H, Wareham NJ, Sewter

CP, Digby JE, Mohammed SN, Hurst JA, Cheetham CH, Earley AR, Barnett AH,

Prins JB, O'Rahilly S 1997 Congenital leptin deficiency is associated with severe early-

onset obesity in humans. Nature 387:903-908

267. Halaas JL, Gajiwala KS, Maffei M, Cohen SL, Chait BT, Rabinowitz D, Lallone

RL, Burley SK, Friedman JM 1995 Weight-reducing effects of the plasma protein

encoded by the obese gene. Science 269:543-546

268. Farooqi IS, Jebb SA, Langmack G, Lawrence E, Cheetham CH, Prentice AM,

Hughes IA, McCamish MA, O'Rahilly S 1999 Effects of recombinant leptin therapy in

a child with congenital leptin deficiency. N Engl J Med 341:879-884

168

269. Friedman JM 1998 Leptin, leptin receptors, and the control of body weight. Nutr Rev

56:s38-46; discussion s54-75

270. Considine RV, Sinha MK, Heiman ML, Kriauciunas A, Stephens TW, Nyce MR,

Ohannesian JP, Marco CC, McKee LJ, Bauer TL, et al. 1996 Serum immunoreactive-

leptin concentrations in normal-weight and obese humans. N Engl J Med 334:292-295

271. Mackintosh RM, Hirsch J 2001 The effects of leptin administration in non-obese

human subjects. Obes Res 9:462-469

272. Kelesidis T, Mantzoros CS 2006 The emerging role of leptin in humans. Pediatr

Endocrinol Rev 3:239-248

273. Bjorbaek C, Elmquist JK, Michl P, Ahima RS, van Bueren A, McCall AL, Flier JS

1998 Expression of leptin receptor isoforms in rat brain microvessels. Endocrinology

139:3485-3491

274. Fei H, Okano HJ, Li C, Lee GH, Zhao C, Darnell R, Friedman JM 1997 Anatomic

localization of alternatively spliced leptin receptors (Ob-R) in mouse brain and other

tissues. Proc Natl Acad Sci U S A 94:7001-7005

275. Mercer JG, Hoggard N, Williams LM, Lawrence CB, Hannah LT, Trayhurn P 1996

Localization of leptin receptor mRNA and the long form splice variant (Ob-Rb) in mouse

hypothalamus and adjacent brain regions by in situ hybridization. FEBS Lett 387:113-

116

276. Tartaglia LA 1997 The leptin receptor. J Biol Chem 272:6093-6096

277. Uotani S, Bjorbaek C, Tornoe J, Flier JS 1999 Functional properties of leptin receptor

isoforms: internalization and degradation of leptin and ligand-induced receptor

downregulation. Diabetes 48:279-286

278. Wang MY, Zhou YT, Newgard CB, Unger RH 1996 A novel leptin receptor isoform in

rat. FEBS Lett 392:87-90

279. Mercer JG, Hoggard N, Williams LM, Lawrence CB, Hannah LT, Morgan PJ,

Trayhurn P 1996 Coexpression of leptin receptor and preproneuropeptide Y mRNA in

arcuate nucleus of mouse hypothalamus. J Neuroendocrinol 8:733-735

280. Cowley MA, Smart JL, Rubinstein M, Cerdan MG, Diano S, Horvath TL, Cone RD,

Low MJ 2001 Leptin activates anorexigenic POMC neurons through a neural network in

the arcuate nucleus. Nature 411:480-484

281. Banks WA, Kastin AJ, Huang W, Jaspan JB, Maness LM 1996 Leptin enters the

brain by a saturable system independent of insulin. Peptides 17:305-311

169

282. Tang-Christensen M, Havel PJ, Jacobs RR, Larsen PJ, Cameron JL 1999 Central

administration of leptin inhibits food intake and activates the sympathetic nervous system

in rhesus macaques. J Clin Endocrinol Metab 84:711-717

283. Lee GH, Proenca R, Montez JM, Carroll KM, Darvishzadeh JG, Lee JI, Friedman

JM 1996 Abnormal splicing of the leptin receptor in diabetic mice. Nature 379:632-635

284. Margetic S, Gazzola C, Pegg GG, Hill RA 2002 Leptin: a review of its peripheral

actions and interactions. Int J Obes Relat Metab Disord 26:1407-1433

285. Balthasar N, Coppari R, McMinn J, Liu SM, Lee CE, Tang V, Kenny CD,

McGovern RA, Chua SC, Jr., Elmquist JK, Lowell BB 2004 Leptin receptor signaling

in POMC neurons is required for normal body weight homeostasis. Neuron 42:983-991

286. van de Wall E, Leshan R, Xu AW, Balthasar N, Coppari R, Liu SM, Jo YH,

MacKenzie RG, Allison DB, Dun NJ, Elmquist J, Lowell BB, Barsh GS, de Luca C,

Myers MG, Jr., Schwartz GJ, Chua SC, Jr. 2008 Collective and individual functions

of leptin receptor modulated neurons controlling metabolism and ingestion.

Endocrinology 149:1773-1785

287. Wilcke M, Walum E 2000 Characterization of leptin intracellular trafficking. Eur J

Histochem 44:325-334

288. Bjorbaek C, Uotani S, da Silva B, Flier JS 1997 Divergent signaling capacities of the

long and short isoforms of the leptin receptor. J Biol Chem 272:32686-32695

289. Li C, Friedman JM 1999 Leptin receptor activation of SH2 domain containing protein

tyrosine phosphatase 2 modulates Ob receptor signal transduction. Proc Natl Acad Sci U

S A 96:9677-9682

290. Friedman JM 1999 Leptin and the regulation of body weight. Harvey Lect 95:107-136

291. Banks AS, Davis SM, Bates SH, Myers MG, Jr. 2000 Activation of downstream

signals by the long form of the leptin receptor. J Biol Chem 275:14563-14572

292. Bates SH, Stearns WH, Dundon TA, Schubert M, Tso AW, Wang Y, Banks AS,

Lavery HJ, Haq AK, Maratos-Flier E, Neel BG, Schwartz MW, Myers MG, Jr. 2003

STAT3 signalling is required for leptin regulation of energy balance but not reproduction.

Nature 421:856-859

293. Bjorbaek C, Kahn BB 2004 Leptin signaling in the central nervous system and the

periphery. Recent Prog Horm Res 59:305-331

294. Niswender KD, Morton GJ, Stearns WH, Rhodes CJ, Myers MG, Jr., Schwartz

MW 2001 Intracellular signalling. Key enzyme in leptin-induced anorexia. Nature

413:794-795

170

295. Belgardt BF, Bruning JC 2010 CNS leptin and insulin action in the control of energy

homeostasis. Ann N Y Acad Sci 1212:97-113

296. Morris DL, Rui L 2009 Recent advances in understanding leptin signaling and leptin

resistance. Am J Physiol Endocrinol Metab 297:E1247-1259

297. Minokoshi Y, Alquier T, Furukawa N, Kim YB, Lee A, Xue B, Mu J, Foufelle F,

Ferre P, Birnbaum MJ, Stuck BJ, Kahn BB 2004 AMP-kinase regulates food intake

by responding to hormonal and nutrient signals in the hypothalamus. Nature 428:569-574

298. Bjorbaek C, Buchholz RM, Davis SM, Bates SH, Pierroz DD, Gu H, Neel BG, Myers

MG, Jr., Flier JS 2001 Divergent roles of SHP-2 in ERK activation by leptin receptors.

J Biol Chem 276:4747-4755

299. Rahmouni K, Sigmund CD, Haynes WG, Mark AL 2009 Hypothalamic ERK

mediates the anorectic and thermogenic sympathetic effects of leptin. Diabetes 58:536-

542

300. Cui H, Cai F, Belsham DD 2006 Leptin signaling in neurotensin neurons involves

STAT, MAP kinases ERK1/2, and p38 through c-Fos and ATF1. Faseb J 20:2654-2656

301. Elias CF, Kelly JF, Lee CE, Ahima RS, Drucker DJ, Saper CB, Elmquist JK 2000

Chemical characterization of leptin-activated neurons in the rat brain. J Comp Neurol

423:261-281

302. Goldstone AP, Morgan I, Mercer JG, Morgan DG, Moar KM, Ghatei MA, Bloom

SR 2000 Effect of leptin on hypothalamic GLP-1 peptide and brain-stem pre-

proglucagon mRNA. Biochem Biophys Res Commun 269:331-335

303. Goldstone AP, Mercer JG, Gunn I, Moar KM, Edwards CM, Rossi M, Howard JK,

Rasheed S, Turton MD, Small C, Heath MM, O'Shea D, Steere J, Meeran K, Ghatei

MA, Hoggard N, Bloom SR 1997 Leptin interacts with glucagon-like peptide-1 neurons

to reduce food intake and body weight in rodents. FEBS Lett 415:134-138

304. Bojanowska E, Nowak, A. 2007 Interactions between leptin and exendin-4, a glucagon-

like peptide-1 agonist, in the regulation of food intake in the rat. Journal of Physiology

and Pharmacology 58:349-360

305. Mercer JG, Moar KM, Findlay PA, Hoggard N, Adam CL 1998 Association of leptin

receptor (OB-Rb), NPY and GLP-1 gene expression in the ovine and murine brainstem.

Regul Pept 75-76:271-278

306. Vrang N, Larsen PJ, Jensen PB, Lykkegaard K, Artmann A, Larsen LK, Tang-

Christensen M 2008 Upregulation of the brainstem preproglucagon system in the obese

Zucker rat. Brain Res 1187:116-124

171

307. Huo L, Gamber KM, Grill HJ, Bjorbaek C 2008 Divergent leptin signaling in

proglucagon neurons of the nucleus of the solitary tract in mice and rats. Endocrinology

149:492-497

308. Marroqui L, Vieira E, Gonzalez A, Nadal A, Quesada I 2011 Leptin downregulates

expression of the gene encoding glucagon in alphaTC1-9 cells and mouse islets.

Diabetologia 54:843-851

309. Beck B 2000 Neuropeptides and obesity. Nutrition 16:916-923

310. Everitt BJ, Hokfelt T 1990 Neuroendocrine anatomy of the hypothalamus. Acta

Neurochir Suppl (Wien) 47:1-15

311. Cepko CL, Roberts BE, Mulligan RC 1984 Construction and applications of a highly

transmissible murine retrovirus shuttle vector. Cell 37:1053-1062

312. De Vitry F, Camier M, Czernichow P, Benda P, Cohen P, Tixier-Vidal A 1974

Establishment of a clone of mouse hypothalamic neurosecretory cells synthesizing

neurophysin and vasopressin. Proc Natl Acad Sci U S A 71:3575-3579

313. Cepko CL 1989 Immortalization of neural cells via retrovirus-mediated oncogene

transduction. Annu Rev Neurosci 12:47-65

314. Fick LJ, Belsham DD 2010 Nutrient sensing and insulin signaling in neuropeptide-

expressing immortalized, hypothalamic neurons: A cellular model of insulin resistance.

Cell Cycle 9:3186-3193

315. Belsham DD 2007 Hormonal regulation of clonal, immortalized hypothalamic neurons

expressing neuropeptides involved in reproduction and feeding. Mol Neurobiol 35:182-

194

316. Dalvi PS, Nazarians-Armavil A, Tung S, Belsham DD 2011 Immortalized neurons for

the study of hypothalamic function. Am J Physiol Regul Integr Comp Physiol

300:R1030-1052

317. Kasckow J, Mulchahey JJ, Aguilera G, Pisarska M, Nikodemova M, Chen HC,

Herman JP, Murphy EK, Liu Y, Rizvi TA, Dautzenberg FM, Sheriff S 2003

Corticotropin-releasing hormone (CRH) expression and protein kinase A mediated CRH

receptor signalling in an immortalized hypothalamic cell line. J Neuroendocrinol 15:521-

529

318. Rasmussen JE, Torres-Aleman I, MacLusky NJ, Naftolin F, Robbins RJ 1990 The

effects of estradiol on the growth patterns of estrogen receptor-positive hypothalamic cell

lines. Endocrinology 126:235-240

319. Kokoeva MV, Yin H, Flier JS 2005 Neurogenesis in the hypothalamus of adult mice:

potential role in energy balance. Science 310:679-683

172

320. Markakis EA, Palmer TD, Randolph-Moore L, Rakic P, Gage FH 2004 Novel

neuronal phenotypes from neural progenitor cells. J Neurosci 24:2886-2897

321. Pencea V, Bingaman KD, Wiegand SJ, Luskin MB 2001 Infusion of brain-derived

neurotrophic factor into the lateral ventricle of the adult rat leads to new neurons in the

parenchyma of the striatum, septum, thalamus, and hypothalamus. J Neurosci 21:6706-

6717

322. Kokoeva MV, Yin H, Flier JS 2007 Evidence for constitutive neural cell proliferation in

the adult murine hypothalamus. J Comp Neurol 505:209-220

323. Perez-Martin M, Cifuentes M, Grondona JM, Lopez-Avalos MD, Gomez-Pinedo U,

Garcia-Verdugo JM, Fernandez-Llebrez P 2010 IGF-I stimulates neurogenesis in the

hypothalamus of adult rats. Eur J Neurosci 31:1533-1548

324. Xu Y, Tamamaki N, Noda T, Kimura K, Itokazu Y, Matsumoto N, Dezawa M, Ide C

2005 Neurogenesis in the ependymal layer of the adult rat 3rd ventricle. Exp Neurol

192:251-264

325. Fineman MS, Bicsak TA, Shen LZ, Taylor K, Gaines E, Varns A, Kim D, Baron AD

2003 Effect on Glycemic Control of Exenatide (Synthetic Exendin-4) Additive to

Existing Metformin and/or Sulfonylurea Treatment in Patients With Type 2 Diabetes.

Diabetes Care 26:2370-2377

326. Jeppesen PB, Gilroy R, Pertkiewicz M, Allard JP, Messing B, O'Keefe SJ 2011

Randomised placebo-controlled trial of teduglutide in reducing parenteral nutrition and/or

intravenous fluid requirements in patients with short bowel syndrome. Gut

327. Hayes MR, Leichner TM, Zhao S, Lee GS, Chowansky A, Zimmer D, De Jonghe

BC, Kanoski SE, Grill HJ, Bence KK 2011 Intracellular signals mediating the food

intake-suppressive effects of hindbrain glucagon-like peptide-1 receptor activation. Cell

Metab 13:320-330

328. Shin ED, Estall JL, Izzo A, Drucker DJ, Brubaker PL 2005 Mucosal adaptation to

enteral nutrients is dependent on the physiologic actions of glucagon-like peptide-2 in

mice. Gastroenterology 128:1340-1353

329. Chomczynski P, Sacchi N 1987 Single-step method of RNA isolation by acid

guanidinium thiocyanate-phenol-chloroform extraction. Anal Biochem 162:156-159

330. Zhou L, Nian M, Gu J, Irwin DM 2006 Intron 1 sequences are required for pancreatic

expression of the human proglucagon gene. Am J Physiol Regul Integr Comp Physiol

290:R634-641

331. Paxinos G, Franklin KBJ 2003 The Mouse Brain in Stereotaxic Coordinates: Compact

Second Edition, Second Edition: Academic Press

332. Kieffer TJ, Habener JF 1999 The glucagon-like peptides. Endocr Rev 20:876-913

173

333. Vrang N, Larsen PJ 2010 Preproglucagon derived peptides GLP-1, GLP-2 and

oxyntomodulin in the CNS: role of peripherally secreted and centrally produced peptides.

Prog Neurobiol 92:442-462

334. Niswender KD, Baskin DG, Schwartz MW 2004 Insulin and its evolving partnership

with leptin in the hypothalamic control of energy homeostasis. Trends Endocrinol Metab

15:362-369

335. Benoit SC, Air EL, Coolen LM, Strauss R, Jackman A, Clegg DJ, Seeley RJ, Woods

SC 2002 The catabolic action of insulin in the brain is mediated by melanocortins. J

Neurosci 22:9048-9052

336. Sheriff S, Chance WT, Iqbal S, Rizvi TA, Xiao C, Kasckow JW, Balasubramaniam

A 2003 Hypothalamic administration of cAMP agonist/PKA activator inhibits both

schedule feeding and NPY-induced feeding in rats. Peptides 24:245-254

337. Hsieh Y-S, Yang S-F, Kuo D-Y 2007 Intracerebral administration of protein kinase A or

cAMP response element-binding protein antisense oligonucleotide can modulate

amphetamine-mediated appetite suppression in free-moving rats. American Journal of

Physiology - Endocrinology And Metabolism 292:E123-E131

338. Duan Z, Bradner J, Greenberg E, Mazitschek R, Foster R, Mahoney J, Seiden MV

2007 8-Benzyl-4-oxo-8-azabicyclo[3.2.1]oct-2-ene-6,7-dicarboxylic Acid (SD-1008), a

Novel Janus Kinase 2 Inhibitor, Increases Chemotherapy Sensitivity in Human Ovarian

Cancer Cells. Molecular Pharmacology 72:1137-1145

339. Blaskovich MA, Sun J, Cantor A, Turkson J, Jove R, Sebti SM 2003 Discovery of

JSI-124 (cucurbitacin I), a selective Janus kinase/signal transducer and activator of

transcription 3 signaling pathway inhibitor with potent antitumor activity against human

and murine cancer cells in mice. Cancer Res 63:1270-1279

340. Bensaude O 2011 Inhibiting eukaryotic transcription: Which compound to choose? How

to evaluate its activity? Transcription 2:103-108

341. Ljungman M 2007 The transcription stress response. Cell Cycle 6:2252-2257

342. Rusinov V, Baev V, Minkov IN, Tabler M 2005 MicroInspector: a web tool for

detection of miRNA binding sites in an RNA sequence. Nucleic Acids Res 33:W696-700

343. Cook KB, Kazan H, Zuberi K, Morris Q, Hughes TR 2011 RBPDB: a database of

RNA-binding specificities. Nucleic Acids Res 39:D301-308

344. Gonzalez M, Boer U, Dickel C, Quentin T, Cierny I, Oetjen E, Knepel W 2008 Loss

of insulin-induced inhibition of glucagon gene transcription in hamster pancreatic islet

alpha cells by long-term insulin exposure. Diabetologia 51:2012-2021

345. Philippe J, Drucker DJ, Habener JF 1987 Glucagon gene transcription in an islet cell

line is regulated via a protein kinase C-activated pathway. J Biol Chem 262:1823-1828

174

346. Philippe J 1991 Structure and pancreatic expression of the insulin and glucagon genes.

Endocr Rev 12:252-271

347. Drucker DJ 2003 Glucagon gene expression. In: Henry HL, Norman, AW, ed.

Encyclopedia of Hormones. San Diego, CA: Academic; 47-55

348. Brubaker PL, Schloos J, Drucker DJ 1998 Regulation of glucagon-like peptide-1

synthesis and secretion in the GLUTag enteroendocrine cell line. Endocrinology

139:4108-4114

349. Tsai B, Yue S, Irwin DM 2007 A novel element regulates expression of the proximal

human proglucagon promoter in islet cells. Gen Comp Endocrinol 151:230-239

350. Drucker DJ, Brubaker PL 1989 Proglucagon gene expression is regulated by a cyclic

AMP-dependent pathway in rat intestine. Proc Natl Acad Sci U S A 86:3953-3957

351. Wang J, Cao Y, Steiner DF 2003 Regulation of proglucagon transcription by activated

transcription factor (ATF) 3 and a novel isoform, ATF3b, through the cAMP-response

element/ATF site of the proglucagon gene promoter. J Biol Chem 278:32899-32904

352. Blencowe B, Brenner S, Hughes T, Morris Q 2009 Post-transcriptional gene

regulation: RNA-protein interactions, RNA processing, mRNA stability and localization.

Pac Symp Biocomput:545-548

353. Jing Q, Huang S, Guth S, Zarubin T, Motoyama A, Chen J, Di Padova F, Lin SC,

Gram H, Han J 2005 Involvement of microRNA in AU-rich element-mediated mRNA

instability. Cell 120:623-634

354. Kato M, Slack FJ 2008 microRNAs: small molecules with big roles - C. elegans to

human cancer. Biol Cell 100:71-81

355. Filipowicz W, Bhattacharyya SN, Sonenberg N 2008 Mechanisms of post-

transcriptional regulation by microRNAs: are the answers in sight? Nat Rev Genet 9:102-

114

356. Eulalio A, Huntzinger E, Izaurralde E 2008 Getting to the root of miRNA-mediated

gene silencing. Cell 132:9-14

357. Evangelisti C, Florian MC, Massimi I, Dominici C, Giannini G, Galardi S, Bue MC,

Massalini S, McDowell HP, Messi E, Gulino A, Farace MG, Ciafre SA 2009 MiR-128

up-regulation inhibits Reelin and DCX expression and reduces neuroblastoma cell

motility and invasiveness. Faseb J 23:4276-4287

358. Liu L, Jiang Y, Zhang H, Greenlee AR, Han Z 2010 Overexpressed miR-494 down-

regulates PTEN gene expression in cells transformed by anti-benzo(a)pyrene-trans-7,8-

dihydrodiol-9,10-epoxide. Life Sci 86:192-198

175

359. Granjon A, Gustin MP, Rieusset J, Lefai E, Meugnier E, Guller I, Cerutti C,

Paultre C, Disse E, Rabasa-Lhoret R, Laville M, Vidal H, Rome S 2009 The

microRNA signature in response to insulin reveals its implication in the transcriptional

action of insulin in human skeletal muscle and the role of a sterol regulatory element-

binding protein-1c/myocyte enhancer factor 2C pathway. Diabetes 58:2555-2564

360. Hamrick MW, Herberg S, Arounleut P, He HZ, Shiver A, Qi RQ, Zhou L, Isales

CM, Mi QS 2010 The adipokine leptin increases skeletal muscle mass and significantly

alters skeletal muscle miRNA expression profile in aged mice. Biochem Biophys Res

Commun 400:379-383

361. Aberg K, Saetre P, Jareborg N, Jazin E 2006 Human QKI, a potential regulator of

mRNA expression of human oligodendrocyte-related genes involved in schizophrenia.

Proc Natl Acad Sci U S A 103:7482-7487

362. Yano M, Okano HJ, Okano H 2005 Involvement of Hu and heterogeneous nuclear

ribonucleoprotein K in neuronal differentiation through p21 mRNA post-transcriptional

regulation. J Biol Chem 280:12690-12699

363. King PH, Levine TD, Fremeau RT, Jr., Keene JD 1994 Mammalian homologs of

Drosophila ELAV localized to a neuronal subset can bind in vitro to the 3' UTR of

mRNA encoding the Id transcriptional repressor. J Neurosci 14:1943-1952

364. Kreymann B, Williams G, Ghatei MA, Bloom SR 1987 Glucagon-like peptide-1 7-36:

a physiological incretin in man. Lancet 2:1300-1304

365. Drucker DJ 2001 Minireview: the glucagon-like peptides. Endocrinology 142:521-527

366. Kendall DM, Riddle MC, Rosenstock J, Zhuang D, Kim DD, Fineman MS, Baron

AD 2005 Effects of Exenatide (Exendin-4) on Glycemic Control Over 30 Weeks in

Patients With Type 2 Diabetes Treated With Metformin and a Sulfonylurea. Diabetes

Care 28:1083-1091

367. Kim D, MacConell L, Zhuang D, Kothare PA, Trautmann M, Fineman M, Taylor K

2007 Effects of once-weekly dosing of a long-acting release formulation of exenatide on

glucose control and body weight in subjects with type 2 diabetes. Diabetes Care 30:1487-

1493

368. Raz I, Hanefeld M, Xu L, Caria C, Williams-Herman D, Khatami H 2006 Efficacy

and safety of the dipeptidyl peptidase-4 inhibitor sitagliptin as monotherapy in patients

with type 2 diabetes mellitus. Diabetologia 49:2564-2571

369. Barrera JG, D'Alessio DA, Drucker DJ, Woods SC, Seeley RJ 2009 Differences in

the central anorectic effects of glucagon-like peptide-1 and exendin-4 in rats. Diabetes

58:2820-2827

370. Perfetti R, Merkel P 2000 Glucagon-like peptide-1: a major regulator of pancreatic

beta-cell function. Eur J Endocrinol 143:717-725

176

371. Baggio LL, Huang Q, Cao X, Drucker DJ 2008 An albumin-exendin-4 conjugate

engages central and peripheral circuits regulating murine energy and glucose

homeostasis. Gastroenterology 134:1137-1147

372. Dalvi PS, Nazarians-Armavil A, Purser MJ, Belsham DD 2012 Glucagon-Like

Peptide-1 Receptor Agonist, Exendin-4, Regulates Feeding-Associated Neuropeptides in

Hypothalamic Neurons in Vivo and in Vitro. Endocrinology 153:2208-2222

373. Nishizawa M, Nakabayashi H, Kawai K, Ito T, Kawakami S, Nakagawa A, Niijima

A, Uchida K 2000 The hepatic vagal reception of intraportal GLP-1 is via receptor

different from the pancreatic GLP-1 receptor. J Auton Nerv Syst 80:14-21

374. Yang H, Egan JM, Wang Y, Moyes CD, Roth J, Montrose MH, Montrose-Rafizadeh

C 1998 GLP-1 action in L6 myotubes is via a receptor different from the pancreatic GLP-

1 receptor. Am J Physiol 275:C675-683

375. Montrose-Rafizadeh C, Yang H, Wang Y, Roth J, Montrose MH, Adams LG 1997

Novel signal transduction and peptide specificity of glucagon-like peptide receptor in

3T3-L1 adipocytes. J Cell Physiol 172:275-283

376. Malendowicz LK, Nowak KW 2002 Preproglucagon derived peptides and thyrotropin

(TSH) secretion in the rat: robust and sustained lowering of blood TSH levels in exendin-

4 injected animals. Int J Mol Med 10:327-331

377. Ferrini F, Salio C, Lossi L, Merighi A 2009 Ghrelin in central neurons. Current

neuropharmacology 7:37-49

378. Lu S, Guan JL, Wang QP, Uehara K, Yamada S, Goto N, Date Y, Nakazato M,

Kojima M, Kangawa K, Shioda S 2002 Immunocytochemical observation of ghrelin-

containing neurons in the rat arcuate nucleus. Neurosci Lett 321:157-160

379. Sato T, Fukue Y, Teranishi H, Yoshida Y, Kojima M 2005 Molecular forms of

hypothalamic ghrelin and its regulation by fasting and 2-deoxy-d-glucose administration.

Endocrinology 146:2510-2516

380. Hori Y, Kageyama H, Guan JL, Kohno D, Yada T, Takenoya F, Nonaka N,

Kangawa K, Shioda S, Yoshida T 2008 Synaptic interaction between ghrelin- and

ghrelin-containing neurons in the rat hypothalamus. Regul Pept 145:122-127

381. Pugazhenthi U, Velmurugan K, Tran A, Mahaffey G, Pugazhenthi S 2010 Anti-

inflammatory action of exendin-4 in human islets is enhanced by phosphodiesterase

inhibitors: potential therapeutic benefits in diabetic patients. Diabetologia 53:2357-2368

382. Kim S, Moon M, Park S 2009 Exendin-4 protects dopaminergic neurons by inhibition

of microglial activation and matrix metalloproteinase-3 expression in an animal model of

Parkinson's disease. J Endocrinol 202:431-439

177

383. Ersahin M, Toklu HZ, Erzik C, Cetinel S, Akakin D, Velioglu-Ogunc A, Tetik S,

Ozdemir ZN, Sener G, Yegen BC 2010 The anti-inflammatory and neuroprotective

effects of ghrelin in subarachnoid hemorrhage-induced oxidative brain damage in rats. J

Neurotrauma 27:1143-1155

384. Kageyama K, Kumata Y, Akimoto K, Takayasu S, Tamasawa N, Suda T 2011

Ghrelin stimulates corticotropin-releasing factor and vasopressin gene expression in rat

hypothalamic 4B cells. Stress

385. Rowe W, Viau V, Meaney MJ, Quirion R 1992 Central Administration of Neurotensin

Stimulates Hypothalamic-Pituitary-Adrenal Activity. Annals of the New York Academy

of Sciences 668:365-367

386. Gil-Lozano M, Perez-Tilve D, Alvarez-Crespo M, Martis A, Fernandez AM,

Catalina PA, Gonzalez-Matias LC, Mallo F 2010 GLP-1(7-36)-amide and Exendin-4

stimulate the HPA axis in rodents and humans. Endocrinology 151:2629-2640

387. MacLusky NJ, Cook S, Scrocchi L, Shin J, Kim J, Vaccarino F, Asa SL, Drucker DJ

2000 Neuroendocrine function and response to stress in mice with complete disruption of

glucagon-like peptide-1 receptor signaling. Endocrinology 141:752-762

388. Fu LY, van den Pol AN 2010 Kisspeptin directly excites anorexigenic

proopiomelanocortin neurons but inhibits orexigenic neuropeptide Y cells by an indirect

synaptic mechanism. J Neurosci 30:10205-10219

389. Qi Y, Henry BA, Oldfield BJ, Clarke IJ 2010 The action of leptin on appetite-

regulating cells in the ovine hypothalamus: demonstration of direct action in the absence

of the arcuate nucleus. Endocrinology 151:2106-2116

390. Whitnall MH 1988 Distributions of pro-vasopressin expressing and pro-vasopressin

deficient CRH neurons in the paraventricular hypothalamic nucleus of colchicine-treated

normal and adrenalectomized rats. J Comp Neurol 275:13-28

391. Sawchenko PE, Pfeiffer SW 1988 Ultrastructural localization of neuropeptide Y and

galanin immunoreactivity in the paraventricular nucleus of the hypothalamus in the rat.

Brain Res 474:231-245

392. Hoosein NM, Gurd RS 1984 Human glucagon-like peptides 1 and 2 activate rat brain

adenylate cyclase. FEBS Lett 178:83-86

393. Ma X, Bruning J, Ashcroft FM 2007 Glucagon-like peptide 1 stimulates hypothalamic

proopiomelanocortin neurons. J Neurosci 27:7125-7129

394. Dominguez G, Kuhar MJ 2004 Transcriptional regulation of the CART promoter in

CATH.a cells. Brain Res Mol Brain Res 126:22-29

178

395. Fick LJ, Cai F, Belsham DD 2009 Hypothalamic preproghrelin gene expression is

repressed by insulin via both PI3-K/Akt and ERK1/2 MAPK pathways in immortalized,

hypothalamic neurons. Neuroendocrinology 89:267-275

396. Harrison RJ, McNeil GP, Dobner PR 1995 Synergistic activation of

neurotensin/neuromedin N gene expression by c-Jun and glucocorticoids: novel effects of

Fos family proteins. Mol Endocrinol 9:981-993

397. Doyle ME, Egan JM 2007 Mechanisms of action of glucagon-like peptide 1 in the

pancreas. Pharmacol Ther 113:546-593

398. Wang X, Murphy TJ 1998 Inhibition of cyclic AMP-dependent kinase by expression of

a protein kinase inhibitor/enhanced green fluorescent fusion protein attenuates

angiotensin II-induced type 1 AT1 receptor mRNA down-regulation in vascular smooth

muscle cells. Mol Pharmacol 54:514-524

399. Mayer CM, Belsham DD 2009 Insulin directly regulates NPY and AgRP gene

expression via the MAPK MEK/ERK signal transduction pathway in mHypoE-46

hypothalamic neurons. Mol Cell Endocrinol 307:99-108

400. Vo N, Klein ME, Varlamova O, Keller DM, Yamamoto T, Goodman RH, Impey S

2005 A cAMP-response element binding protein-induced microRNA regulates neuronal

morphogenesis. Proc Natl Acad Sci U S A 102:16426-16431

401. Han VK, Hynes MA, Jin C, Towle AC, Lauder JM, Lund PK 1986 Cellular

localization of proglucagon/glucagon-like peptide I messenger RNAs in rat brain. J

Neurosci Res 16:97-107

402. Drucker DJ, Erlich P, Asa SL, Brubaker PL 1996 Induction of intestinal epithelial

proliferation by glucagon-like peptide 2. Proc Natl Acad Sci U S A 93:7911-7916

403. Tsai CH, Hill M, Asa SL, Brubaker PL, Drucker DJ 1997 Intestinal growth-

promoting properties of glucagon-like peptide-2 in mice. Am J Physiol 273:E77-84

404. Tsai CH, Hill M, Drucker DJ 1997 Biological determinants of intestinotrophic

properties of GLP-2 in vivo. Am J Physiol 272:G662-668

405. Benjamin MA, McKay DM, Yang PC, Cameron H, Perdue MH 2000 Glucagon-like

peptide-2 enhances intestinal epithelial barrier function of both transcellular and

paracellular pathways in the mouse. Gut 47:112-119

406. Wojdemann M, Wettergren A, Hartmann B, Hilsted L, Holst JJ 1999 Inhibition of

sham feeding-stimulated human gastric acid secretion by glucagon-like peptide-2. J Clin

Endocrinol Metab 84:2513-2517

407. Dube PE, Forse CL, Bahrami J, Brubaker PL 2006 The essential role of insulin-like

growth factor-1 in the intestinal tropic effects of glucagon-like peptide-2 in mice.

Gastroenterology 131:589-605

179

408. Leen JL, Izzo A, Upadhyay C, Rowland KJ, Dube PE, Gu S, Heximer SP, Rhodes

CJ, Storm DR, Lund PK, Brubaker PL 2011 Mechanism of action of glucagon-like

peptide-2 to increase IGF-I mRNA in intestinal subepithelial fibroblasts. Endocrinology

152:436-446

409. Jennes L, Stumpf WE, Kalivas PW 1982 Neurotensin: topographical distribution in rat

brain by immunohistochemistry. J Comp Neurol 210:211-224

410. Williams G, Bing C, Cai XJ, Harrold JA, King PJ, Liu XH 2001 The hypothalamus

and the control of energy homeostasis: different circuits, different purposes. Physiol

Behav 74:683-701

411. Orskov C, Holst JJ, Knuhtsen S, Baldissera FG, Poulsen SS, Nielsen OV 1986

Glucagon-like peptides GLP-1 and GLP-2, predicted products of the glucagon gene, are

secreted separately from pig small intestine but not pancreas. Endocrinology 119:1467-

1475

412. Wynne K, Bloom SR 2006 The role of oxyntomodulin and peptide tyrosine-tyrosine

(PYY) in appetite control. Nat Clin Pract Endocrinol Metab 2:612-620

413. Dogrukol-Ak D, Tore F, Tuncel N 2004 Passage of VIP/PACAP/secretin family across

the blood-brain barrier: therapeutic effects. Curr Pharm Des 10:1325-1340

414. Schmidt PT, Naslund E, Gryback P, Jacobsson H, Hartmann B, Holst JJ, Hellstrom

PM 2003 Peripheral administration of GLP-2 to humans has no effect on gastric

emptying or satiety. Regul Pept 116:21-25

415. Jeppesen PB, Hartmann B, Thulesen J, Graff J, Lohmann J, Hansen BS, Tofteng F,

Poulsen SS, Madsen JL, Holst JJ, Mortensen PB 2001 Glucagon-like peptide 2

improves nutrient absorption and nutritional status in short-bowel patients with no colon.

Gastroenterology 120:806-815

416. Paul M, Poyan Mehr A, Kreutz R 2006 Physiology of local renin-angiotensin systems.

Physiol Rev 86:747-803

417. Mietlicki EG, Nowak EL, Daniels D 2009 The effect of ghrelin on water intake during

dipsogenic conditions. Physiol Behav 96:37-43

418. Lovshin JA, Huang Q, Seaberg R, Brubaker PL, Drucker DJ 2004

Extrahypothalamic expression of the glucagon-like peptide-2 receptor is coupled to

reduction of glutamate-induced cell death in cultured hippocampal cells. Endocrinology

145:3495-3506

419. Velazquez E, Ruiz-Albusac JM, Blazquez E 2003 Glucagon-like peptide-2 stimulates

the proliferation of cultured rat astrocytes. Eur J Biochem 270:3001-3009

180

420. Camby I, Salmon I, Bourdel E, Nagy N, Danguy A, Brotchi J, Pasteels JL, Martinez

J, Kiss R 1996 Neurotensin-mediated effects on astrocytic tumor cell proliferation.

Neuropeptides 30:133-139

421. Sato M, Nakahara K, Goto S, Kaiya H, Miyazato M, Date Y, Nakazato M, Kangawa

K, Murakami N 2006 Effects of ghrelin and des-acyl ghrelin on neurogenesis of the rat

fetal spinal cord. Biochem Biophys Res Commun 350:598-603

422. Zhang W, Lin TR, Hu Y, Fan Y, Zhao L, Stuenkel EL, Mulholland MW 2004

Ghrelin stimulates neurogenesis in the dorsal motor nucleus of the vagus. J Physiol

559:729-737

423. Wang Y, Guan X 2010 GLP-2 potentiates L-type Ca2+ channel activity associated with

stimulated glucose uptake in hippocampal neurons. Am J Physiol Endocrinol Metab

298:E156-166

424. Kageyama H, Takenoya F, Shiba K, Shioda S 2010 Neuronal circuits involving ghrelin

in the hypothalamus-mediated regulation of feeding. Neuropeptides 44:133-138

425. Gahete MD, Rubio A, Cordoba-Chacon J, Gracia-Navarro F, Kineman RD, Avila J,

Luque RM, Castano JP 2010 Expression of the ghrelin and neurotensin systems is

altered in the temporal lobe of Alzheimer's disease patients. J Alzheimers Dis 22:819-828

426. Asakawa A, Inui A, Fujimiya M, Sakamaki R, Shinfuku N, Ueta Y, Meguid MM,

Kasuga M 2005 Stomach regulates energy balance via acylated ghrelin and desacyl

ghrelin. Gut 54:18-24

427. Kishimoto M, Okimura Y, Nakata H, Kudo T, Iguchi G, Takahashi Y, Kaji H,

Chihara K 2003 Cloning and characterization of the 5(')-flanking region of the human

ghrelin gene. Biochem Biophys Res Commun 305:186-192

428. Dalvi PS, Belsham DD 2012 Glucagon-Like Peptide-2 Directly Regulates Hypothalamic

Neurons Expressing Neuropeptides Linked to Appetite Control in Vivo and in Vitro.

Endocrinology 153:2385-2397

429. Schwartz MW, Baskin DG, Bukowski TR, Kuijper JL, Foster D, Lasser G,

Prunkard DE, Porte D, Jr., Woods SC, Seeley RJ, Weigle DS 1996 Specificity of

leptin action on elevated blood glucose levels and hypothalamic neuropeptide Y gene

expression in ob/ob mice. Diabetes 45:531-535

430. Schwartz MW, Marks JL, Sipols AJ, Baskin DG, Woods SC, Kahn SE, Porte D, Jr.

1991 Central insulin administration reduces neuropeptide Y mRNA expression in the

arcuate nucleus of food-deprived lean (Fa/Fa) but not obese (fa/fa) Zucker rats.

Endocrinology 128:2645-2647

431. Inokuchi A, Oomura Y, Nishimura H 1984 Effect of intracerebroventricularly infused

glucagon on feeding behavior. Physiol Behav 33:397-400

181

432. Honda K, Kamisoyama H, Saito N, Kurose Y, Sugahara K, Hasegawa S 2007

Central administration of glucagon suppresses food intake in chicks. Neurosci Lett

416:198-201

433. Dakin CL, Small CJ, Batterham RL, Neary NM, Cohen MA, Patterson M, Ghatei

MA, Bloom SR 2004 Peripheral oxyntomodulin reduces food intake and body weight

gain in rats. Endocrinology 145:2687-2695

434. Ghatei MA, Uttenthal LO, Christofides ND, Bryant MG, Bloom SR 1983 Molecular

forms of human enteroglucagon in tissue and plasma: plasma responses to nutrient

stimuli in health and in disorders of the upper gastrointestinal tract. J Clin Endocrinol

Metab 57:488-495

435. Roberge JN, Brubaker PL 1991 Secretion of proglucagon-derived peptides in response

to intestinal luminal nutrients. Endocrinology 128:3169-3174

436. Kieffer TJ, McIntosh CH, Pederson RA 1995 Degradation of glucose-dependent

insulinotropic polypeptide and truncated glucagon-like peptide 1 in vitro and in vivo by

dipeptidyl peptidase IV. Endocrinology 136:3585-3596

437. Tavares W, Drucker DJ, Brubaker PL 2000 Enzymatic- and renal-dependent

catabolism of the intestinotropic hormone glucagon-like peptide-2 in rats. Am J Physiol

Endocrinol Metab 278:E134-139

438. Hartmann B, Harr MB, Jeppesen PB, Wojdemann M, Deacon CF, Mortensen PB,

Holst JJ 2000 In vivo and in vitro degradation of glucagon-like peptide-2 in humans. J

Clin Endocrinol Metab 85:2884-2888

439. Jasleen J, Ashley SW, Shimoda N, Zinner MJ, Whang EE 2002 Glucagon-like

peptide 2 stimulates intestinal epithelial proliferation in vitro. Dig Dis Sci 47:1135-1140

440. Sorensen LB, Flint A, Raben A, Hartmann B, Holst JJ, Astrup A 2003 No effect of

physiological concentrations of glucagon-like peptide-2 on appetite and energy intake in

normal weight subjects. Int J Obes Relat Metab Disord 27:450-456

441. Larsen PJ, Tang-Christensen M, Jessop DS 1997 Central administration of glucagon-

like peptide-1 activates hypothalamic neuroendocrine neurons in the rat. Endocrinology

138:4445-4455

442. Bavec A, Hallbrink M, Langel U, Zorko M 2003 Different role of intracellular loops of

glucagon-like peptide-1 receptor in G-protein coupling. Regul Pept 111:137-144

443. Koole C, Wootten D, Simms J, Valant C, Sridhar R, Woodman OL, Miller LJ,

Summers RJ, Christopoulos A, Sexton PM 2010 Allosteric ligands of the glucagon-

like peptide 1 receptor (GLP-1R) differentially modulate endogenous and exogenous

peptide responses in a pathway-selective manner: implications for drug screening. Mol

Pharmacol 78:456-465

182

444. Dickson SL, Shirazi RH, Hansson C, Bergquist F, Nissbrandt H, Skibicka KP 2012

The glucagon-like peptide 1 (GLP-1) analogue, exendin-4, decreases the rewarding value

of food: a new role for mesolimbic GLP-1 receptors. J Neurosci 32:4812-4820

445. Rinaman L 2010 Ascending projections from the caudal visceral nucleus of the solitary

tract to brain regions involved in food intake and energy expenditure. Brain Res 1350:18-

34

446. Alhadeff AL, Rupprecht LE, Hayes MR 2012 GLP-1 neurons in the nucleus of the

solitary tract project directly to the ventral tegmental area and nucleus accumbens to

control for food intake. Endocrinology 153:647-658

447. Clancy B, Darlington RB, Finlay BL 2001 Translating developmental time across

mammalian species. Neuroscience 105:7-17

448. Clancy B, Kersh B, Hyde J, Darlington RB, Anand KJ, Finlay BL 2007 Web-based

method for translating neurodevelopment from laboratory species to humans.

Neuroinformatics 5:79-94

449. Attardi B, Tsujii T, Friedman R, Zeng Z, Roberts JL, Dellovade T, Pfaff DW,

Chandran UR, Sullivan MW, DeFranco DB 1997 Glucocorticoid repression of

gonadotropin-releasing hormone gene expression and secretion in morphologically

distinct subpopulations of GT1-7 cells. Mol Cell Endocrinol 131:241-255

450. Northcutt RG, Muske LE 1994 Multiple embryonic origins of gonadotropin-releasing

hormone (GnRH) immunoreactive neurons. Brain Res Dev Brain Res 78:279-290

451. Chaillou E, Baumont R, Chilliard Y, Tillet Y 2002 Several subpopulations of

neuropeptide Y-containing neurons exist in the infundibular nucleus of sheep: an

immunohistochemical study of animals on different diets. J Comp Neurol 444:129-143

452. Bain J, Plater L, Elliott M, Shpiro N, Hastie CJ, McLauchlan H, Klevernic I,

Arthur JS, Alessi DR, Cohen P 2007 The selectivity of protein kinase inhibitors: a

further update. Biochem J 408:297-315

453. Gharbi SI, Zvelebil MJ, Shuttleworth SJ, Hancox T, Saghir N, Timms JF,

Waterfield MD 2007 Exploring the specificity of the PI3K family inhibitor LY294002.

Biochem J 404:15-21

454. Lachey JL, D'Alessio DA, Rinaman L, Elmquist JK, Drucker DJ, Seeley RJ 2005

The role of central glucagon-like peptide-1 in mediating the effects of visceral illness:

differential effects in rats and mice. Endocrinology 146:458-462

455. Le Bihan D, Urayama S, Aso T, Hanakawa T, Fukuyama H 2006 Direct and fast

detection of neuronal activation in the human brain with diffusion MRI. Proc Natl Acad

Sci U S A 103:8263-8268

456. 2008 Cell imaging: Light activated. Nature 456:826-827

183

457. Shibata C, Naito H, Jin XL, Ueno T, Funayama Y, Fukushima K, Hashimoto A,

Matsuno S, Sasaki I 2001 Effect of glucagon, glicentin, glucagon-like peptide-1 and -2

on interdigestive gastroduodenal motility in dogs with a vagally denervated gastric

pouch. Scand J Gastroenterol 36:1049-1055

458. Pellissier S, Sasaki K, Le-Nguyen D, Bataille D, Jarrousse C 2004 Oxyntomodulin

and glicentin are potent inhibitors of the fed motility pattern in small intestine.

Neurogastroenterol Motil 16:455-463

459. Day JW, Ottaway N, Patterson JT, Gelfanov V, Smiley D, Gidda J, Findeisen H,

Bruemmer D, Drucker DJ, Chaudhary N, Holland J, Hembree J, Abplanalp W,

Grant E, Ruehl J, Wilson H, Kirchner H, Lockie SH, Hofmann S, Woods SC,

Nogueiras R, Pfluger PT, Perez-Tilve D, DiMarchi R, Tschop MH 2009 A new

glucagon and GLP-1 co-agonist eliminates obesity in rodents. Nat Chem Biol 5:749-757

460. Hoffman GE, Smith MS, Fitzsimmons MD 1992 Detecting steroidal effects on

immediate early gene expression in the hypothalamus. Neuroprotocols 1:52-66

461. Wetsel WC, Eraly SA, Whyte DB, Mellon PL 1993 Regulation of gonadotropin-

releasing hormone by protein kinase-A and -C in immortalized hypothalamic neurons.

Endocrinology 132:2360-2370

462. Vijayan E, Carraway R, Leeman SE, McCann SM 1988 Use of antiserum to

neurotensin reveals a physiological role for the peptide in rat prolactin release. Proc Natl

Acad Sci U S A 85:9866-9869

463. Gully D, Canton M, Boigegrain R, Jeanjean F, Molimard JC, Poncelet M, Gueudet

C, Heaulme M, Leyris R, Brouard A, et al. 1993 Biochemical and pharmacological

profile of a potent and selective nonpeptide antagonist of the neurotensin receptor. Proc

Natl Acad Sci U S A 90:65-69

464. Szentirmai É, Kapás L, Sun Y, Smith RG, Krueger JM 2009 The preproghrelin gene

is required for the normal integration of thermoregulation and sleep in mice. Proceedings

of the National Academy of Sciences 106:14069-14074