374
Characterization of mechanisms of myocardial remodeling in genetic models of cardiac hypertrophy Andrea A. Domenighetti Submitted in total fulfillment of the requirements of the degree of Doctor of Philosophy December 2005 Department of Physiology The University of Melbourne, Australia Produced on archival quality paper

Characterization of mechanisms of myocardial remodeling in

Embed Size (px)

Citation preview

Characterization of mechanisms of myocardial

remodeling in genetic models of cardiac hypertrophy

Andrea A. Domenighetti

Submitted in total fulfillment of the requirements of

the degree of Doctor of Philosophy

December 2005

Department of Physiology

The University of Melbourne, Australia

Produced on archival quality paper

ii

To Esther and Guido Domenighetti

iii

ABSTRACT

Introduction and aims:

Cardiac hypertrophy is clinically defined as a relative increase in heart size associated

with a thickening of the ventricular wall. It is a common feature of individuals suffering

from different cardio-vascular or metabolic conditions and leads to heart failure. The

structural, functional and molecular mechanisms which induce hypertrophy

independent of hemodynamic alterations are poorly characterized. In this study,

questions about whether cardiac-specific neuro-endocrine activation or metabolic

imbalance are sufficient to induce hypertrophic structural and functional remodeling are

addressed using genetically manipulated mouse models of primary cardiac

hypertrophy.

Two different transgenic mouse models of blood pressure-independent cardiac

hypertrophy (i.e. of primary cardiac hypertrophy) were investigated: 1) a cardiac-

specific angiotensinogen-overexpressing transgenic mouse, the TG1306/1R; and 2) a

muscle-specific glucose transporter 4 (GLUT4) knock-out mouse model, the GLUT4-

KO. It was hypothesized that cardiac-specific activation of the renin-angiotensin

system or a decrease in cardiac glucose uptake are sufficient to induce the heart to

undergo pathological hypertrophy and failure. Experimental investigations included

cellular and tissue morphometric analysis combined with assessment of isolated adult

cardiomyocyte contractility and gene expression profiling using RT-PCR and cDNA

microarray assays.

Cardiac and cardiomyocyte hypertrophy in the TG1306/1R mouse:

The present study demonstrates that cardiac remodeling in TG1306/1R transgenic (TG)

mice is associated with cardiomyocyte hypertrophy but not fibrosis, when compared to

age-matched wild-type (WT) littermates. Molecular analysis showed an approximate

iv

10-fold upregulation of the angiotensinogen (Agt) mRNA, associated with an

approximate 25% reduction in GLUT4 protein level in the hearts of TG mice.

Analysis of isolated cardiomyocyte isotonic shortening showed age-dependent

impairment of contractility in TG cardiomyocytes. Comparisons between age-matched

TG and WT mycoytes showed that both rates of shortening and lengthening (maximal

rate of cell shortening [MRS] and maximal rate of cell lengthening [MRL]) were

reduced by 15-35% in the TG myocytes relative to WT. This indicates a reduction in

inotropic and lusitropic performance in the TG cardiomyocytes, independent of age.

The slowest contraction kinetics were observed in myocytes from older TG hearts.

Genotype-dependent prolongation of the contraction cycle was also evident at both

ages in TG cardiomyocytes.

In TG hearts, cardiomyocyte dysfunction was associated with a 5-fold downregulation

of the sarcoplasmic reticulum (SR) calcium ATPase pump (SERCA2) and upregulation

of the sodium-calcium exchanger NCX1.1 and the sodium-hydrogen exchanger NHE-1

mRNA levels. No significant differences were observed in the cardiac expression of the

ryanodine receptor RyR2 mRNA between WT and TG mice.

Microarray expression profiling showed that angiotensin II- (Ang II) stimulated cardiac

hypertrophy is predominantly associated with upregulation of pro-hypertrophic genes

involved in protein biosynthesis and cell differentiation. Differential gene expression

analysis of TG hearts also detected a decrease in expression of mRNA involved in

glucose and mitochondrial fatty acid metabolism relative to WT, and a possible

compensatory activation of peroxisomal free fatty acid beta oxidation.

Cardiac and cardiomyocyte remodeling in the GLUT4-KO mouse:

GLUT4 ‘Knock-out’ mice expressing the Lox+/+ and Cre+/- genetic constructs (LLC),

develope cardiac and cardiomyocyte hypertrophy accompanied by a 6-fold increase in

myocardial collagen content, when compared to genetic controls expressing the Lox+/+

(LL) construct only. In LLC mice cardiac and cardiomyocyte remodeling was

associated with a 99% reduction in GLUT4 protein and an approximate 2-fold

upregulation of cardiac Agt mRNA expression, when compared to age-matched LL

v

mice. Interestingly, an 85% reduction in GLUT4 protein was also observed in hearts of

LL mice, when compared to age-matched control C57BL6 mice (i.e. WT). This

GLUT4 reduction in LL mice relative to WT was also associated with a modest but

significant increase in cardiac and cardiomyocyte dimensions. Regression analysis

showed that suppression of GLUT4 protein levels in hearts to a level below 5% of WT

level was linked to a gross hypertrophic and pro-fibrotic cardiac phenotype in LLC

mice.

Analysis of isolated cardiomyocyte isotonic shortening showed age-dependent

impairment of contractility in LLC cardiomyocytes compare to LL. Analysis of the

specific changes in contractile cycle parameters showed that at both ages LLC

cardiomyocytes exhibited significant impairment of contractile performance and cycle

timing. MRS and MRL were reduced by 20-25% in the myocytes of both 15-20 and 35-

40 week-old LLC mice relative to myocytes from age-matched LL mice. The

maximum shortening (%S) attained by LLC myocytes was 25-35% less than that

achieved by age-matched LL myocytes. Genotype-dependent prolongation of the

contraction cycle was also evident at both ages in LLC cardiomyocytes.

In LLC hearts, cardiomyocyte dysfunction was coupled to a ~2-fold downregulation of

SERCA2 protein levels, in association with a downregulation of the NCX1.1 and an

upregulation of NHE-1 mRNA expression. A significant downregulation was observed

in cardiac mRNA levels of the ryanodine receptor RyR2 mRNA in LLC mice

compared with LL mice.

Microarray expression profiles revealed that cardiac remodeling in LLC mice was

associated with strong downregulation of genes involved in mitochondrial energy

production and upregulation of genes involved in cell proliferation and tissue

inflammation. The same analysis demonstrated a downregulation of enzymes involved

in beta-oxidation and mitochondrial electron transport chain and upregulation of

enzymes involved in glycolysis and gluconeogenesis. These data suggest a shift away

from the production of aerobically derived ATP involving mitochondrial oxidation of

NADH equivalents, to an anaerobic pathway where glucose and glycogen are

metabolized to pyruvate and then reduced to lactate by lactate dehydrogenase.

vi

In conclusion:

Activation of the intra-cardiac renin-angiotensin system or decreased cardiac glucose

uptake lead to cardiac and cardiomyocyte remodeling, resulting in cardiac hypertrophy

and failure in mice. The severity of cardiomyocyte functional remodeling is determined

by the extent of cellular and tissue hypertrophy and myocardial collagen deposition.

The present study supports the concept that cardiac-specific activation of the renin-

angiotensin system or a decrease in cardiac glucose uptake are sufficient ‘triggers’ to

activate pathologic cardiac remodeling, the severity of which is dependent on the

degree of activation of the intra-cardiac renin-angiotensin system and the extent of

glucose metabolic perturbation.

vii

STATEMENT OF AUTHORSHIP

This is to certify that:

This thesis comprises only my original work towards the PhD.

Due acknowledgment has been made in the text to all other material used.

This thesis is less than 100,000 words in length, exclusive of tables, figures and

references.

Andrea A. Domenighetti

December 2005

viii

ACKNOWLEDGMENTS

Looking back on my sojourn in Australia I have numerous people to thank, affirming

that research is meaningless without collaborative effort. In particular, I would like to

acknowledge the following people - without them I would not be submitting this

Thesis. Whether my gratitude is for intellectual supervision, or for making my time as

an overseas PhD student immensely enjoyable, both are equally acknowledged.

A/Prof. Lea M. Delbridge, for giving me the opportunity to undertake a PhD, guiding

me and encouraging me to keep persevering and to pursue science at high levels. Thank

you for giving me the freedom to explore the subject in my own way, while at the same

time offering frequent physiological insight into exciting results. Thank you also for the

good times together with Chris, Alex and Kim.

Prof. Thierry Pedrazzini (Department of Medicine, University of Lausanne Medical

School, Switzerland), for shipping the TG1306/1R transgenic mouse strain to Australia

- it was not an easy task! Thank you for your unconditional support and for acting as a

‘stand-in’ supervisor during the time spent in Lausanne.

Prof. Joseph Proietto (Department of Medicine, Repatriation Hospital, Melbourne),

for providing the GLUT4-KO mice. Thank you for the pleasant moments spent in your

laboratory and with your family!

Prof. Stephen Harrap, for allowing me to undertake my research in the Department of

Physiology at the University of Melbourne, and for keeping my cardiovascular system

fit with exciting tennis games. I’m already missing the Federer style backhands and

smashes!

Prof. Hans R. Brunner and Prof. Trefor Morgan, for beginning the collaboration

between the Department of Physiology in Melbourne and the Division of Hypertension

ix

in Lausanne, Switzerland. Your collaborative efforts gave me the fantastic opportunity

to spend four years of my life in Australia. A personal acknowledgment to Prof.

Trefor Morgan for sharing his ‘savoir faire’ and his passion for winemaking and wine

tasting. I immensely enjoyed picking grapes at Mount Charlie!

Dr. Neil Williams, for giving me my first experience of Australian friendship and

hospitality. Thank you for hosting me in your house for more than one month and for

introducing me to the tasteful and joyful world of Australian gastronomy and wine!

Dr. Stephen M. Richards (The School of Medicine, University of Tasmania), Dr.

Gregory Jones (Department of Medicine and Surgery, Otago University, New

Zealand), Mrs. Faye Doherty (Department of Anatomy and Cell Biology, Melbourne

University), Dr. Cory Griffiths (Forensic Science Service, Tasmania), Dr. Sofianos

Andrikopoulos (Department of Medicine, The Royal Melbourne Hospital), Dr. Garry

Myers (The Institute for Genomic Research -TIGR- Maryland, USA), Dr. Matthew

Ritchie (WEHI Bioinformatics Group, Melbourne), Dr. Gordon Smyth (WEHI

Bioinformatics Group, Melbourne) and Dr. Robert Di Nicolantonio (variously current

or past laboratory colleagues and/or collaborators), for collectively teaching me more

than just the basics of nucleotide amplification, histology, cardiomyocyte contractility,

Western blotting, cDNA microarray assays and statistical analysis. Thank you for

sharing your valued time, your computer desktops and your bench spaces with me.

Your warm friendship and your ‘bucatini ca pummarola’ will always be with me

(Robert)!

Other ‘Delbridge lab’ members, past and present, including Danielle Hart, Petcharat

Trongtorsak, Venne Danes, Claire Curl, Clare Lax, Catherine Huggins, Nadine

Khalil, Emad Abro, Enzo Porrello, and Anna Caldwell. Thank you for your

scientific and technical participation. I will not forget the laughter, the moments shared

during our never-ending lab meetings and the Whitlams (Clare)!

The personnel of the Department of Physiology, in particular Chris Adamidis, Joyce

Kelly, Lesley Robinson, Christine Hofsteter, Karin Diamond, Charles

Chlebowczyk, Philip Dubbin and Jim Pringle. Thank you for your prompt

x

administrative and technical support, the numerous roaming profiles created and

deleted, the tennis games, the Christmas parties and the friendship.

All my Australian mates, especially my colleague and housemate David Plant. It was

just simply too good for words! Thank you Gordon, ‘Jack’, ‘Mattoon’, ‘Bryan’,

James, Sarah, Angela, Tania, the office mates, the other PhD students of the

Department, together with Debby, Angus, Christina, Abby and Michi. Thank you for

the excursions to Mt Bulla, the Great Ocean Road, the parties and the nightlife in

Melbourne - eating, drinking, dancing, singing and keeping me (in)sane!

My beloved Sylvie Lurot. Thank you for your patience and your efficient instruction

on ‘Microsoft Word and Adobe Acrobat for dummies’. Without your support, this

thesis would not have had the same ‘look’.

Last but not least, an extra thanks to Catherine Huggins for the special and wonderful

job done in assembling the final version of this document in Melbourne for remote

submission.

The Roche Research Foundation (Fli7stm 98-120), the Swiss National Science

Foundation and the University of Melbourne are warmly acknowledged for

scholarship and financial support.

xi

INDEX CHAPTER I: General introduction and literature review 1 1. INTRODUCTION

2

2. DEFINITION OF CARDIAC HYPERTROPHY 4 2.1. Concentric cardiac hypertrophy 4 2.2. Eccentric cardiac hypertrophy 5 2.3. Ventricular dilation

6

3. FAMILIAL FORMS OF HEART DISEASES 7 3.1. Familial hypertrophic cardiomyopathy 7 3.2. Dilated cardiomyopathy 8 3.3. Congenital heart diseases 9 3.4. Cardiac arrhythmias 9 3.5. Coronary diseases

10

4. HYPERTROPHY AS A PRELUDE TO HEART FAILURE 11 4.1. Definition of heart failure 11 4.2. Transition from cardiac hypertrophy to heart failure 12 4.3. Incidence and clinical causes of heart failure

14

5. ALTERED EXCITATION-CONTRACTION COUPLING IN CARDIAC HYPERTROPHY AND FAILURE

15

5.1. Electrical changes in cardiac hypertrophy and heart failure 16 5.2. Calcium cycling in cardiac hypertrophy and heart failure 17

5.2.1. Calcium cycling in compensated cardiac hypertrophy 17 5.2.2. Calcium cycling in decompensated cardiac remodeling 18 5.2.3. The sodium/calcium exchanger (NCX) 19

5.3. Potassium currents in cardiac hypertrophy and heart failure 20 5.4. Sympathetic stimulation and electrical impairment 21 5.5. Cell-to-cell coupling and electrical communication 21 5.6. The sodium/hydrogen exchanger (NHE) and pH regulation in cardiac hypertrophy

23

6. MOLECULAR PATHWAYS FOR CARDIAC DEVELOPMENT, HYPERTROPHY AND FAILURE

26

6.1. Embryonic heart development and cell fate determination 26 6.1.1 BMP and Wnt signaling during embryonic development 27 6.1.2 Nkx and GATA signaling 28 6.1.3 MEF2 and cardiomyocyte differentiation 29

6.2. Cardiomyocyte growth and hypertrophy in the adult heart 30 6.2.1. Different signaling pathways for different hypertrophic stimuli 31 6.2.2. Signaling pathways through seven-transmembrane receptors 31 6.2.3. Signaling through mitogen-activated protein kinases (MAPKs) 33

xii

6.2.4. Calcium-dependent pathways of cardiac hypertrophy induction 33 6.3. Cardiomyocyte death 34 6.4. Cardiomyocyte growth and death: a fine balancing act 36 6.5. The fibroblast fraction: collagen production and cardiomyocyte hypertrophy 38 6.6. Extracellular matrix turnover and fibrosis 39

6.6.1 Matrix degradation 39 6.6.2 Matrix biosynthesis 39 6.6.3 Matrix cell adhesion and transduction of signaling pathways 40

7. THE RENIN-ANGIOTENSIN SYSTEM AND CARDIAC REMODELING 42

7.1. An overview 42 7.2. The different components of the renin-angiotensin system 43

7.2.1. Renin 43 7.2.2. The angiotensin-converting enzyme 44 7.2.3. The angiotensinogen precursor 45 7.2.4. Angiotensin peptides 46 7.2.5. Angiotensin II 47 7.2.6. The AT1 receptor 47 7.2.7. The AT2 receptor 49

7.3. Cardiac components of the renin-angiotensin system 51

8. METABOLIC REGULATION OF CARDIAC HYPERTROPHY 52 8.1. Metabolism in the normal heart 52

8.1.1. Fatty acid metabolism 52 8.1.2. Glucose metabolism 53

8.2. Metabolic adaptation in cardiac hypertrophy and failure 54 8.3. Cardiac metabolism in type 2 diabetes and insulin resistant states 55

8.3.1. Diabetes mellitus 55 8.3.2. Insulin resistance 56 8.3.3. Cardiac glucose transport in states of insulin resistance 57 8.3.4. Cardiac fatty acid regulation in states of insulin resistance 58

8.4. Diabetic cardiomyopathy 59 8.4.1. Metabolic disturbances in diabetic cardiomyopathy 59 8.4.2. Mechanical dysfunction in diabetic cardiomyopathy 60

9. AIMS 61 CHAPTER II: General methods 72 1. EXPERIMENTAL MODELS 73

1.1. Ethics approval 73 1.2. The transgenic angiotensinogen overexpressing mouse (TG1306/1R) 73 1.3. The GLUT4 transporter knock-out mouse (GLUT4-KO) 74

2. CARDIAC WEIGHT INDEX 75

3. WHOLE HEART HISTOLOGY AND MORPHOMETRY 76

3.1. Heart preparation and staining 76 3.2. Histological imaging 76

xiii

4. ADULT CARDIOMYOCYTE CELL ISOLATION 78

5. ADULT CARDIOMYOCYTE CONTRACTILITY 79 5.1. Experimental conditions 79 5.2. Recording and analysis 80

6. PROTEIN EXTRACTION AND WESTERN BLOTTING 81

7. TOTAL RNA EXTRACTION AND RT-PCR 82

7.1. RNA extraction 82 7.2. Optimization of RT-PCR conditions 82

8. cDNA MICROARRAY ASSAYS 85

8.1. Principles of DNA microarray analysis 85 8.2. RNA and cDNA preparation for hybridization 85 8.3. Pre-hybridization and hybridization of cDNA microarray slides 86 8.4. Slide washing and scanning 87 8.5. Statistical normalization of cDNA microarray data and clone ranking 88 8.6. Gene Ontology classification of selected transcript candidates 90

9. STATISTICS 92 CHAPTER III: Diverse evolving cardiac and cardiomyocyte phenotypes in Ang II-induced and insulin resistant cardiac hypertrophy

103

1. INTRODUCTION 104

1.1. Cardiac adaptation in response to environmental changes 104 1.2. In vivo models of Ang II-induced cardiac remodeling 104

1.2.1. Cardiac remodeling regression by Ang II suppression 104 1.2.2. Overexpressing or knocking out the intra-cardiac RAS 105 1.2.3. Overexpressing the AT receptors 105 1.2.4. Knocking out the AT receptors 107 1.2.5. Knocking out the angiotensinogen gene 109 1.2.6. Overexpressing Ang II fusion protein directly in the heart 110 1.2.7. Overexpressing the angiotensinogen gene in the heart 110

1.3. Transgenic and knock out models for the GLUT4 112 1.3.1. Overexpressing or knocking out the GLUT4 112 1.3.2. The GLUT4-KO mouse model 113

1.4. In summary 117 1.5. Aims 117

2. METHODS 119

2.1. Whole heart histology and morphometry 119 2.1.1. Procedures 119 2.1.2. Experimental groups 119

2.2. Longevity data and homozygote TG1306/1R mice 120 2.3. RNA and protein extraction and quantification 120

xiv

2.3.1. RNA extraction, RT-PCR and Western blotting 120 2.3.2. Experimental groups 121

2.4. Statistical considerations and presentation of the results 122

3. RESULTS 123 3.1. Cardiac and cardiomyocyte remodeling in TG1306/1R mice 123

3.1.1. Survival and the hypertrophic phenotype in TG mice 123 3.1.2. Myocardial and chamber remodeling in TG hearts 123 3.1.3. Cardiomyocyte hypertrophy but not fibrosis in TG hearts 124 3.1.4. Differential gene expression in TG hearts 125 3.1.5. Decreased GLUT4 protein content in TG hearts 125

3.2. Cardiac and cardiomyocyte remodeling in GLUT4-KO mice 126 3.2.1. Myocardial and chamber remodeling in GLUT4-KO hearts 126 3.2.2. Cardiomyocyte hypertrophy and fibrosis in LLC hearts 126 3.2.3. Differential gene expression in hearts from LLC mice 127 3.2.4. Decreased GLUT4 protein content in LL and LLC hearts 128

4. DISCUSSION 129

4.1. Cardiac and cardiomyocyte remodeling in TG1306/1R mice 129 4.1.1. Concentric hypertrophy and ventricular dilation in TG mice 129 4.1.2. Cardiac and cardiomyocyte remodeling but not fibrosis in TG 130 4.1.3. Differential gene expression profiles in TG hearts 131 4.1.4. Decreased GLUT4 protein levels in hearts of TG mice 133

4.2. Cardiac and cardiomyocyte remodeling in LLC mice 133 4.2.1. Cardiomyocyte hypertrophy and fibrotic remodeling in LLC 133 4.2.2. Decreased cell-to-cell connectivity in LLC hearts 135 4.2.3. Decreased GLUT4 protein levels in hearts of LLC and LL mice 136

5. IN SUMMARY 137 CHAPTER IV: Impaired cardiomyocyte contractility and differential expression of calcium and proton transporters in ageing models of Ang II-induced and insulin resistant cardiac hypertrophy

150

1. INTRODUCTION 151

1.1. E-C coupling with increased pacing frequency in the mouse 151 1.2. Modulation of cardiomyocyte contractility by Ang II 152

1.2.1. Acute modulation of myocyte contractility by Ang II 152 1.2.2. Mechanisms for differential inotropic effects of Ang II on cardiac tissue

154

1.2.3. Regulation of the sodium/hydrogen exchanger (NHE) by Ang II 155 1.2.4. Regulation of the sodium/calcium exchanger (NCX) by Ang II 156 1.2.5. Regulation of SR calcium ATPase (SERCA2) by Ang II 156 1.2.6. Regulation of the ryanodine receptors RyR by Ang II 157 1.2.7. Regulation of the voltage-operated calcium channels by Ang II 158 1.2.8. Integrating the mechanisms of Ang II modulation of cardiomyocyte inotropy

159

xv

1.3. Diabetic cardiomyopathy and cardiomyocyte contractility 161 1.3.1. Cardiac dysfunction related to diabetic cardiomyopathy 161 1.3.2. Altered pH homeostasis in diabetic cardiomyopathy 164 1.3.3. Altered calcium homeostasis in diabetic cardiomyopathy 165

1.4. Insulin resistant cardiomyopathy: evidence of activation of the renin-angiotensin system

166

1.5. Aims 167

2. METHODS 168 2.1. Cardiomyocyte contractility 168

2.1.1. Myocyte preparation and recording 168 2.1.2. Experimental groups 168 2.1.3. Recording protocols and data analysis for cardiomyocyte contractility 169

2.2. RNA and protein extraction and quantification 170 2.2.1. Procedures 170 2.2.2. Experimental groups 170

2.3. Statistical considerations and presentation of the results 171

3. RESULTS 173 3.1. Stability of contractile performance 173 3.2. Impaired cardiomyocyte contractility in Agt overexpressing mice 173

3.2.1. Decreased inotropy, lusitropy and prolonged cycle time at 5 Hz in TG 173 3.2.2. Age- and frequency-dependent alterations in cardiomyocyte function 174 3.2.3. Differential gene and protein expression profiles 175

3.3. Impaired cardiomyocyte contractility in GLUT4-deficient mice 176 3.3.1. Decreased inotropy, lusitropy and prolonged cycle timing at 5 Hz 176 3.3.2. Age- and frequency-dependent alterations in cardiomyocyte function 177 3.3.3. Differential gene and protein expression profiles 178 3.3.4. After-contraction in LLC cardiomyocytes at 1.5 Hz. 178

4. DISCUSSION 179

4.1. Decreased contractile performance in TG cardiomyocytes 179 4.2 Increased twitch duration in TG cardiomyocytes 182 4.3. Decreased inotropy and lusitropy in LLC cardiomyocytes 184 4.4. Increased twitch duration in LLC cardiomyocytes 186 4.5. Negative contraction-frequency relationship in mouse cardiomyocytes 187 4.6. After-contraction, matters of calcium overload? 189

5. IN SUMMARY 190

xvi

CHAPTER V: Microarray analysis of global changes in gene expression in Ang II-induced and insulin resistant cardiac hypertrophy

207

1. INTRODUCTION 208

1.1. Application of microarray analysis 208 1.2. cDNA versus oligonucleotide microarray assays 209 1.3. Application of microarray analysis to the study of heart development and cardiovascular diseases

210

1.3.1. Heart and cardiomyocyte development 210 1.3.2. Myocardial infarction and ischemia 212 1.3.3. Cardiac hypertrophy and heart failure 213 1.3.4. Hereditary hypertrophic cardiomyopathy 216 1.3.5. Uncomplicated and hypertension-associated human obesity 216

1.4. In summary 217 1.5. Aims 218

2. METHODS 219

2.1. RNA and cDNA preparation for hybridization 219 2.2. cDNA microarray assays and experimental groups 219

2.2.1. Choice of array design 219 2.2.2. TG versus WT hearts 220 2.2.3. LLC versus LL hearts 220

3. RESULTS 221

3.1. Differential gene expression and gene clustering in TG mice 221 3.1.1. Differential gene expression in TG mice 221 3.1.2. Gene Ontology classification of candidates 221

3.2. Differential gene expression and gene clustering in LLC mice 222 3.2.1. Differential gene expression in LLC hearts 222 3.2.2. Gene Ontology classification in LLC hearts 223

3.3. Comparative analysis of TG1306/1R and GLUT4-KO 224 3.3.1. Common and distinct gene sets altered in TG and LLC 224 3.3.2. Comparative GO analysis between TG and LLC 224

4. DISCUSSION 226

4.1. Differential gene expression in TG1306/1R mice 226 4.1.1. Evidence for general transcriptional upregulation in TG hearts 226 4.1.2. Evidence for enhanced protein synthesis in TG hearts 227 4.1.3. Overexpression of ribosomal subunits in TG hearts 228 4.1.4. Overexpression of chaperones in TG hearts 229 4.1.5. Overexpression of calreticulin in TG hearts 231 4.1.6. SR and ER stress: are they related phenomena? 232 4.1.7. Differential regulation of Wnt signaling in TG hearts 233 4.1.8. Cytoskeletal remodeling in TG hearts 235 4.1.9. Differential protein turnover and catabolism in TG hearts 236 4.1.10. Are sex hormones involved in cardiac remodeling? 237

xvii

4.1.11. Metabolic disturbances in TG hearts 239 4.1.12. Carbonic anhydrase and NHE1 upregulation 241

4.2. Differential gene expression in LLC hearts 242 4.2.1. Metabolic and mitochondrial impairment in LLC hearts 242 4.2.2. Metal and iron homeostasis in LLC hearts 244 4.2.3. Genes associated with hereditary and non-hereditary cardiomyopathy in LLC hearts

245

4.2.4. Adrenomedullin receptor overexpression in LLC hearts 246 4.2.5. Cell cycle regulation and fibroblast proliferation in LLC hearts 246

4.3. Common set of genes altered in TG and LLC hearts 248 4.3.1. Common downregulated genes 248 4.3.2. Reciprocally regulated genes 249

5. IN SUMMARY 251 CHAPTER VI: Mechanisms of cardiac remodeling in Ang II-induced and insulin resistant cardiac hypertrophy: a comparative overview

263

1. MORPHOLOGIC CHANGES IN CARDIAC HYPERTROPHY 264

2. MECHANICAL CHANGES IN CARDIAC HYPERTROPHY 266

3. DIFFERENTIAL GENE EXPRESSION OF METABOLIC

SUBSTRATES IN CARDIAC HYPERTROPHY 268

4. IN CONCLUSION 271 Bibliography 279

xviii

INDEX OF TABLES AND FIGURES

CHAPTER I: Figure I-1: Intracellular ionic homeostasis during cardiomyocyte contraction and relaxation

62

Figure I-2: Embryonic heart formation 63 Figure I-3: Signaling through 7TM receptors and G-protein 64 Figure I-4: Signaling through EGF receptor transactivation 65 Figure I-5: The MAPK signaling cascade 66 Figure I-6: Hypertrophic growth through calcineurin and NF-ATc 67 Figure I-7: Mechanisms of caspase-induced apoptosis 68 Figure I-8: The renin-angiotensin system (RAS) 69 Figure I-9: Generation and degradation of angiotensin peptides 70 Figure I-10: PPARα-dependent metabolic genes transcription 71 CHAPTER II: Figure II-1: Histological analysis of myocardial collagen density 93 Figure II-2: Histological analysis of myocardial nuclei count 94 Figure II-3: Depiction of cardiac perfusion and myocyte isolation apparatus 95 Figure II-4: Depiction of cardiomyocyte contractile function evaluation 96 Figure II-5: Illustration of RT-PCR amplification and gel analysis 97 Figure II-6: Illustration of microarray ‘target’ and ‘probe’ species 98 Figure II-7: Summarized microarray hybridization protocol 99 Figure II-8: Microarray hybridization and visualization depiction 100 Figure II-9: Depiction of a MA plot after normalization 101 Figure II-10: Side-by-side box plots of the M-values from a microarray experiment

102

CHAPTER III: Table III-1: Age-dependent cardiac structural remodeling in WT and TG mice

138

Table III-2: Age-dependent cardiac structural remodeling in LL and LLC mice

139

Figure III-1: Decreased survival and dilated hypertrophy in TG hearts 140 Figure III-2: Ang II expression levels and severity of remodeling 141 Figure III-3: Agt overexpression and downregulation of Cx43 in TG and WT hearts

142

Figure III-4: GLUT4 protein expression and negative correlation with CWI in TG and WT hearts

143

Figure III-5: Cardiac remodeling in LL and LLC hearts 144 Figure III-6: Fibrosis and non-myocyte proliferation in LLC hearts 145 Figure III-7: Diffuse and focal/reparative fibrosis in LLC hearts 146 Figure III-8: Agt and Cx43 gene expression profiles in LLC and LL hearts 147 Figure III-9: GLUT4 protein expression and correlation with CWI in LLC and LL hearts

148

Figure III-10: GLUT4 protein expression and correlation with CWI in LLC and LL versus WT hearts

149

xix

CHAPTER IV: Table IV-1: Isotonic shortening of adult cardiomyocytes from WT and TG mice

191

Table IV-2: Isotonic shortening of adult cardiomyocytes from LL and LLC mice

192

Table IV-3: Summary of the multiple-way ANOVA analysis for the TG and WT myocytes

193

Table IV-4: Summary of the multiple-way ANOVA analysis for the LLC and LL myocytes

194

Figure IV-1: Effects of acute Ang II modulation of cardiomyocyte contractility

195

Figure IV-2: Effects of chronic Ang II modulation of cardiomyocyte contractility

196

Figure IV-3: Protocol summary for the contractility experiments 197 Figure IV-4: Evaluation of %S and Tf prior to and after the frequency ramp (TG)

198

Figure IV-5: Evaluation of %S and Tf prior to and after the frequency ramp (LLC)

199

Figure IV-6: Pacing responses of isotonically shortening adult myocytes from WT and TG

200

Figure IV-7: Pacing responses of isotonically shortening adult myocytes from WT and TG

201

Figure IV-8: mRNA and protein expression profiles in WT and TG hearts 202 Figure IV-9: Pacing responses of isotonically shortening adult myocytes from LL and LLC mice

203

Figure IV-10: Pacing responses of isotonically shortening adult myocytes from LL and LLC

204

Figure IV-11: mRNA and protein expression profiles in LL and LLC hearts 205 Figure IV-12: After-contraction in LLC myocytes 206

xx

CHAPTER V: Table V-1: Upregulated clones in TG ventricle 252 Table V-2: Downregulated clones in TG ventricle 253 Table V-3: Upregulated clones in LLC ventricle 254 Table V-4: Downregulated clones in LLC ventricle 255

Figure V-1: Gene Ontology classification of genes in TG hearts according to the cellular location

256

Figure V-2: Gene Ontology classification in TG hearts according to the molecular function

257

Figure V-3: Gene Ontology classification in TG hearts according to the biological process

258

Figure V-4: Gene Ontology classification of genes in LLC mice according to the cellular location

259

Figure V-5: Gene Ontology classification in LLC hearts according to the molecular function

260

Figure V-6: Gene Ontology classification in LLC hearts according to the biological process

261

Figure V-7: Depiction of the physical and biochemical interactions between CAR - NHE-1 - AE3

262

CHAPTER VI: Figure VI-1: Comparative analysis of cardiac remodeling in WT, TG, LL and LLC mice

273

Figure VI-2: Pacing responses of isotonically shortening myocytes from WT and LL (1.5-5.0 Hz)

274

Figure VI-3: Pacing responses of isotonically shortening myocytes from WT and LL (1.5-5.0 Hz)

275

Figure VI-4: mRNA and protein expression profiles in WT and LL hearts 276 Figure VI-5: Pacing responses of isotonically shortening myocytes from TG and LLC (1.5-5.0 Hz)

277

Figure VI-6: Pacing responses of isotonically shortening myocytes from TG and LLC (1.5-5.0 Hz)

278

GRAPHICS COPYRIGHT ACKNOWLEDGEMENTS

Chapter I: http://www.myogen.com/discovery/cardiac.php

1

Chapter V: http://www.tqnyc.org/NYC040844/Mitosis.htm 207

Bibliography: http://www.usask.ca/antiquities/Collection/Rosetta_Stone.html

279

xxi

ABBREVIATIONS (excluding gene nomenclature)

7TM Seven-transmembrane (receptor)

ACE Angiotensin converting enzyme

Agt Angiotensinogen

Ang Angiotensin (e.g. Ang II= Angiotensin II)

ANP Atrial natriuretic peptide

AT Angiotensin receptor (e.g. AT1 and AT2= Angiotensin receptor 1 and 2)

β1 and 2-AR Beta 1 & 2 adrenoceptor subtypes

BMP Bone morphogenetic proteins

BNP Brain natriuretic peptide

CICR Calcium-induced calcium release

CWI Cardiac weight index (heart weight/body weight ratio)

Cx Connexin (e.g. Cx43= Connexin 43)

E-C Excitation-contraction (coupling)

ECM Extracellular matrix

EGF Epidermal growth factor

EGFR EGF receptor

ER Endoplasmic reticulum

EST Expressed sequence tag

FABP Fatty acid binding protein

FFA Free fatty acid

FGF Fibroblast growth factor (e.g. FGF-2= Fibroblast growth factor 2)

GATA G-A-T-A zing finger transcription factor family

GLUT Glucose transporter (e.g. GLUT4= Glucose transporter 4)

GLUT4-KO GLUT4 deletion mouse model (collective term includes all littermates,

LL + LLC)

GO Gene ontology

GPCR G protein-coupled receptor

HSP Heat shcok protein

xxii

ICaL L-type calcium current

IK Delayed rectifier potassium current

INCX Inward current mediated by the NCX

IDDM Insulin-dependent (i.e. Type 1) diabetes mellitus

IGF Insulin-like growth factor (e.g. IGF-I= Insulin-like growth factor I)

IL Interleukin

KO Knock-out

Lo Cell resting length

LL Lox+/+ Cre-/- (i.e. control) littermate from the GLUT4-KO mouse line

LLC Lox+/+ Cre+/- (i.e. knock-out) littermate from the GLUT4-KO mouse line

LQT Long QT syndrome

LV Left ventricle

MAPK Mitogen-activated protein kinase

MEF Myocyte-specific enhancer factor

MHC Myosin heavy chain (e.g. β-MHC= β-myosin heavy chain)

MI Myocardial infarction

MLC Myosin light chain

MMP Metalloproteinase

MRL Maximal rate of cell lengthening

MRS Maximal rate of cell shortening

MW Molecular weight

NaKATPase Sodium-Potassium ATPase

NCX Sodium-Calcium exchanger

NHE Sodium-Hydrogen Exchanger (e.g. NHE-1= Sodium-Hydrogen

Exchanger 1)

NIDDM Non insulin-dependent (i.e. Type 2) diabetes mellitus

Nkx Homeobox gene family

NO Nitric Oxide

iNOS Inducible nitric oxide synthase

PDGF Platelet-derived growth factor

PKC Protein kinase C

PPARα Peroxisome proliferator activated receptor α

RAS Renin-angiotensin system

xxiii

ROS Reactive oxygen species

RTK Tyrosine-kinase receptor

RV Right ventricle

RyR Ryanodine receptor

%S Maximum cell shortening (% resting length)

SERCA2 Calcium ATPase of the sarcoplasmic reticulum

SHR Spontaneously hypertensive rat (strain)

SR Sarcoplasmic reticulum

SRF Serum response factor

SRP Signal recognition particles

STZ Streptozotocin

Tf Time at return to resting cell length

Tf - Tm Duration of cell lengthening

Tf - To Duration of contractile cycle

Tm Time at maximum shortening

Tm - To Duration of cell shortening

To Excitation-contraction coupling latency

TAC Trans-aortic constriction

TG Transgenic littermate from the TG1306/1R mouse line

TG1306/1R Transgenic model (collective term includes all littermates, TG and WT)

TGF Transforming growth factor (e.g. TGF-β1= Transforming

growth factor-β1)

TNF Tumor necrosis factor (e.g. TNF-α= Tumor necrosis factor alpha)

%tsa Percent total sectional area

Wnt Wnt genes (drosophila melanogaster)

WT Wild-type (i.e. control) littermate from the TG1306/1R mouse line

identical with C57BL6

xxiv

PUBLICATIONS

Refereed Papers:

Containing material presented in this Thesis:

Domenighetti AA, Wang Q, Egge Mr, Richards SM, Pedrazzini T and Delbridge LMD

(2005). Angiotensin II-mediated phenotypic cardiomyocyte remodeling leads to age-

dependent cardiac dysfunction and failure. Hypertension, 46: 426-432.

Kaczmarczyk SJ, Andrikopoulos S, Favaloro J, Domenighetti AA, Dunn A, Ernst M,

Grail D, Fodero-Tavoletti M, Huggins CE, Delbridge LMD, Zajac JD and Proietto J

(2003). Threshold effects of GLUT4 deficiency on cardiac glucose uptake and

development of hypertrophy. J. Mol. Endocrinol., 31(3): 449-459.

Additional relevant publications:

Porrello ER, Huggins CE, Curl CL, Domenighetti AA, Pedrazzini T, Delbridge LMD,

Morgan TO (2004). Elevated dietary sodium intake exacerbates myocardial

hypertrophy associated with cardiac-specific overproduction of angiotensin II. J. Renin

Angiotensin Aldosterone Syst., 5(4): 169-175.

Lax CJ, Domenighetti AA, Pavia JM, Di Nicolantonio R, Curl CL, Morris MJ,

Delbridge LMD (2004). Transitory reduction in angiotensin AT2 receptor expression

level in post-infarct remodeling in rat myocardium. Clin. Exp. Pharmacol. Physiol.,

31(8): 512-517.

xxv

Huggins CE, Domenighetti AA, Pedrazzini T, Pepe S, Delbridge LMD (2003).

Elevated intracardiac angiotensin II leads to cardiac hypertrophy and mechanical

dysfunction in normotensive mice. J. Renin Angiotensin Aldosterone Syst., 4(3): 186-

190.

Communications:

Domenighetti, Ritchie, Smyth, Pedrazzini, Proietto, Delbridge (2004). Gene

expression profiling reveals distinct sets of genes altered during hormonally and

metabolically induced cardiac hypertrophies. J. Mol. Cell. Cardiol., 37: 303.

Porrello, Morgan, Huggins, Domenighetti, Pedrazzini, Delbridge (2004). Effect of

different sodium intakes on cardiac size in mice producing angiotensin II locally in the

heart. J. Hypertension, 22 (Suppl I): S10.

Domenighetti, Ritchie, Wang, Proietto, Smyth, Delbridge, Pedrazzini (2003). Gene

expression profiling reveals both common and distinct sets of genes altered during

hormonally and metabolically induced cardiac hypertrophy. Proc. Annual meeting of

the Swiss Cardiovascular Research and Training Network, Bern - Switzerland.

Huggins, Domenighetti, Pedrazzini, Pepe, Delbridge (2003). Chronic overexpression

of cardiac angiotensin suppresses basal ex vivo heart function and cardiomyocyte

contractility in hypertrophic transgenic mice. European Soc Cardiol Congress, Vienna,

Aug 30-Sept 3. Abs 22548.

Huggins, Domenighetti, Pedrazzini, Pepe, Delbridge (2003). Effects of elevated

cardiac AngII on ventricular function in mice. Proc. I.S.H.R., Melbourne, Australia.

August 7-9. J.Mol.Cell.Cardiol.

Domenighetti, Wang, Pedrazzini, Delbridge (2002). Chronic Ang II-induced

stimulation in the heart generates E-C coupling dysfunctions and leads to heart failure.

Annual meeting of the Swiss Cardiovascular Research and Training Network, Bern -

Switzerland.

xxvi

Domenighetti, Pedrazzini, Delbridge (2002). Chronic overproduction of cardiac Ang II

in transgenic mice induces cardiac hypertrophy, E-C coupling abnormalities and heart

failure. Gordon Conference on Cardiac Regulatory Mechanisms, Connecticut, - USA.

Domenighetti, Pedrazzini, Delbridge (2001). Abnormal contractile function in

cardiomyocytes from transgenic mice overexpressing Ang II in the heart. Proc. Satellite

congress to the IUPS on the Renin-Angiotensin-Aldosterone System, Melbourne -

Australia.

Domenighetti, Andrikopoulos, Kaczmarczyk, Favaloro, Zajac, Proietto, Delbridge

(2001). A deficiency in GLUT4 transporters induces fibrotic and hypertrophic cardiac

remodelling. World meeting of the International Union of Physiological Sciences

(IUPS), Christchurch - New Zealand.

Domenighetti AA (2001). ‘Myocardial remodeling in diabetic and metabolic-induced

cardiac hypertrophy’. Invited seminar at the Department of Medicine, Royal Melbourne

Hospital, Melbourne - Australia.

CHAPTER I

General introduction and literature review

Chapter I

2

1. INTRODUCTION

Cardiac hypertrophy, clinically identified as an increase in ventricular wall thickness, is

a common feature of individuals suffering from various cardiovascular or metabolic

conditions, such as hypertension, myocardial infarction or type II diabetes. In addition

to pathological conditions, in healthy populations and in athletes there is also

considerable genetic and physiological variation in cardiac mass. In general, cardiac

hypertrophy is associated with microscopic and macroscopic rearrangements of the

normal structure of the heart. This is linked with differential gene expression, leading to

cellular and interstitial changes occurring within the cardiac tissue. As a response to

these alterations the cardiac tissue changes in shape and function, and eventually

progresses toward failure. Although the multiple mechanisms underlying the transition

to heart failure are not fully understood and characterized, novel key concepts and

definitions have emerged in support of the idea that failure is not a simple matter of

reduced contractility and pump dysfunction. Heart failure is increasingly recognized to

be a highly complex clinical syndrome incorporating multiple extra-cardiac and intra-

cardiac features, including differential gene expression, neuro-endocrine activation and

changes in metabolism. Since cardiac hypertrophy development comprises many of the

changes associated with progressive heart failure, there is general acceptance that an

understanding of the molecular, structural and functional mechanisms that cause the

heart to remodel and to increase in size will largely contribute to knowledge of the

mechanisms underlying heart failure.

In this regard, experimental genetically manipulated animal models provide valuable

experimental tools for elucidating the molecular and functional mechanisms

responsible for the development of cardiac hypertrophy and the transition from the state

of compensated hypertrophy to dilation and failure. In particular, genetic animal

models of blood pressure-independent cardiac hypertrophy (i.e. of primary cardiac

hypertrophy) have been generated to address the question of whether cardiac-specific

neuro-endocrine activation and metabolic imbalance are necessary and/or sufficient to

induce the heart to remodel.

Chapter I

3

In the present work, an examination of two different genetic mouse models of primary

cardiac hypertrophy is presented: 1) a cardiac-specific angiotensinogen-overexpressing

transgenic mouse, the TG1306/1R; and 2) a glucose transporter (GLUT4) knock-out

mouse model, the GLUT4-KO. Experimental investigations included cellular and tissue

morphometric analysis on the heart of these rodents (Chapter III), combined with

experiments of cardiomyocyte contractility (Chapter IV) and cardiac gene expression

profiling by cDNA microarray assays (Chapter V). The objective was to provide new

insight into the mechanisms that are common and unique in the development of

hypertrophic states induced either by the cardiac-overproduction of the bioactive

peptide angiotensin II (Ang II) or a cardiomyocyte-specific glucose metabolic

deficiency (Chapter VI). In this initial Chapter a literature survey and the general

concepts concerning the mechanisms and the stimuli that lead to cardiac hypertrophy

and heart failure are presented, while Chapter II describes the general methods applied

in the experimental studies to follow (Chapter III- V).

Chapter I

4

2. DEFINITION OF CARDIAC HYPERTROPHY

Cardiac hypertrophy can be defined as a physiologic and pathologic condition where

the heart changes in size, shape and function in response to a variety of extra-cardiac

and intra-cardiac stimuli. Cardiac hypertrophy is itself a predictor of cardiovascular

morbidity and mortality, independent of hypertension and coronary diseases (Koren et

al., 1991; Levy et al., 1996; Schmieder and Messerli, 2000; Drazner et al. 2004). In

general, cardiac hypertrophy involves two major components of the heart: the cellular

fraction and the extracellular matrix (ECM). The degree of myocardial remodeling is

mainly determined by changes in collagen concentration or in collagen isoform

expression and cell growth (hypertrophy), proliferation (hyperplasia) or death (by

apoptosis or necrosis).

Cardiac hypertrophy is often associated with or even induced by other intra-cardiac

events, including cell necrosis, ischemia, infarction, interstitial fibrosis and vascular

sclerosis. The normal myocardial wall architecture is remodeled and the

cardiomyocytes increase either in width or in length or in both directions. Clinically,

there are two forms of anatomic cardiac hypertrophy that could progressively lead to

heart failure: these are named concentric and eccentric hypertrophies (Hunter and

Chien, 1999; Morgan and Delbridge, 1999).

2.1. Concentric cardiac hypertrophy

This form of hypertrophy is often associated with hypertension and other states of

pressure overload (e.g. aortic stenosis). It is generally believed that in this case

cardiomyocytes tend to increase in width rather than length (Hunter and Chien, 1999).

This is due to the assembling of extra contractile-protein units in parallel. Wall

thickening reduces cardiac chamber size and decreases stroke volume as a

Chapter I

5

consequence. Concentric hypertrophy is initiated by increased diastolic stretch.

Abnormal ventricular relaxation impedes atrial emptying of blood, which can passively

stretch these thin-walled chambers and cause atrial dilation, a clinically prevalent

feature of the concentric hypertrophic heart. Moreover it has been demonstrated that

concentric cardiac hypertrophy is a significant predictor of coronary heart disease (de

Simone and Palmieri, 2002; Malmqvist et al., 2002), depressed midwall systolic

function (Schillaci et al., 2002), increased QT interval duration (Oikarinen et al., 2001),

microalbuminuria in hypertensive patients (Tsioufis et al., 2002), and increased

incidence of pathologic events in patients with uncomplicated myocardial infarction

(Carluccio et al., 2000).

2.2. Eccentric cardiac hypertrophy

This form of cardiac hypertrophy is generally associated with volume overload states.

During eccentric hypertrophy, the cardiomyocytes are believed to increase in length

rather than width, because extra contractile-protein units are assembled in series.

Eccentric hypertrophy results from increased systolic stress (Katz, 2002) and is almost

always accompanied by an increase in ventricular wall thickness, which reflects

myocyte hypertrophy as well as a variable increase in interstitial fibrosis. In athletes a

physiologic eccentric hypertrophy can occur, which is caused by an increase in

workload. In this case, hypertrophy is not acutely associated with a particular

pathology, although some long-term clinical implications after deconditioning cannot

be excluded (Neri Serneri et al., 2001; Pelliccia et al., 2002). Generally, the intra-

ventricular volume expansion stretches cardiomyocytes and improves pressure-volume

relationships within the heart so as to augment cardiac output. Eccentric ventricular

hypertrophy has also been associated with left atrial enlargement in hypertensive

patients (Gerdts et al., 2002) and increased QT interval duration (Oikarinen et al.,

2001).

Chapter I

6

2.3. Ventricular dilation

Ultimately, eccentric and concentric forms of cardiac hypertrophies may degenerate

into ventricular dilation. In dilated hypertrophy regional myocardial injury and ageing

is associated with progressive and deleterious thinning of the ventricular wall and

impaired systolic function. Studying the molecular and physiological pathways leading

to acquired concentric and eccentric cardiac hypertrophy and cardiac dilation is a

complicated matter. This is generally attributable to the nature of the multifactorial

origin of these forms of hypertrophy, which are often linked with and triggered by

multiple underlying pathologies.

There is evidence suggesting that concentric and eccentric forms of remodeling and

ventricular dilation share some common important intracellular signaling pathways. For

example, transgenic and cardiac-specific over-expression of tropomodulin generates

mice with dilated cardiomyopathy (Sussman et al., 1998a). Interestingly, the dilated

phenotype can be prevented by treating the same mice with inhibitors of the calcineurin

pathway, a well-known stimulus for concentric cardiac hypertrophy (Sussman et al.,

1998b). Cross-breeding of the tropomodulin transgenic mice in heterozygous and

homozygous forms with calcineurin over-expressing mice demonstrate that different

cardiac hypertrophic phenotypes (that is, concentric and eccentric) coexist with the

same genetic background (Sussman et al., 2000). These results also suggest that

distinctions made between concentric, eccentric and dilated forms of hypertrophy to

explain different clinical pathologies of the heart may be oversimplified, especially

when molecular and cellular mechanisms are taken in account.

Chapter I

7

3. FAMILIAL FORMS OF HEART DISEASES

In recent years there has been significant progress in delineating causes of hereditary

heart diseases and primary cardiomyopathies. Today it is widely accepted that most

congenital ‘idiopathic’ heart defects are known to result from heritable gene mutations.

This is exemplified by the recurrence of heart defects in the same family. Identification

of mutations in patients with cardiomyopathies and various other heart malformations

has revealed substantial molecular complexity in the etiologies of these disorders.

3.1. Familial hypertrophic cardiomyopathy

In contrast to acquired forms of cardiac hypertrophy, primary/familial hypertrophic

cardiomyopathies are not secondary to other underlying pathologies. This form of

hypertrophy is an autosomal dominant disease of the cardiac muscle that is

characterized by disproportionate symmetric or asymmetric hypertrophy of the left

ventricle and cardiomyocyte disarray. Cardiac mass is increased due to left ventricular

wall thickening, often with particular involvement of the interventricular septum. As a

consequence of hypertrophy, the left ventricular chamber volumes are diminished,

resulting in the appearance of a muscle-bound heart. Despite the presence of even

markedly abnormal ventricular morphology and histopathology, systolic function in

familial hypertrophic cardiomyopathy is usually excellent and can often appear supra-

normal. Yet, most affected individuals develop mild to moderate symptoms of

shortness of breath (dyspnea) and chest pain (angina) due to impaired diastolic

relaxation of the hypertrophic heart (Spirito et al, 1997). In addition, patients are at

increased risk of developing heart failure, atrial and ventricular arrhythmias, and

sudden death. Indeed, hypertrophic cardiomyopathy is the most common cause of

sudden death in the young, accounting for 48% of sudden death in individuals less than

35 years of age (Basso et al., 2001). Furthermore, unrecognized familial hypertrophic

Chapter I

8

cardiomyopathy is the most common cause for sudden death in athletes (Maron et al.,

1996; Maron, 2003).

Several genes have been mapped to familial hypertrophic cardiomyopathy. The

majority encode sarcomeric proteins in the heart muscle (e.g. β-myosin heavy chain,

cardiac troponin T, tropomyosin, actinin) (reviewed in Bonne et al., 1998). Mutations

in cardiac actin, troponin I and titin are also currently recognized as rare causes of

familial hypertrophic cardiomyopathy. An important corollary to the obvious

conclusion that sarcomere mutations in hypertrophic cardiomyopathy alter the

molecular processes of muscle contraction is that these events also activate pathways

for myocyte growth. Although little is currently known about signaling pathways that

link sarcomere force production to myocyte growth, and ultimately to cardiac

hypertrophy, evidence suggests a critical role for calcium in signal transduction. For

instance, it has been demonstrated that myosin mutations cause fundamental

dysregulation of sarcomere calcium requirement and cycling (Fatkin et al., 2000).

3.2. Dilated cardiomyopathy

Defined by ventricular dilation and diminished contractile function, dilated

cardiomyopathy is a prevalent world-wide disorder that is estimated to affect 40-50

cases per 100,000 (Codd et al., 1989). Dilated cardiomyopathy causes heart failure,

serious arrhythmias and thromboembolic events, all of which account for the premature

mortality of the disease (Elliot, 2000). Although cardiac mass is increased, there is

often only modest ventricular wall hypertrophy while atrial and ventricular chambers

can be distended. Myocyte hypertrophy and interstitial fibrosis are difficult to observe.

Unlike familial hypertrophic cardiomyopathy, substantial distortion of cell architecture

or myocyte disarray is not a feature of dilated cardiomyopathy. Early clinical

manifestations of the disease are vague and often not attributed to heart failure. When

underlying causes such as coronary artery disease, chronic alcohol abuse, thyroid

disease or viral infection are excluded as etiologies, a diagnosis is often made of

‘idiopathic dilated cardiomyopathy’ (Kasper et al., 1994). Familial studies and

echocardiography of relatives of affected individuals have demonstrated that 25-30% of

Chapter I

9

‘idiopathic’ dilated cardiomyopathy is caused by inherited gene mutations (Grunig et

al., 1998; Seidman and Seidman, 2001). Interestingly, left ventricular dilation and

systolic dysfunction develop in ~15% of patients with familial hypertrophic

cardiomyopathy.

3.3. Congenital heart diseases

Congenital heart diseases are a group of structural cardiac anomalies present at birth,

affecting ~0.4% of live born infants (Ferencz et al., 1985; Roskes et al., 1990). These

are caused by abnormal cardiac morphogenesis and development. Such disorders

include defects in the valves and chambers associated with specific chromosomal

abnormalities (e.g. trisomy 21/Down syndrome or Noonan syndrome) and with genetic

dysmorphic syndromes. As an example, the search for the keyword ‘congenital heart

disease’ on the internet-based databank OMIM (Online Mendelian Inheritance in Men,

found at http://www.ncbi.nlm.gov/), reveals more than 300 entries corresponding to

tens of chromosome loci and hundred of potentially related genes, the mutation of

which could lead to various pathologic cardiac phenotypes, such as septation defects,

stenosis, neural crest defects, tetralogy of Fallot, supravalvular aortic stenosis, or

coronary artery obstructive lesions. Examples of syndromes that have been linked to

specific genetic mutations are the Holt-Oram syndrome (linked to the genes TBX-5 and

NKX2.5), the Williams syndrome (linked to the gene for elastin), the Alagille

syndrome (linked to mutations in JAG1), Down syndrome (linked to region on

chromosome 21 that contains about 28 known genes) and the Noonan syndrome (linked

to PTPN11).

3.4. Cardiac arrhythmias

Cardiac arrhythmias are a common cause of morbidity and mortality and may be

associated with genetic defects or altered expression in calcium, potassium or sodium

channels (Basso et al., 2001). Long QT (LQT) syndrome, a condition of cardiac

Chapter I

10

arrhythmias, is characterized by prolongation of the QT interval on electrocardiograms

(Wang et al., 1998; Basso et al., 2001). Various forms of inherited long QT have been

reported, including the autosomal recessive LQT of the Jervell-Lange-Nielsen

syndrome (with mutation found in the potassium channels KCNQ1 and KCNE1), the

autosomal dominant LQT of the Romano-Ward syndrome (with mutations in the

potassium channel KCNQ1) and the autosomal dominant LQT of the Andersen

cardiodysrhythmic periodic paralysis (caused by mutations in the potassium channel

KCNJ2).

3.5. Coronary diseases

Coronary artery diseases are characterized by a gradual accumulation of vascular

plaque materials including inflammatory cells and molecules, cholesterol, complex

lipids and lipoproteins and varying amounts of fibrous connective tissue containing

numerous smooth-muscle cells, collagen, T-cells and others. The clinical symptoms of

coronary heart disease, such as angina pectoris, myocardial infarction and death, result

from mild to moderate atherosclerotic lesions (atherosclerosis) that progress from fatty

streaks to fibrous plaques, and ultimately severe plaque rupture. The majority of

coronary heart disease cases are considered to be multifactorial and result from the

interaction of multiple genetic and environmental factors. Various genes, implicated in

conferring an increased risk of coronary heart disease and myocardial infarction,

include the angiotensin-converting enzyme (ACE) (Cambien et al., 1992; Ruiz et al.,

1994; Schachter et al., 1994), angiotensinogen (Agt) (Katsuya et al., 1995) and genes

regulating lipoprotein metabolism (Schachter et al., 1994).

Chapter I

11

4. HYPERTROPHY AS A PRELUDE TO

HEART FAILURE

4.1. Definition of heart failure

Although the term chronic heart failure is commonly used, there is no clear definition

for this condition. The American College of Cardiology and the American Heart

Association (ACC/AHA) joint task force ‘Guidelines for the evaluation and

management of heart failure’ defines heart failure as a “complex clinical syndrome that

can result from any structural or functional cardiac disorder that impairs the ability of

the ventricle to fill with and eject blood. The cardinal manifestations of heart failure are

dyspnea and fatigue, which may limit exercise tolerance, and fluid retention, which

may lead to pulmonary and peripheral edema” (ACC/AHA, 2001). However, such

pathophysiological conceptualization of heart failure is rapidly evolving currently to

take into account multiple extra- and intra-cardiac features, including myocardial

remodeling, metabolic disturbances, neuro-humoral activation and cytokine release.

It is now apparent to physicians and researchers that patients often go through a period

of latent or even asymptomatic left ventricular hypertrophy and diastolic dysfunction

before the development of overt signs and symptoms of heart failure (Vasan et al.,

1997). In many cases the development of cardiac hypertrophy leading to heart failure is

preceded by or associated with a cascade of ‘succeeding events’. These events can take

multiple forms, namely 1) an acute insult, like a myocardial infarction, a myocarditis or

a vessel atherosclerotic disease; 2) hemodynamic changes, such as the gradual

development of systemic hypertension or the onset of a valvular insufficiency; 3)

neuro-humoral activation, involving the upregulation of hormones, cytokines and

neuropeptides; 4) a metabolic imbalance, caused by hyperlipidemia, hyperglycemia,

insulin resistance, diabetes and obesity; 5) other systemic diseases, such as chronic

obstructive lung diseases, renal diseases or hyperthyroidism; and finally 6) the

Chapter I

12

expression of mutated genes, as in idiopathic and familial forms of hypertrophic and

dilated cardiac remodeling. In synthesis, heart failure is multifactorial and in many

cases its genesis is determined by the coexistence and the auto-induction of one or

more of the above events occurring in association with the process of normal

senescence (Smith, 1985; Kannel and Belanger, 1991; Julien, 1997; Adams, 2001).

4.2. Transition from cardiac hypertrophy to heart failure

Cardiac and cardiomyocyte hypertrophy in response to pathologic conditions has

traditionally been considered a compensatory and adaptive response required to sustain

cardiac output in the face of new pathologic working conditions. Prolonged or chronic

hypertrophy however is associated with a significant increase in the risk for decreased

cardiac output, development of ventricular arrhythmias, myocardial dilation and sudden

death, independently of the causes triggering hypertrophy (Levy et al., 1990; Vakili et

al., 2001). This notion is further confirmed by observations made in clinical trials

where inhibition or even regression of cardiac hypertrophy by certain drugs, such as

angiotensin-converting enzyme inhibitors (ACE-I), lowers the risk for several

endpoints, including progression to heart failure and death, whereas persistence of

cardiac and cardiomyocyte hypertrophy predicts adverse outcomes (Beltrami et al.,

1994; Mathew et al., 2001). According to the previous definition of heart failure

(“ventricular dysfunction with symptoms”), the transition from compensated

hypertrophy to failure could therefore indicate the limits of the molecular and

phenotypic adaptation process of the heart to pathologic working conditions.

A search for molecular markers characterizing heart failure and the transition of

compensated hypertrophy to heart failure is only beginning. Among potential clinical

markers of heart failure are the circulating natriuretic peptides, such as the atrial and

brain natriuretic peptides (ANP and BNP). These molecules are secreted by the atria

and the ventricles respectively to maintain normal cardiovascular and renal homeostasis

(Clerico and Emdin., 2004). Recent clinical studies have demonstrated that both ANP

and BNP are potentially useful in diagnosis and prognosis of heart failure, where

plasma concentration of these peptides is increased (Maisel et al., 2003). In addition,

Chapter I

13

BNP also appears to be useful in assessing the risk and predicting the outcome in

patients with myocardial infarction and failure. Natriuretic peptides are the only US

Food and Drug Administration (FDA)-cleared biomarkers for diagnosis of heart failure

to date. In addition to natriuretic peptides, plasma concentrations of various other

molecular markers have been reported to be disturbed in heart failure patients. Some of

these include mammalian cardenolides (Jortani and Valdes, 2001), various cytokines

such as interleukin-6 and tumor necrosis factor alpha (IL-6 and TNF-α) (Cicoira et al.,

2001; Parissis et al., 2003), troponin I and T (Missov and De Marco, 1999),

adrenomedullin (Tsuruda et al., 2003) and leptin (Filippatos et al., 2000; Perrego et al.,

2005).

The majority of intra-cardiac molecules which would allow characterization of the

transition from compensated to decompensated states are probes specific for the

extracellular matrix (ECM) component of the heart, such as fibronectin and the shift

between types of collagen isoforms (Pauschinger et al., 1999). The expression of

transforming growth factor-beta 1 (TGF-β1) and osteopontin (Pauschinger et al., 1999;

Singh et al., 1999), as well as the activation of metalloproteinases (MMPs) (Iwanaga et

al., 2002) are also key markers of the transition from cardiac hypertrophy to heart

failure. Among the cytoskeletal and sarcomeric changes that could enhance the

transition toward failure, depressed microtubule depolymerization (Tagawa et al.,

1998) and changes in troponin I and T function (Noguchi et al., 2003) are proposed.

Finally, cardiac beta-adrenergic desensitization (Castellano and Bohm, 1997; Tse et al.,

2000), increased protein Gαq-dependent myocyte apoptosis (Dorn and Hahn, 2004),

inducible nitric oxide synthase (iNOS) expression and nitric oxide (NO) production

(Horinaka et al., 2003), as well as a shift in sodium/potassium ATPase (NaKATPase)

isoforms (Fedorova et al., 2004) could be directly involved in the phenotypic and

mechanical changes observed in the heart during the transition to failure.

Experimentally, the age-dependent transition from compensated cardiac hypertrophy to

heart failure has been rarely studied due to the absence of valid experimental models.

Spontaneously hypertensive rats (SHR) develop impaired myocardial function and

biventricular congestive heart failure, after a long period (18 months) of compensatory

cardiac hypertrophy due to systemic hypertension (Bing et al., 2002; Boluyt et al.,

Chapter I

14

1995). A similar phenotype can be obtained in chronic pressure overload introduced

gradually in young normotensive rats after aortic banding (Boluyt et al., 2005).

Badenhorst et al. (2003) showed that chronic beta-adrenergic receptor activation

initiates the progression from compensated concentric hypertrophy to cardiac

dysfunction primarily through tissue remodeling in SHR. Dahl salt-sensitive rats fed a

high-salt diet after the age of 6 weeks develop concentric left ventricular hypertrophy at

11 weeks and heart failure with ventricular dilation at 15-20 weeks (Morii et al., 1998).

In this model, the heart failure transition is associated with a marked decrease in

myocardial contractility and with ventricular remodeling. Stenosis of the descending

aorta in guinea pigs lead to similar results, with compensated hypertrophy for up to~10

weeks after banding and the presence of heart failure after 20 weeks (Randhawa and

Singal, 1992).

4.3. Incidence and clinical causes of heart failure

Recent estimates highlight the high incidence of heart failure throughout the Western

world (Cubillos-Garzon et al., 2004; Lee et al., 2004; Remme et al., 2004). It is the only

cardiovascular disease of rapidly increasing prevalence. World-wide, it is estimated

that 3.6 million new cases occur each year. Reasons for the rising incidence of chronic

heart failure include an increased awareness amongst clinicians leading to an increase

in the rate of diagnoses, an increase in the rate of survival after an acute event (such as

myocardial infarction) as a result of recent therapeutic developments (AIRE Study

Investigators, 1993), and the advancing age of the population in the Western world

(Steward et al., 2003). The prevalence ranges from 1-2% in the middle-aged to 10-15%

in octogenarians (Cowie et al., 1997). Amongst other risk factors, hypertension,

coronary artery diseases, ischemia, left ventricular hypertrophy, sympathetic

overactivity and type 2 diabetes are important risk factors for the development of heart

failure (Wei et al., 1998; Arnold et al., 2003; Drazner et al., 2004; Elhendy et al., 2004;

Palatini and Julius, 2004).

Chapter I

15

5. ALTERED EXCITATION-CONTRACTION COUPLING

IN CARDIAC HYPERTROPHY AND FAILURE

Contraction of the cardiac muscle cell is initiated by transmission of spontaneously

generated electrical action potentials. Upon sodium-dependent cardiomyocyte

membrane depolarization (Figure I-1, A.), calcium enters the cytosol mainly through

voltage-operated calcium channels generating L-type calcium currents (ICaL) (Bers,

2001) (Figure I-1, B.). This calcium entry triggers calcium release from the

sarcoplasmic reticulum (SR). This step is regulated by the activation and inactivation of

calcium release channels, namely the ryanodine receptors (RyR) of the SR (Bers, 2001;

Marks, 2001), a process termed calcium-induced calcium release (CICR) (Figure I-1,

C.). Under certain conditions, calcium can also enter the cell via sarcolemmal

sodium/calcium exchanger (NCX) (Figure I-1, D.). Application of laser scanning

confocal microscopy techniques and calcium-sensitive fluorescent indicators has made

it possible to sustain the hypothesis that the initial phase of excitation-contraction (E-C)

coupling in cardiac cells is locally regulated by spatial and temporal

summation/coupling of single ICaL, NCX operation and SR calcium ‘sparks’ (Egger and

Niggli, 1999; Berridge et al., 2000; Niggli and Egger, 2002). These processes raise

intracellular calcium concentrations from ~100 nM to ~1 μM, causing calcium binding

to troponin C which serves to remove the inhibition of the attachment of actin to

myosin and produces shortening of the cell (Figure I-1, E.).

For relaxation to occur, calcium must be removed from the cytoplasm. The main means

of lowering cytoplasmic calcium to terminate contraction make use of active pumping

of calcium into the SR by the calcium ATPase SERCA2 (Figure I-1, F.). A component

of calcium is also removed from the cell by the NCX (Negretti et al., 1993; Bassani et

al., 1994) (Figure I-1, G.). The relative contribution of the NCX in calcium extrusion is

species- and ICaL density-dependent (Bassani et al., 1994; Bers, 2001). In addition, the

relative contribution of SERCA2 and NCX to elimination of cytosolic calcium can also

change upon altering steady-state conditions, such as increasing stimulation frequency

Chapter I

16

(Pieske et al., 1999; Maier et al., 2000). Using paired rapid cooling contractures Maier

et al. (2000) showed that increasing stimulation frequency from 0.25 to 3 Hz in rabbit

myocardium increased the apparent SERCA2 contribution from 28% to 65%, while in

rats this contribution increased only from 62% to 70%, suggesting that in the latter

species SERCA2-dependent calcium removal was already nearly maximal at low

frequency. Active extrusion of calcium across the sarcolemmal membrane by

sarcolemmal ATPase and mitochondrial uptake also occurs but it represents only 2-3%

of the total calcium transport during relaxation (Bassani et al., 1992; Berridge et al.,

2000) (Figure I-1, H.). In parallel to calcium-dependent relaxation, recovery of

excitability by repolarization of the cell membrane to resting levels (around -90 mV) is

achieved by the activation of potassium-dependent hyperpolarizing currents (Figure I-

1, I.).

5.1. Electrical changes in cardiac hypertrophy and heart failure

In recent years, electrical remodeling has emerged as an important pathophysiological

mechanism in the development of mechanical dysfunction, pump failure and sudden

death. When detected by electrocardiography, cardiac hypertrophy and heart failure are

commonly associated with the development of QT interval prolongation or dispersion

and life-threatening arrhythmias (Kulan et al., 1998; Pogwizd and Bers, 2004). Clinical

studies suggest that 50% of sudden deaths associated with heart failure are attributable

to ventricular tachycardia or fibrillation, the other half being coupled to

bradyarrhythmias or electromechanical dissociation (Janse, 2004). Mechanisms

underlying these arrhythmias are multifactorial, but they stem from disordered

electrical currents arising from differential expression or activity of ion channels,

increasing action potential duration. The resulting delay in the recovery of excitability,

a consistent feature of ventricular hypertrophy, predisposes to delayed

afterdepolarizations, especially at higher frequencies (Hart, 1994; Armoundas et al.,

2001; Janse, 2004; Pogwizd and Bers, 2004). The most consistent electrophysiological

changes associated with action potential prolongation and triggered activity involve 1)

abnormalities in intracellular calcium handling, including an increase in NCX activity;

2) a reduction in potassium currents, including the transient outward (Ito) and the

Chapter I

17

inward rectifier (IK1) currents; and 3) impaired beta-adrenergic stimulation. Additional

mechanisms include myocardial fibrosis, altered cell-to-cell junctional communication,

myocardial electrical and cellular heterogeneity, all of which predispose to re-entrant

mechanisms of arrhythmia.

5.2. Calcium cycling in cardiac hypertrophy and heart failure

It has been proposed that calcium cycling plays a major role in the development of

electrical and mechanical dysfunction in cardiac hypertrophy and failure (Wickenden et

al., 1998; Bers, 2001). Calcium-dependent cardiomyocyte contractility is indeed

affected during cardiomyocyte remodeling. This aspect will be detailed in Chapter IV,

but in general terms mechanical dysfunction in decompensated cardiac hypertrophy and

failure is normally paralleled by time- and amplitude-dependent modifications of the

calcium transient (Beuckelmann et al., 1992; Wickenden et al., 1998), and the

expression or activity of several calcium-handling proteins, including SERCA2 (Dipla

et al., 1999; Lim et al., 1999; Terracciano et al., 2001; Zhong et al., 2001; Hasenfuss

and Pieske, 2002), L-type calcium channels (He et al., 2001; Chen et al., 2002), RyR

receptors (Marks et al., 2002) and NCX (Dipla et al., 1999; Hasenfuss and Pieske,

2002; Sipido et al., 2002).

5.2.1. Calcium cycling in compensated cardiac hypertrophy

In general, there is evidence to suggest that modification of the calcium signaling is

dependent on the stage and severity of cardiac hypertrophy. In compensated

experimental cardiac hypertrophy, differential gene expression of calcium handling

proteins is modest or unchanged (De la Bastie et al., 1990; Feldman et al., 1993; Kiss et

al., 1995; Arai et al., 1996), while ICaL, systolic calcium and calcium transients are

generally increased (Brooksby et al., 1993; Kagaya et al., 1996; Mukherjee and

Spinale, 1998; Armoundas et al., 2001; Wang et al., 2001). In compensated cardiac

hypertrophy, differences in systolic calcium that exist between hypertrophied and

normal cardiomyocytes can be abolished by voltage-clamp conditions, where the

Chapter I

18

duration of depolarization is normalized (Brooksby et al., 1993). This suggests that

action potential prolongation directly correlates with elevated systolic calcium and

increased contractility (inotropy) in these hypertrophied myocytes. Thus, in

compensated cardiac hypertrophy, action potential prolongation often stems from

upregulation of inward currents, such as ICaL, which is active in ‘phase 2’ of the action

potential, and the inward current mediated by the NCX (INCX), which is active in ‘phase

3’. It could be suggested that in compensated cardiac hypertrophy, a compensatory

prolongation of the action potential would increase the amplitude of the calcium

transient and lead to a positive inotropic response of the heart to the new working and

environmental conditions.

5.2.2. Calcium cycling in decompensated cardiac remodeling

Regardless of maintained action potential prolongation, in maladaptive cardiac

hypertrophy and in heart failure, systolic calcium flux is markedly reduced

(Beuckelmann et al., 1992; Pieske et al., 1995), ICaL maybe unchanged or sometimes

decreased, while the calcium transient is significantly prolonged and SR calcium

uptake activity reduced (Beuckelmann et al., 1992; Mukherjee et al., 1995; Delbridge et

al., 1997). In animal models of severe hypertrophy, these events correlate with

downregulation of SR calcium-handling proteins, such as RyR, SERCA2 and

phospholamban (De la Bastie et al., 1990; Naudin et al., 1991; Feldman et al., 1993; Go

et al., 1995; Kiss et al., 1995; Hittinger et al., 1999; Milnes and MacLeod, 2001). This

is also confirmed in human end-stage and dilated heart failure (Flesch et al., 1996;

Hasenfuss et al., 1997). Unlike the early changes that lead to action potential

prolongation, it appears that changes in SR calcium handling are maladaptive.

Therefore, an imbalance of these regulatory mechanisms, generally leading to a shift of

the circulation of calcium from intracellular routes (cytosol and SR) to extracellular

routes (trans-sarcolemmal), could be hypothetically responsible for time-dependent

impaired diastolic and systolic function and could promote the onset of failure. These

mechanisms will be further detailed in Chapter IV dealing with cardiac remodeling

induced by cardiac Ang II overproduction and diabetic cardiomyopathy.

Chapter I

19

5.2.3. The sodium/calcium exchanger (NCX)

The NCX importantly contributes to control of intracellular calcium concentrations,

exporting cytoplasmic calcium by electrogenically exchanging it for extracellular

sodium (Figure I-1). It catalyzes the bidirectional exchange of 3 sodium ions for a

single calcium ion. Thus, one net positive charge per reaction cycle moves inwardly

during relaxation.

The density of the NCX generated calcium flux is generally found to be increased or

unchanged in cardiac hypertrophy and heart failure in humans, experimental models

and in cultured hypertrophic cardiomyocytes (Studer et al., 1994; Reinecke et al., 1997;

Brooks et al., 2000; Hobai and O’Rourke, 2000; Wang et al., 2001). Forward-mode

exchanger function (sodium in and calcium out, during repolarization) compensates for

defective SR calcium removal at the expense of depletion of the SR releasable pool of

calcium with repetitive stimulation (Hasenfuss et al., 1996; Pieske et al., 1996), thus

increasing depolarizing current. Reverse mode exchange (sodium out and calcium in,

during depolarization) has been suggested to provide inotropic support to the failing

heart (Flesch et al., 1996; Mattiello et al., 1998).

The major exception to this is probably represented by models of diabetic and

metabolic cardiomyopathy, where the NCX has been consistently shown to be

downregulated (Hattori et al., 2000; Kashihara et al., 2000; Duan et al., 2003).

Interestingly some have reported a dissociation between NCX current (INCX) density

and NCX protein expression in compensated cardiac hypertrophy, where NCX1 – the

major exchanger transcript in the heart – is overexpressed to an extent similar to that

reported in heart failure, but INCX is decreased (Wang et al., 2001). Similar

downregulation of INCX was described in the G protein (Gαq)-overexpressing mouse

model of mild cardiac hypertrophy (Mitarai et al., 2000) and dissociation of NCX

protein and activity has been described in cardiac hypertrophy (Boateng et al., 2001).

These findings contrast with heart failure, in which most, although not all (Hasenfuss et

al., 1999) studies have documented increased exchanger activity in parallel to NCX1

Chapter I

20

mRNA or protein overexpression. In some experimental models the imbalance of this

current may therefore be related to the action potential prolongation and arrhythmias.

5.3. Potassium currents in cardiac hypertrophy and heart failure

As previously mentioned, repolarization in the mammalian heart is achieved primarily

by the activity of potassium-selective ionic currents, although the exact molecular

composition of these currents varies considerably from species to species. Ventricular

myocytes express several distinct classes of voltage-dependent potassium channels.

The inward rectifier potassium current (IK1) maintains the resting membrane potential

and contributes to the terminal phase of repolarization in the ventricular myocytes. An

important potassium current is the calcium-independent transient outward current, Ito.

Unlike the inward rectifier, Ito is expressed in heart cells in a species- and cell type-

specific fashion. This current plays a crucial role in the early phase of repolarization.

The delayed rectifier K current (IK), composed of molecularly distinct rapid (IKr) and

slow (IKs) components, is important in ‘phase 3’ of repolarization (reviewed in

Armoundas et al., 2001).

It is proposed that of all the mechanisms that may contribute to action potential

prolongation, depression of Ito appears to be the most reproducible (Beuckelmann et al.,

1993; Cerbai et al., 1994; Kaab et al., 1996; Li et al., 2002). The peak Ito density is also

reduced during aging and can represent a very early event in the response to decreased

pump performance. Nuss et al. (1996) reported that failing cardiomyocytes with 66%

reduction in Ito (and prolonged action potential duration) exhibit normalized

repolarization time when transfected with potassium channel genes (ShK).

Other altered potassium currents could also impact on the action potential profile of

hypertrophied cardiomyocytes, including IK1 (Brooksby et al., 1993; Kaab et al., 1996)

and the pacemaker current If. The latter is a current that contributes to diastolic

depolarization in pacemaking cells (i.e. sinus node and Purkinje cell) and neonatal

cardiomyocytes (Er et al., 2003). If activates slowly on repolarization and deactivates

rapidly with depolarization, supporting a mixed sodium and potassium current (van

Chapter I

21

Ginneken and Giles, 1991). In heart failure, hypertrophy, and atrial fibrillation, If

current densities and/or mRNA levels of its molecular correlate hyperpolarization-

activated cyclic nucleotide-gated (HCN) channels are also recorded in hypertrophic

ventricular myocytes and are increased compared with controls (Cerbai et al., 1997;

Hoppe et al., 1998; Cerbai et al., 2001; Fernandez-Velasco et al., 2003). These

observations support a potential contribution of If to the arrhythmogenesis of working

myocardium under pathological conditions (Zorn-Pauly et al., 2004; Michels et al.,

2005).

5.4. Sympathetic stimulation and electrical impairment

Decompensated cardiac hypertrophy and heart failure are fundamentally accompanied

by a chronic activation of the sympathetic nervous system and such increased activity

often precedes the onset of overt clinic symptoms of heart failure (Bohm et al., 1997;

Esler et al., 1997). The beta- and alpha-adrenergic signaling pathways are known to

significantly affect the function of a number of ion channels and transporters in the

heart. The net effect of beta-adrenergic stimulation is to shorten the ventricular action

potential duration due to an increase in the current density and a hyperpolarizing shift

of the activation of IK (Hartzell and Duchatelle-Gourdon, 1993; Terrenoire et al., 2005),

despite beta receptor stimulation of depolarizing current through the L-type calcium

channel. Alpha-adrenergic receptor stimulation inhibits several potassium currents in

the mammalian heart, including Ito, IK1 and IK in rat ventricle with the net effect of

prolonging action potential duration (Fedida et al., 1993) and promoting early

afterdepolarizations.

5.5. Cell-to-cell coupling and electrical communication

As previously mentioned, a characteristic consequence of hypertrophic remodeling is

the increased risk for fatal ventricular arrhythmias. Alterations in action potential

duration, disturbed calcium handling, and ventricular re-entry circuits arising from

Chapter I

22

regions of slow, inhomogeneous conduction, and tissue conduction block may

contribute to arrhythmogenesis. The observed alterations in cell-to-cell conduction of

electrical impulses may be due to changes in the expression pattern and composition of

the gap junction proteins. Gap junctions are mainly located in the intercalated discs of

cardiomyocytes and consist of multiple gap junction channels. A gap junction channel

is built of gap junction proteins (connexins); six connexins interact to form a connexon

(hemichannel) on one cell surface, which aligns head-to-head with a connexon on the

opposing cell surface together forming an intercellular channel (Saetz et al., 2003).

Twenty distinct connexin genes have been identified in the human genome. Three types

of connexins, connexin-43 (Cx43), -40 (Cx40), and -45 (Cx45), are expressed in heart

(Willecke et al., 2002). Cx43 is abundant in atrial and ventricular myocardium (Davis

et al., 1995). Cx40 is expressed in atrial myocytes, coronary vascular endothelium and

the His-Purkinje conducting system. Cx45 is observed in the sinoatrial and

atrioventricular nodes and small amounts colocalize with Cx43 in adult ventricular

myocardium (Coppen et al., 1998). A major role of gap junctions in the myocardium is

to enable rapid and coordinated electrical excitation, a prerequisite for normal rhythmic

cardiac function, and probably also to facilitate intercellular exchange of small

molecules, such as regulatory proteins and metabolites.

The cardiac hypertrophic response is a dynamic continuum in which progressive

changes in protein phosphorylation and gene expression first create adaptive structural

and functional changes but which can eventually lead to increasingly maladaptive

changes, culminating in heart failure. Electrical propagation velocity first increases in

hypertrophied ventricles but then decreases as hypertrophy becomes more severe and

chronic (Cooklin et al., 1997; McIntyre and Fry, 1997). An increase in electrical

propagation velocity in compensated hypertrophy is generally associated with increased

connexin levels, increased number of gap junctions, and enhanced intercellular

coupling (Darrow et al., 1996; Dodge et al., 1998; Kostin et al., 2004). Decrements in

conduction velocity, a general feature of chronic left ventricular hypertrophy in

humans, may be related to discontinuities in extracellular resistance caused by

interstitial fibrosis (Li et al., 1999) and an increase in intercellular resistance

attributable to gradual decrease in connexin expression or phosphorylation (Danik et

al., 2004).

Chapter I

23

In general, increased junctional resistance, due to qualitative or quantitative changes in

the expression of gap junctional connexins, may delay conduction of action potentials

and predispose the heart to re-entrant arrhythmias. It has been demonstrated that

ventricular Cx43 mRNA and protein expression is reduced in patients with chronic

ischemic and non-ischemic heart diseases and decompensated cardiac hypertrophy

(Peters et al., 1993; Dupont et al., 2001; Kitamura et al., 2002; Kostin et al., 2004).

Mice with conditional inactivation of the Cx43 gene exclusively in cardiomyocytes

show profound conduction defects and develop ventricular arrhythmias and sudden

death by 2 months of age (Gutstein et al., 2001), while knock-out mice for Cx40 show

prolongation of the PR interval on the electrocardiogram, evidence of bundle-branch

block and slow conduction in the His-Purkinje conducting system (Tamaddon et al.,

2000).

At the protein level, connexin phosphorylation, modulating cardiomyocyte gap junction

permeability, is regulated by protein kinase C (PKC)-dependent intracellular signaling

(Doble et al., 2000; Bowling et al., 2001). For example, administration of the pro-

hypertrophic octapeptide Ang II to the extracellular fluid rapidly reduces gap junction

conductance in hearts of cardiomyopathic hamsters (De Mello, 1996), and also in

normal rat hearts (De Mello and Altieri, 1992). This effect is blocked by the AT1

receptor blocker losartan and also by staurosporin, a PKC inhibitor (De Mello, 1996).

5.6. The sodium/hydrogen exchanger (NHE) and pH regulation in

cardiac hypertrophy

The sodium/hydrogen exchanger (NHE), and more specifically the NHE-1 isoform, is a

ubiquitous protein in mammalian cells. It is an integral membrane protein which

exchanges one intracellular proton (H+) for an extracellular sodium (Na+) ion, thereby

protecting cells from intracellular acidification. Indeed, the major function of NHE-1 is

the regulation of intracellular pH, but it also participates in regulation of sodium fluxes

and cell volume (Orlowski and Grinstein, 2003). NHE-1 is activated by decreases in

intracellular pH, but also by numerous hormones which can activate protein kinases,

Chapter I

24

such as mitogen activated protein kinases (MAPKs) and p90rsk (Dostal and Baker,

1998; Moor and Fliegel, 1999; Takahashi et al., 1999). The activation of NHE-1 is

associated with a variety of downstream events, including cell proliferation and

differentiation (Kapus et al., 1994; Wang et al., 1997), apoptosis in mesenchymal cells

and survival in epithelial cells (Wu et al., 2003; Sun et al., 2004), cytoskeletal

organization and cell migration (Denker et al., 2000; Denker and Barber, 2002).

NHE-1 is the predominant isoform expressed in cardiac tissue and it is one of the most

important mechanisms by which protons are extruded from cardiomyocytes.

Importantly, there is extensive evidence that NHE-1 modulates cardiomyocyte

contractility (discussed in Chapter IV) and a number of lines of evidence suggests that

NHE-1 may represent a key factor mediating hypertrophic responses in the heart. The

NHE has been shown to be upregulated in pressure-overload hypertrophy in rabbits and

in stretch-activated cardiomyocytes in culture (Takewaki et al., 1995), as well as in

stretched papillary muscles from adult cats (Cingolani et al., 1998). Furthermore, the

NHE gene has been shown to be upregulated in isolated hearts ex vivo in response to

ischemia or injury (Gan et al., 1999; Karmazyn et al., 1999), while development and

ageing seems to be associated with a downregulation of the gene (Chen et al., 1995;

Haworth et al., 1997).

In vitro studies have demonstrated that NHE-1 inhibitors (e.g. amiloride and cariporide)

block hypertrophic responses to various stimuli. For example, stretch-induced

stimulation of protein synthesis in neonatal cardiac myocytes, as well as stretch-

induced alkalinization in feline papillary muscles, can be blocked by NHE inhibitors

(Cingolani et al., 1998; Yamazaki et al., 1998), as can noradrenalin-induced protein

synthesis in cultured neonatal rat cardiomyocytes (Hori et al., 1990). In vivo

experiments demonstrated that NHE-1 inhibition attenuates and reverses post-infarction

remodeling and heart failure in Sprague-Dawley rats after coronary artery ligation

(Yoshida and Karmazyn, 2000; Chen et al., 2004), whereas cariporide treatment

completely prevented the development of fibrosis, left ventricular remodeling and

contractile dysfunction in beta 1-adrenergic receptor transgenic mice (Engelhardt et al.,

2002). Similar treatment prevented hypertrophy and cellular electrical and ionic

remodeling in pressure- and volume overload-induced cardiac hypertrophy in rabbits

(Baartscheer et al., 2005).

Chapter I

25

The precise mechanism for NHE-1 involvement in the hypertrophic response and in

heart failure remains to be determined, however it is suggested that intracellular sodium

overload could be the link between abnormal NHE-1 activity and hypertrophic growth

in cardiomyocytes (Gu et al., 1998; Hayasaki-Kajiwara et al., 1999; Chahine et al.,

2005). In particular, activation of the NHE would increase intracellular levels of

sodium and activate PKC-mediated hypertrophic stimuli. Alternatively, NHE-1

involvement in cardiomyocyte hypertrophy could be mediated by the activation of

various kinases (e.g. MAPKs) resulting in cell growth (Yamazaki et al., 1998).

Investigations into the role of the NHE-1 in cardiac disease have also been examined in

the particular setting of ischemia and reperfusion. Although clinical trials have not yet

provided unequivocal evidence, it is experimentally established that NHE-1 inhibitors

applied during ischemia and/or reperfusion protect the heart against ischemic damage

ex vivo and in vivo (Myers et al., 1998; Xiao and Allen, 2000; Yoshida and Karmazyn,

2000; Chen et al., 2004).

Chapter I

26

6. MOLECULAR PATHWAYS FOR CARDIAC

DEVELOPMENT, HYPERTROPHY AND FAILURE

6.1. Embryonic heart development and cell fate determination

The processes of cell growth, proliferation and death in the heart involve qualitative

and quantitative changes in gene expression (Anversa et al., 1998). Cardiomyocyte

hypertrophy is characterized by cell structural rearrangement and growth. This is

caused by an increase in protein biosynthesis induced by gene reprogramming.

Changes in gene expression are linked with reactivation of immediate early genes and

genes that lead to expression of fetal and embryonic isoforms of structural and

secretory proteins (Fambrough et al., 1999), such as β-myosin heavy chain (β-MHC),

skeletal α-actin and ANP (Lompre et al., 1984; van Bilsen and Chien, 1993). For

instance, the shift in actin and MHC isoforms has important implications for

cardiomyocyte contractility, since fetal β-MHC and skeletal α-actin are associated with

slower shortening kinetics and lower energy-cost contraction (Barany, 1967; Alpert and

Mulieri, 1982). Since cardiac hypertrophy is a form of heart ‘re-modeling’ led by the

reactivation of embryonic gene expression patterns, unraveling the mechanisms

underlying the early modeling and development of the embryonic cardiovascular

system is of particular interest.

Embryonic heart development commences long before the appearance of electrically

functional, beating heart tissue. The heart initially forms as a simple two-layered tube

(the outer myocardial layer and the endocardium). Later the epicardium covers the

myocardium to form the outer surface of the heart. The endocardium will form the

endothelial lining of the heart, while the myocardium will constitute the cardiac muscle

mass and the epicardium will be the source of cardiac fibroblasts and coronary arteries

(for reviews, see Lough and Sugi, 2000; Männer et al., 2001). Endocardium,

epicardium and myocardium share a common lineage as they are all derived from the

Chapter I

27

same group of mesodermal cells. After the heart tube has formed, it goes through a

complex series of looping movements and tissue elaborations that result in septation

and valve formation and generate the mature, multichambered organ (Olson and

Srivastava, 1996; Moorman and Christoffels, 2003).

The earliest event in heart formation is commitment of mesodermal cells to a

‘cardiogenic fate’ and their migration into antero-lateral regions of the embryo during

late gastrulation. The heart and the derivatives of the blood islands are the first

mesodermal tissues to differentiate after gastrulation in embryos. Cells that migrate

anterior and lateral to the primitive streak in early gastrulation contribute to heart

tissue, whereas cells that move into the posterior lateral plate form the blood islands

and blood precursors (Garcia-Martinez and Schoenwolf, 1993). In this process,

morphogenic movements and cardiac fate determination are believed to be regulated by

multiple growth and morphogenic factors secreted from the underlying endodermal

cells, including bone morphogenetic proteins (BMPs), transforming growth factor-beta

(TGF-β) and fibroblast growth factors (FGFs), (Beddington and Robertson, 1999;

Azhar et al., 2003; Sugi and Markwald, 2003). Many, if not all of these growth and

morphogenic factors are differentially regulated in the adult heart during cardiac

hypertrophy and failure.

6.1.1 BMP and Wnt signaling during embryonic development

BMP signals from the lateral regions of the embryo are required for heart formation. In

particular BMP2 and BMP4 signaling in mammalian neural crest derivatives is

essential for outflow tract development and may regulate a crucial

proliferation/differentiation signal for the ventricular myocardium (Schlange et al.,

2000; Stottmann et al., 2004). Marvin et al. (2001) proposed a model for the chick

embryo in which a dorsal-ventral BMP gradient intersecting an anterior-posterior

wingless (Wnt) signal gradient would induce cardiogenesis in a region of high BMP

and low Wnt activity. Wnt genes, related to wingless in Drosophila melanogaster,

encode a number of secreted proteins that play critical roles in the development of

many organisms, especially in cell fate and patterning (Figure I-2). Wingless itself

collaborates with decapentaplegic, a BMP homologue and member of the TGF-β

Chapter I

28

superfamily, to specify the rudimentary heart tube in flies (Frasch, 1995). In the

absence of Wnt proteins, cells undertake active measures to maintain low levels of the

Wnt signaling protein β-catenin. This is realized by a glycogen synthase kinase 3β

(GSK-3β)-dependent ubiquitination of β-catenin (Liu et al., 2002). Wnt binding to the

frizzled and LRP5/6 receptors activates some associated downstream components,

leading to inactivation of GSK-3β, thereby stabilizing and inducing an accumulation of

β-catenin. Accumulation and translocation of β-catenin to the nucleus enhances

transcription of Wnt-responsive genes.

6.1.2 Nkx and GATA signaling

Homeodomain transcription factors comprise a large family of DNA binding factors

that regulate transcription and development. Many homeodomain genes arranged in

genomic clusters determine anterior-posterior patterning, while others determine the

fate of cells in specific tissues. The proliferation of cardiac myocytes and their

differentiation early in development are dependent on the coordinate expression and

action of serum response factor (SRF), GATA4 and the homeodomain factor Nkx2-5.

All three of these factors are expressed in developing cardiomyocytes and induce

expression of cardiac genes. In addition, the Hop (Homeodomain Only Protein) gene

encodes a factor expressed early in cardiac development that is involved in cardiac

differentiation. The influence of Hop on the opposing processes of cardiomyocyte

differentiation and proliferation reflect the interaction of Hop with SRF and the duel

role SRF plays. In early cardiac development, Hop opposes differentiation induction by

SRF, while at later stages Hop opposes the proliferation induced by SRF. Interestingly,

transgenic mice that overexpress Hop develop severe cardiac hypertrophy, cardiac

fibrosis, and premature death (Kook et al., 2003).

The vertebrate homeobox gene Nkx2-5 (homologous to tinman in D. melanogaster and

first identified in the mouse) is essential for early development of precardiac tissues,

suggesting that expression may be an early indication of commitment to the cardiac

lineage. Expression of Nkx2-5 within the vertebrate heart is limited to the muscular

tissues and the same expression continues in the adult heart in fully differentiated

Chapter I

29

cardiac muscle suggesting that this gene is important for maintaining heart gene

expression, not just for early heart development (Lints et al., 1993). BMP signaling is

crucial in the regulation of Nkx2.5 expression and specification of the embryonic

cardiac lineage (Brown et al., 2004).

GATA proteins are zing-finger transcription factors that bind to a DNA element that

includes the sequence GATA- hence the name. Like the tinman-related homeodomain

proteins, GATA proteins are capable of interacting with other protein in order to form

active transcription complexes. Three members of the GATA family, GATA4, GATA5

and GATA6, are expressed during heart development in both myocardial and

endocardial tissue. Transcription of all three genes commences very early in precardiac

tissue, soon after the tissue is specified, and continues in the myocardial and

endocardial layers of the heart tube and adult heart (Molkentin et al., 1997). GATA

binding sites are present in the regulatory regions of a number of cardiac specific genes,

including α-myosin heavy chain (α-MHC), myosin light chain 2 (MLC2), myocyte-

specific enhancer factor 2 (MEF2), cardiac troponin C, cardiac actin and atrial

natriuretic peptide. These are markers of cardiomyocyte differentiation and are

generally re-expressed in their fetal isoforms during adult cardiac hypertrophy.

6.1.3 MEF2 and cardiomyocyte differentiation

Members of the myocyte-specific enhancer factor 2 (MEF2) family are important for

the formation of muscle tissue in vertebrates. The mouse genome contains four genes,

MEF2A, MEF2B, MEF2C, and MEF2D, which are related to the single D.

melanogaster D-MEF2 gene. While the different vertebrate MEF2 genes are expressed

in a wide variety of muscle cell types, all are expressed in cardiac muscle at some time

during embryonic development. The MEF2 binding sequence is present in the

promoters of a number of cardiac specific sequences, including, α-MHC and MLC2. In

mice lacking MEF2C function, several cardiac differentiation markers including ANP,

cardiac actin and α-MHC are not expressed. Although a normal linear heart tube forms,

it fails to undergo looping. This may be due to the altered expression of other

transcription factors, like Hand1 and Hand2, which are involved in establishing

Chapter I

30

regional identity within the heart. Interestingly, MEF2 binding activity is necessary for

regulation of the GLUT4 glucose transporter gene promoter in muscle and adipose

tissue, thus suggesting that normal expression of MEF2 transcription factors is essential

for the glucose metabolism of the embryonic and adult heart (Thai et al., 1998; Michael

et al., 2001).

6.2. Cardiomyocyte growth and hypertrophy in the adult heart

Because of their contractile activity and numeric contribution to the final heart mass,

cardiomyocytes are fundamentally involved in all forms of cardiac remodeling. Growth

of the heart during embryogenesis occurs primarily through proliferation (hyperplasia)

of cardiac myocytes. Soon after birth, cardiomyocytes withdraw largely from the cell

cycle and subsequent growth occurs predominantly through differentiation and increase

in myocyte size (hypertrophy) rather than number. It has been generally accepted that

adult cardiomyocytes are terminally differentiated cells, which have lost their capacity

to divide (Liu and Olson, 2002). However new evidence suggests that myocyte

differentiation and tissue renewal from a pool of undifferentiated stem cells contribute

in a limited but significant way to the remodeling processes in the adult heart (Orlic et

al., 2001; Anversa et al., 2002; Penn et al., 2004; Rosenblatt-Velin et al., 2005).

It is important to highlight the fact that in terms of gene expression, there are

differences between a developing embryo and an adult patient undergoing cardiac

hypertrophy. In the first case, the embryonic expression of gene sequences are part of a

program highly regulated in space and time which deletes vestigial structures, controls

cell numbers and physiologically models structures (van den Hoff et al., 2000). In adult

life, re-expression of fetal gene programs in the heart mainly serves to maintain the

heart in a new compensatory and homeostatic state by counterbalancing deleterious

environmental changes. Therefore, in the second case, the endogenous program

interacts with new variables like ageing, mechanical stress, neuro-humoral secretion

and metabolic changes which are absent during fetal or embryonic life. In general

terms, adult cardiomyocyte hypertrophy results in intracellular production of extra

contractile proteins and sarcomeric reorganization. This sarcomeric and morphological

Chapter I

31

reorganization is associated with alterations in cardiomyocyte contractility, which

eventually will lead to impaired myocardial function. The adult hypertrophic response

is frequently characterized as a ‘fetal recapitulation’ growth pattern. As the findings to

be presented in the Thesis demonstrate, such a concept is erroneously over-simplified.

6.2.1. Different signaling pathways for different

hypertrophic stimuli

One general aim to understand the molecular mechanisms of adult cardiomyocyte

hypertrophy and remodeling is to attribute distinct phenotypic features to defined

stimuli and signaling pathways. Multiple interacting signal transduction pathways are

initiated by mediators of hypertrophic phenotypes in vitro and in vivo. Such mediators

include mechanical stretch, cytokines, growth factors, catecholamines, vasoactive

peptides and hormones. In general terms, the transmission of these signals involves

three common phases: 1) activation of receptors or adhesion molecules that are mainly

located at the cell surface, 2) transduction of the signal through activation of various

second messengers, and 3) initiation of cytosolic and nuclear events that lead to

differential gene expression and modification of cell physiology.

6.2.2. Signaling pathways through seven-transmembrane

receptors

An enormous family of over 1000 genes encodes receptor proteins that are

characterized by a signature seven-transmembrane (7TM) configuration. Members of

this family comprise receptors for many hormones, peptide and non-peptide, including

the receptors for Ang II (i.e. AT1 and AT2 receptor subtypes), endothelin-1 (ETA and

ETB receptor subtypes) and noradrenaline (α- and β-AR adrenoreceptors), which play

an important role in regulating cell physiology in the heart. The 7TM receptors are

commonly referred to as G-protein-coupled receptors (GPCRs), because most of them

transduce their signals by activating heterotrimeric G-proteins. Signals by GPCR

ligands result in transcriptional translation of immediate early genes, such as Fos and

Chapter I

32

Jun, and other genes involved in long-term remodeling of heart tissue and the

physiological response to stress in the heart, such as fetal isoforms of actin and myosin

filaments, ANP and BNP.

In the traditional view of 7TM-receptor signaling, a single cell-surface receptor protein

is activated by the binding of a single agonist ligand, and this ligand-activated receptor

can then activate many G-proteins, one at a time (Figure I-3). However, an abundance

of new information has been generated in recent years relating to the signaling events

triggered through GPCRs. It is becoming clear that receptors are not isolated entities,

but that they interact with other receptors already present or recruited to the vicinity,

which results in new signaling models where cell-surface receptors function as dimers

or even higher multimers. Several examples of functional GPCR heterodimers have

been identified. For example, the Ang II-activated AT2 receptor binds the AT1 receptor

and antagonizes its function (AbdAlla et al., 2001); while the AT1 receptor binds to the

bradykinin receptor B2 causing increased activation of Gαq and GαI proteins and

modulation of the endocytotic pathway for both receptors (AbdAlla et al., 2000).

Moreover, heterodimerization of β1- and β2-AR inhibits the agonist-promoted

internalization of the β2-AR and its ability to activate the ERK1/2 (see below)

signaling pathway (Lavoie et al., 2002).

There is increasing evidence to suggest that mitogenic and growth responses to receptor

activation by peptides such as Ang II and endothelin-1 may be mediated also by

GPCR-mediated transactivation of tyrosine-kinase receptors (RTKs), such as the

receptors mediating epidermal growth factor (EGF), platelet-derived growth factor

(PDGF) or insulin-like growth factor (IGF) activity (Thomas et al., 2002) (Figure I-4).

For example, AT1 receptor-elicited tyrosine phosphorylation and activation of

epidermal growth factor receptor (EGFR) stimulates downstream activation of ERK1/2

and vascular smooth muscle cell hyperplasia (Shah et al., 2004). In rat smooth muscle

cells, both Ang II-induced nuclear proto-oncogene expression and increase in c-Fos

protein are prevented by treatment with EGFR kinase inhibitor (Thomas et al., 2002).

Ang II-mediated EGFR transactivation also plays a role in p70 ribosomal protein S6-

kinase-induced protein synthesis leading to hypertrophy (Eguchi et al., 1999).

Furthermore, recent studies have demonstrated that EGFR activation is involved in Ang

Chapter I

33

II-induced vascular contraction and it has been demonstrated that Ang II can directly

promote cardiomyocyte growth via AT1 receptor-mediated transactivation of EGFR in

vitro (Thomas et al., 2002).

6.2.3. Signaling through mitogen-activated protein

kinases (MAPKs)

Mitogen-activated protein kinases are a family of serine/threonine protein kinases that

mediate nuclear transduction of extracellular signals by intracellular protein

phosphorylation, leading to a cascade of transcription factor activation, altered gene

expression and trophic cellular responses. Mammalian MAPKs are grouped into six

major subfamilies: a) ERK1/2 (also known as p42-kDa MAPK and p44-kDa MAPK,

respectively), b) c-Jun N-terminal/stress-activated protein kinases (JNK/SAPK), c) p38

MAPK, d) ERK6, p38-like MAPK, e) ERK3, and f) ERK5 (also called Big MAPK 1)

(Robinson and Cobb, 1997) (Figure I-5). Signals from cell surface receptors such as

GPCRs and growth factor RTKs are transduced, directly or via small G proteins such as

Ras and Rac, to multiple tiers of protein kinases that amplify these signals and/or

regulate reciprocally. MAPK-dependent signaling pathways have been associated with

cellular growth and apoptosis, cellular differentiation and transformation and vascular

contraction. ERK1/2 is activated in response to growth and differentiation factors,

whereas JNKs and p38 MAPK are usually activated in response to inflammatory

cytokines and cellular stress (Pellieux et al., 2000).

6.2.4. Calcium-dependent pathways of cardiac hypertrophy

induction

It has been shown that cardiac hypertrophy can be induced by the calcium-dependent

phosphatase calcineurin after Ang II and adrenalin stimulation (Molkentin et al., 1998).

Calcineurin dephosphorylates NF-AT3 transcription factors. NF-AT3s translocate to

the nucleus and activate cardiac zinc-finger transcription factor GATA4, resulting in

synergistic activation of cardiac transcription (Figure I-6). Interestingly, GATA4 has

Chapter I

34

also been shown to be required for transcriptional activation of genes for AT1A

receptors and different isoforms of MHC during hypertrophy (reviewed in Molkentin

and Olson, 1997). Usage of specific anti-calcineurin oligonucleotides suppressed the

decrease in expression of adult genes (e.g. α-MHC, GLUT4 and SERCA2) during

pressure-overload cardiac hypertrophy in rats (Depre et al., 1999a). Because the

calcineurin/NF-AT3 signal transduction pathway is only one of several signaling

systems shown to be capable of inducing hypertrophy, an important question is whether

this pathway is integrated with, or independent of, other hypertrophic signaling

systems. For example, certain transgenic mouse lines exhibiting cardiac hypertrophy do

not respond to calcineurin inhibition (Sussman et al., 1998b) and hypertrophy in the

SHR is not prevented by inhibition of the calcineurin pathway (Zhang et al., 1999),

indicating the existence of calcineurin-independent mechanisms of cardiac growth.

Recently, it has been shown in transgenic mice that the intracellular calcium-binding

protein calmodulin (CaMK) potently stimulates transcription factor myocyte-specific

enhancer factor (MEF2) activity through a posttranslational mechanism in the heart in

vivo (Passier et al., 2000). This activity is sufficient to induce cardiac hypertrophy in

these mice.

6.3. Cardiomyocyte death

Myocyte demise can occur by three mechanisms in the heart: apoptosis, autophagic cell

death and necrosis (Fesik, 2000; Klionsky and Emr, 2000; Knaapen et al., 2001). These

forms of cell death have a different impact on the morphology of the myocardium

(Anversa et al., 1998). Apoptosis is a form of programmed cell death which ‘shapes’

the organs during embryonic development and which also allows tissue and cell

renewal in the adult. Apoptosis does not generally lead to the activation of

inflammatory reactions, vascular proliferation or collagen deposition. It is characterized

by cell shrinkage, chromatin condensation, DNA fragmentation, membrane blebbing

and formation of apoptotic bodies (Majno and Joris, 1995). The physiological process

of apoptosis can become maladaptive as it is observed in the heart after myocardial

infarction or during the development of heart failure (Olivetti et al., 1997). Indeed,

myocyte apoptosis constitutes the earlier and primary insult of myocardial infarction,

Chapter I

35

which involves nearly 85% of cells in the ischemic region (Bardales et al., 1996).

Furthermore, myocyte apoptosis without replacement is an important mechanisms

involved in the transition from cardiac hypertrophy to heart failure occurring after

ischemia, reperfusion and myocardial infarction in humans (Olivetti et al., 1997). In

addition, apoptosis has been detected in cardiomyocytes experiencing

hypoxia/reoxygenation (Tanaka et al., 1994) and mechanical stretch (Cheng et al.,

1995) and in animal models of cardiac ischemia and reperfusion injury (Gottlieb et al.,

1994; Fliss and Gattinger, 1996). Various apoptotic pathways are active in mammalian

cells, including cardiomyocytes. One of the best characterized pathways is the death

receptor signaling pathway (Figure I-7). The death receptors are a class of cell

membrane receptors belonging to the TNF receptor gene superfamily, such as Fas or

TNFR1 (Ashkenazi and Dixit, 1998). Upon binding by their cognate ligand, they

initiate mitochondrial-mediated cellular apoptosis. The receptor-ligand mediated

apoptotic pathway is not the only mechanism in the apoptotic process: myocyte

apoptosis can also be induced by oxidative stress (Kumar and Jugdutt, 2003), by

GPCR-mediated intracellular signaling cascade (Wencker et al., 2003), and possibly by

calcium and calcineurin-dependent signaling activation (Jayaraman and Marks, 2000;

Braz et al., 2003).

Several proteins constitute a critical checkpoint in cell death. These proteins contain

agonists and antagonists of apoptosis, and alterations in their ratio determine the ‘life or

death’ outcome for a cell. Cytochrome c is localized on the outside of the inner

mitochondrial membrane and in the intermembrane space. It has an important function

in the intracellular electron transport chain reaction for the production of ATP. During

the apoptotic process cytochrome c is released in the cytosol. There it binds to the

apoptosis protease activating factor (Apaf-1) which is present in mammalian cells and

is analogous to Ced-4 in C. elegans. Cytochrome c and Apaf-1 form a complex

together with dATP, which subsequently activates procaspase-9 to caspase-9. This

results finally in the activation of procaspase-3 into caspase-3 which leads to the

characteristic morphological consequences (Figure I-7). Other examples are Bcl-2, a

protein that promotes cell survival, and Bax, which facilitates the activation of an

endogenous cell death pathway (Oltvai et al., 1993). Bax forms heterodimers with Bcl-

2, depressing its protective influence on cell viability; if Bax homodimers dominate,

cells are more predisposed to undergo apoptosis. The tumor suppressor gene p53 is a

Chapter I

36

transcriptional regulator of the Bcl-2 and Bax genes and the induction of p53

downregulates Bcl-2 and upregulates Bax in the cells. However, it has been shown that

the attenuation of Bcl-2 and the enhanced expression of Bax alone cannot trigger

apoptosis (Miyashita and Reed, 1995) and that some other factors or death stimuli are

necessary. Over-expression of p53 in adult ventricular myocytes in vitro upregulates

the transcription of Agt and the AT1 receptor subtype leading to generation of Ang II

and apoptosis (Pierzchalski et al., 1997). This suggests that Ang II may be one of the

pro-death stimuli involved in myocyte programmed cell death.

As with apoptosis, autophagic cell death is regulated and is associated with DNA

fragmentation, but unlike apoptosis it is caspase-independent and morphologically

resembles necrosis (Klionsky and Emr, 2000). Finally, necrosis is a rapid and

irreversible process that occurs when cells are severely damaged. According to Majno

and Joris (1995) necrosis is not a form of cell death but the end stage of any cell death

process, including apoptosis. Necrosis involves swelling of the cell and its organelles,

chromatin condensation, disruption of mitochondria, membrane rupture, cell lysis and

inflammatory immune response (Van Cruchten and Van Den Broeck, 2002). In the

heart, necrosis ultimately results in the generation of tissue fibrosis, diffuse or focal in

nature.

6.4. Cardiomyocyte growth and death: a fine balancing act

The interactions between various pro-growth and pro-death signaling pathways are

complex. Transition to heart failure is usually preceded by cardiac and cardiomyocyte

hypertrophy and, therefore, it has been suggested that hypertrophy could render

cardiomyocytes more sensitive to apoptosis. This notion is supported by various

studies, including work on transgenic mice overexpressing the G-protein Gαq, which

couples to receptors transducing various pro-hypertrophic and neuro-humoral signals.

In these mice, cardiac hypertrophy is associated with increased cardiomyocyte

apoptotic tendency and increases the likelihood of heart failure under stress conditions

(Yussman et al., 2002). In contrast, some hypertrophic signaling molecules appear to be

clearly protective against apoptosis, promoting survival during stress conditions. This is

Chapter I

37

the case for the cytokine receptor gp130, a receptor subunit shared in common with

other cytokines in the IL-6 family, acting via cardiotrophin1-dependent inhibition of

apoptosis (Yasukawa et al., 2001). The gp130 subunit also influences insulin-like

growth factor 1 (IGF-1) and phosphatidylinositol 3-kinase-dependent pathways (Lee et

al., 1999; Aikawa et al., 2000). A further complication is that some hypertrophic

signaling molecules have both anti- and pro-apoptotic properties. As an example, the

JNK can promote apoptosis via Fas-ligand (FasL) or Bcl2-mediated pathways

(Tournier et al., 2000). In contrast, other studies have demonstrated anti-apoptotic

effects of JNK activation in neonatal cardiomyocytes and differentiated mouse

embryonic stem cells (Minamino et al., 1999; Andreka et al., 2001). The anti-apoptotic

effects could be dependent on the activity of JNK as a transcriptional regulator of gene

expression (Minamino et al., 1999; Tournier et al., 2000). Another example of a

hypertrophic signaling molecule with dual effects on apoptosis is the

calcium/calmodulin-regulated phosphatase calcineurin. Activation and overexpression

of calcineurin confers a protective effect against apoptosis in vivo and in vitro (De

Windt et al., 2000). However, calcineurin-dependent β-adrenergic stimulation in

cardiomyocytes can promote the mitochondrial release of cytochrome c and apoptosis

(Saito et al., 2000).

At present, whether hypertrophic signaling pathways are pro-apoptotic or anti-apoptotic

remains an open question. The seemingly contradictory findings for certain signaling

molecules and pathways may be due to the complex interaction between various stimuli

and their duration or intensity on different cellular fractions of the myocardium

(cardiomyocytes, fibroblasts, etc.). Indeed, the mechanical and paracrine interaction

between cardiomyocyte and non-myocyte cells, as well as the dynamic interface

between the cellular and interstitial matrix of the myocardium adds further complexity

to shaping the cardiac response to hypertrophic stimuli.

Chapter I

38

6.5. The fibroblast fraction: collagen production and

cardiomyocyte hypertrophy

Fibroblasts represent the majority of the non-myocyte cellular component of the heart.

During cardiac hypertrophy, fibroblasts can undergo different morphological and

physiological changes: they can express smooth muscle α-actin and embryonic MHC

and develop a contractile phenotype (Frangogiannis et al., 2000), and they can

proliferate and produce ECM components, including collagen I and III, fibronectin and

laminin (Weber and Brilla, 1991; Pauschinger et al., 1999). As a consequence, fibrous

material deposition alters the organized matrix and the interconnections between

cardiomyocytes and adjacent vascular structures. The degree of fibrosis affects the

oxygenation and the metabolism of the cells in the myocardium. Fibrosis contributes to

myocardial stiffness, facilitates arrhythmogenic events, impairs myocyte contractility

and hampers systolic ejection (Swynghedauw, 1999).

Cardiac fibroblasts are also thought to actively contribute to hormone and growth

factor-induced cardiomyocyte hypertrophy. Putative pathways for humoral and

mechanical regulation of protein synthesis and hypertrophy involve cross-talk between

cardiomyocyte and fibroblast populations within the myocardium. For instance, when

the pro-hypertrophic peptide Ang II is added to purified fibroblasts and the supernatant

from the fibroblasts is added to myocytes in culture a significant increase in stimulation

of protein synthesis (and therefore cardiomyocyte hypertrophy) is obtained (Harada et

al., 1997). There is indeed growing evidence that fibroblasts (and other non-myocyte

cells) may secrete growth factors and peptides, such as endothelin-1 that stimulate

myocyte hypertrophy. According to this hypothesis, Harada et al. (1997) demonstrated

that only co-cultures of cardiomyocytes and fibroblasts respond to Ang II and TGF-β1

with growth signals. In light of these findings it can also be postulated that the degree

of cardiac and cardiomyocyte hypertrophy observed could depend on the qualitative

and quantitative presence of fibroblasts to promote paracrine interaction between

cardiomyocytes and fibroblasts in the myocardium.

Chapter I

39

6.6. Extracellular matrix turnover and fibrosis

The turnover of the extracellular matrix in the hypertrophic myocardium is regulated by

non-myocyte and myocyte activation leading to differential gene expression and

cellular secretion of both pro- and anti-fibrotic substrates.

6.6.1 Matrix degradation

Matrix metalloproteinases (MMP) are the zinc-containing, calcium-dependent

endopeptidases (e.g. collagenases, gelatinases, stromelysin, matrylisin) that maintain

homeostasis of tissue structure, including the myocardium, by digesting the various

components of the ECM. Thus, an increase in MMP activity may result in fibrillar

collagen degradation, myocyte slippage, ECM remodeling, and progressive ventricular

dilation (Olivetti et al., 1990; Gunja-Smith et al., 1996; Spinale et al., 2000). Although

several lines of evidence suggest that MMP activation and the initial development of

myocardial hypertrophy are independent events (Chancey et al., 2002), a number of

studies have demonstrated a role for increased MMP activity and abundance in the

development of experimental congestive heart failure and end-stage cardiomyopathic

disease in humans (Berry et al., 2004; Janicki et al., 2004). These data suggest that

MMP activation in the myocardium contributes to dilation (with or without

hypertrophic growth) and a consequent increase in wall stress leading to mechanical

dysfunction. Several stimuli have been linked to MMP activation in the myocardium.

The role of inflammatory cytokines, such as interleukin-1β (IL-1β), TNFα and IL-6,

and reactive oxygen species (ROS) in the production of MMPs has been demonstrated

in various models in vitro and in vivo.

6.6.2 Matrix biosynthesis

In general terms, there are two main forms of matrix biosynthesis leading to myocardial

fibrosis: 1) reparative fibrosis, occurring after ischemia, infarction or during

Chapter I

40

senescence, resulting from a scarring process in which small and large areas of necrosis

heal after direct insults; and 2) reactive fibrosis, which is suggested to be a fibrogenic

response to intra-cardiac cell activation caused by various neuro-humoral factors,

inflammatory cytokines and stretch (reviewed in Schwinghedauw, 1999). Fibrosis is

often associated with ischemia, senescence, inflammatory diseases, diabetes and

expression of growth hormones, such as Ang II and TGF-β1 (Weber and Brilla, 1991;

Shimizu et al., 1993; Pinto et al., 2000). Myocardial fibrosis is defined not only as a

quantitative change in the concentration of matrix collagen in the interstitium, but also

as a qualitative change in collagen type, organization and cross-links. Type I and III are

the predominant types of collagen in the adult heart (Pauschinger et al., 1999). Various

changes in the composition of collagen types and cross-links have been reported in

experimental models, as well as in patients with heart failure (Bishop et al., 1990;

Chapman et al., 1990). It is believed that quantitative elevations of collagen type I, and

collagen type I cross-linking, increase myocardial stiffness, while increases in type III

collagen may facilitate myocardial compliance. During cardiac remodeling normal

collagens are progressively degraded by MMP and are replaced by fibrous interstitial

deposits of poorly cross-linked collagens (Gunja-Smith et al., 1996), which may lead to

dilation of the ventricles.

6.6.3 Matrix cell adhesion and transduction of signaling

pathways

ECM substances provide cells with a structural, chemical and mechanical substrate that

is essential for normal development and responses to pathophysiological signals.

Glycoprotein transmembrane receptors termed integrins are the primary link between

ECM ligands and cytoskeletal structures. They serve both as adhesive receptors and

mechanotransducers of signaling events (Giancotti and Ruoslahti, 1999; Hynes, 1999).

Integrins orchestrate multiple functions in the intact organism including organogenesis,

regulation of gene expression, cell proliferation, differentiation, migration and death.

Integrins themselves signal through a host of pathways, but it is rare for any signaling

pathway to function in isolation. Extensive studies have been performed to show that

integrins and growth factors form a particularly robust synergy (Miyamoto et al., 1996).

Chapter I

41

RTKs, GPCRs and cytokines have all been linked to this response. Terracio et al.

(1991) were amongst the first investigators to document the importance of integrins in

the myocardium and describe how integrin expression was changed with hypertrophy.

Sadoshima and Izumo (1993a) subsequently utilized mechanical stimulation of cultured

myocytes to simulate hemodynamic loading of the intact myocardium. They showed

how a stretch stimulus activated a variety of intracellular signaling pathways (many of

which are known to be involved in growth factor and integrin signaling cascades) and

also provoked paracrine release of factors such as Ang II (Sadoshima et al., 1993).

Ang II stimulation of cardiac fibroblasts has been shown in several studies to modulate

integrin localization and expression (Burgess et al., 1994). Importantly, osteopontin, a

cytokine which binds to several cell surface receptors, including integrins, was shown

to be produced by both cardiac fibroblasts and myocytes and to be upregulated in the

hypertrophic and failing ventricle of humans and animals (Graf et al., 1997; Hsueh et

al., 1998). Osteopontin production is upregulated by Ang II and recent work has shown

that mice deficient in osteopontin have reduced Ang II-mediated cardiac fibrosis and

hypertrophy, likely due to the decreased adhesion to ECM and reduced cell

proliferation by the osteopontin-deficient cardiac fibroblasts (Collins et al., 2004).

Chapter I

42

7. THE RENIN-ANGIOTENSIN SYSTEM AND

CARDIAC REMODELING

7.1. An overview

By the end of the 19th century, the pressor effect of renal extracts had been described

and the putative substance renin was named based on its origin in the kidney (Basso

and Terragno, 2001). In the following century, the work of two independent groups

established the concept of an endocrine renin-angiotensin system (RAS). The RAS

comprises a cascade of enzymatic reactions resulting in the formation of angiotensin II

from the substrate angiotensinogen. In this classic concept, the process is initiated when

the enzyme renin acts on liver-derived Agt, an alpha-globulin, to release the

decapeptide angiotensin I (Ang I). This decapeptide has limited intrinsic

pharmacological activity, and is cleaved by angiotensin-converting enzyme (ACE),

which is present in the endothelium of the lung vessels and in the heart, to yield the

highly active octapeptide Ang II (Iwai et al., 1995). Ang II has several actions related

to blood pressure and fluid homeostasis. It is an extremely powerful vasoconstrictor; it

enhances sympathetic tone and stimulates aldosterone secretion from the adrenal gland,

increasing salt and water reabsorption (Figure I-8). Ang II undergoes hydrolysis by

aminopeptidase to yield the heptapeptide angiotensin III (Ang III), which is also

pharmacologically active. Further cleavage yields peptides with little or no pressor

activity but which possibly exert other effects, see below).

This traditional concept of the RAS has been expanded more recently to recognize the

co-existence of local level RAS in many tissue and organs, including arteries and heart

(discussed below). Several, if not all of the components of the RAS have been

identified in cardiac and vascular tissues. While a circulating RAS has a critical role in

preserving short-term cardio-renal homeostasis through acute changes in vascular tone,

aldosterone release and renal sodium and water reabsorption, a local cardiovascular

Chapter I

43

RAS influences long term control of vascular resistance and myocardial function

through Ang II-mediated contractile, mitogenic and growth promoting effects. For

instance tissue-specific RAS plays an important role in the pathogenesis of

hypertension, since drugs which inhibit Ang II production (ACE inhibitors) or its

binding to specific receptors (AT1 blockers), beneficially reduce blood pressure in

hypertensive patients. In addition, RAS blockade improves cardiac function and

metabolism in patients suffering from diabetes and/or cardiovascular diseases beyond

blood pressure reduction (Anan et al., 2004; Egan et al., 2004; Wachtell et al., 2005).

7.2. The different components of the renin-angiotensin system

7.2.1. Renin

Renin is a protease with high substrate specificity. It is both the initiating and the rate-

limiting factor in the production of Ang II in humans. Renin is a glycoprotein with a

half-life in the circulation of about 15 minutes. It is stored in the juxtaglomerular cells

of the renal afferent arteriole in an inactive form called prorenin and secreted in

response to several different physiologic stimuli (sympathetic activity and

catecholamines, sodium depletion, hypotension, hemorrhage, constriction of renal

artery or aorta and heart failure among others) (Guyton and Hall, 1996). Renin occurs

in organs other than the kidney, e.g. in the brain, where it is implicated in the regulation

of neural activities.

The mouse has a kidney-type renin gene, Ren-1, which is located on mouse

chromosome 1 (Chirgwin et al., 1984). In some mouse strains, such as IRC and NMRI,

the male sub-maxillary gland also secretes large amounts of renin (Weaver et al.,

1991). These mice have a second renin locus, Ren-2, also on chromosome 1. As

mentioned earlier, renin, and in general terms the RAS, play an important role in the

development of hypertension and the cardiac pathologies associated with a rise in blood

pressure. Transgenic over-expression of cardiac Ang II in one-renin gene strain mice

(such as C57BL6 or BALB/c) induces cardiac hypertrophy and systemic inhibition of

Chapter I

44

renin secretion which maintains normal blood pressure, while the same over-expression

in two-renin gene strains (such as the NMRI) results in cardiac hypertrophy associated

with elevated blood pressure (Mazzolai et al., 1998). This suggests that Agt expression,

not renin, is the rate-limiting factor in the production of Ang II in the mouse.

The renin gene has been genetically linked to high blood pressure in different rat

models of hypertension, such as the salt-sensitive Dahl rat and the SHR (Rapp et al.,

1989; Pravenec et al., 1991). Moreover, Mullins et al. (1990) demonstrated that

introduction of the mouse Ren-2 gene into the rat genome induces severe and fulminant

hypertension, associated with low plasma active renin and suppressed kidney renin

(and high extra-renal transgene expression).

7.2.2. The angiotensin-converting enzyme

ACE, or kininase II, is a rather non-specific metalloenzyme (containing zinc) that

cleaves dipeptide units from peptide substrates with diverse amino acid sequences.

ACE does not degrade Ang II, but plays an important role in blood pressure regulation

and electrolyte balance by hydrolyzing Ang I to Ang II. The enzyme is also able to

inactivate bradykinin, a potent vasodilator. This enzyme has a wide tissue distribution

and plays many physiological roles. The ACE gene encodes 2 isozymes. The somatic

ACE isozyme is expressed in many tissues, including vascular endothelial cells (mainly

in lungs), renal epithelial cells, and testicular Leydig cells, whereas the testicular or

germinal ACE isozyme is expressed only in sperm (Ramaraj et al., 1998). The

importance of ACE in circulatory homeostasis is well documented. ACE is abundantly

present as a membrane-bound enzyme on the surface of vascular endothelial cells, and

also ACE circulates in plasma at variable concentrations. Plasma levels of circulating

ACE are determined by an insertion (I)/deletion (D) polymorphism situated in intron 16

of the ACE gene. It is known as the ACE/ID polymorphism. Schunkert et al. (1994)

found an association between blood pressure-independent left ventricular hypertrophy

and the D/D (homozygote for the deletion) genotype of ACE. Moreover, experimental

studies have shown that ACE gene expression is increased in myocardial tissue after

aorto-caval shunt or volume-overload hypertrophy (Iwai et al., 1995; Lear et al., 1997)

and in pressure overload left ventricular hypertrophy (Schunkert et al., 1990). Pinto et

Chapter I

45

al. (1993) have also demonstrated that the cardiac ACE activity is activated early after

the induction of heart failure, suggesting that this enzyme is pathologically involved in

the early stage of heart failure.

More recently, by EST database searching and cDNA library screening Tipnis et al.

(2000) cloned a second full-length ACE cDNA (ACE2), sharing about 40% homology

with ACE. Northern blot analysis has detected high expression of ACE2 in kidney,

testis and heart. ACE2 cleaves Ang I and Ang II into Ang [1-9] and Ang [1-7]

respectively, but is not able to cleave bradykinin (Tipnis et al., 2000; Oudit et al., 2003)

(Figure I-9). ACE2 is not inhibited by classic ACE inhibitors such as captopril.

Targeted disruption of ACE2 in mice resulted in a severe cardiac contractility defect,

increased Ang II levels and upregulation of hypoxia-induced genes in the heart

(Crackower et al., 2002). Genetic ablation of ACE on an ACE2 mutant background

completely rescues the cardiac phenotype, suggesting that ACE2 is an essential

regulator of heart function in vivo and may have antagonizing ACE effects.

7.2.3. The angiotensinogen precursor

Although some other possible significant roles are not excluded (Menard et al., 1991),

the only well proven function of Agt is to provide a reservoir for Ang I, cleaved from

its N-terminal, usually by the enzyme renin. As previously mentioned, the renin-

angiotensinogen reaction is known to be the rate limiting step in the generation of Ang

II, with the exception of the mouse species, where Agt synthesis seems to be rate-

limiting. There are two variants of the Agt protein, that is a low (~50 KDa) and a high

(350-550 KDa) molecular weight (MW) variant (Campbell et al., 1985). The high MW

variant is incompletely characterized physiologically and represents less than 5% of

circulating Agt in men and menstruating women. High MW Agt markedly responds to

estrogen levels, rising to 16% in pregnancy (Tewksbury and Dart, 1982). It is also

remarkable that no function is known for the 442 residue protein produced by the

release of the decapeptide Ang I from the Agt precursor. The concentrations of the

residue protein in the plasma are nearly undetectable (Menard et al., 1983). The major

source of circulating Agt is the liver (Campbell, 1987), but Agt mRNA has been

Chapter I

46

documented in the renal cortex and medulla, adrenal gland, heart, aorta, placenta,

ovary, adipose tissue, saphenous vein and brain (Stronetta et al., 1988; Paul et al., 1993;

Yang et al., 1994). Among the stimuli that increase plasma angiotensinogen and hepatic

mRNA levels, are sex steroids (estrogens) (Krattenmacher et al., 1994), thyroid

hormones (Hong-Brown and Deschepper, 1992), Ang II (Klett et al., 1996) and

glucocorticoids (Ron et al., 1990). Mutated forms of the Agt gene have also been

genetically linked to the pathogenesis of hypertension (Jeunemaitre et al., 1992) and

coronary heart diseases (Katsuya et al., 1995) in humans.

7.2.4. Angiotensin peptides

All Ang peptides are derived from the precursor Agt. When given intravenously, Ang I

is so rapidly converted into Ang II that the pharmacological responses are

indistinguishable (provided that converting enzymes have not been inhibited). Ang II

has a short half-life (a minute or so) and it is immediately degraded by several

peptidases (Figure I-9). Of its metabolites, the heptapeptidase Ang III and the

hexapeptidase Ang IV are believed to retain significant activity. In fact, Ang II and

Ang III transduce their intracellular signals through the same receptors, possibly at

different sites and have similar affinity for at least two receptors (AT1B and AT2

receptors in rodents) (Sandberg et al., 1992; Kambayashi et al., 1993). Ang III

reproduces all the effects of Ang II but with lower potency (Blair-West et al., 1971;

Kono et al., 1975). The hexapeptide Ang IV binds to specific sites (presumptive AT4

receptors) in the brain and in peripheral tissues, including the blood vessels, endothelial

cells, heart, adrenals and kidneys. Among the cardiovascular actions of Ang IV, there is

the vasodilatory effect on lung vessels (Patel et al., 1998) and stimulation of fibroblast

proliferation in the heart (Wang et al., 1995). The pressor activity of Ang III and Ang

IV in humans is 20% (Blair-West et al., 1971; Kono et al., 1975) and 0.2% (Kono et al.,

1982) respectively, of the pressor effect of Ang II.

The heptapeptide Ang [1-7] may also play a significant role as an active molecule in

the vasculature and in the heart. According to Loot et al. (2002), chronic infusion (8

weeks) of Ang [1-7] improved endothelial aortic function and coronary perfusion and

preserved cardiac function in an experimental rat model of heart failure induced by

Chapter I

47

ligation of the left coronary artery. Furthermore, Ang [1-7] produced a significant

increase in cardiac output and stroke volume in anesthetized Wistar rats (Sampaio et

al., 2003). Finally Ang [1-7] has been shown to potentiate the vasodilator and

hypotensive effects of bradykinin in normotensive and hypertensive rats (Paula 1995;

Fernandes et al., 2001) and in porcine coronary vessels (Tom et al., 2001).

7.2.5. Angiotensin II

Numerous and increasingly complex cellular responses to Ang II are recognized. The

‘classical’ actions of Ang II (e.g. regulation of blood pressure, plasma volume and

sympathetic nervous activity) are mediated by the endocrine effects of circulating

Ang II. The autocrine and paracrine effects of locally formed Ang II are believed to

play a key role in the development of cardiovascular disease, including cardiac

hypertrophy, myocardial infarction, hypertension and atherosclerosis. As previously

mentioned, Ang II promotes its effects by acting directly through specific receptors,

indirectly through the release of other factors, and via cross-talk with intracellular

signaling pathways of other vasoactive agents, growth factors and cytokines. Ang II

receptors are coupled to multiple, specific signaling cascades, leading to diverse

biological actions. In mammalian cells, Ang II binds to two distinct high-affinity

plasma membrane receptors, AT1 and AT2 (Feng and Douglas, 2001). Two other Ang

II receptors have been described, AT3 and AT4 (Swanson et al., 1992), however the

pharmacology of these receptors has not been fully characterized.

7.2.6. The AT1 receptor

The AT1 receptor belongs to the 7TM class of GPCR. Two subtypes of the AT1

receptor have been identified, namely AT1a and AT1b, sharing 96% of protein sequence

similarity (Ye and Healy, 1992; Burson et al., 1994; Clauser et al., 1996). In humans

and rodent AT1a is the principal receptor in the vasculature, heart, brain, kidney, lung,

liver, adrenal gland and fetal pituitary, while AT1b predominates in the adult pituitary

and is only expressed in specific regions of the adrenal gland (zona glomerulosa) and

kidney (glomeruli). AT1b expression is detectable in the heart at low levels in rodents

Chapter I

48

(<10% AT1a) and may constitute a pre-junctional variant of the receptor (Guimaraes

and Pinheiro, 2005). A mutation in the SHR AT1b receptor has been linked with cardiac

hypertrophy, and may have a human homologue (Di Nicolantonio et al., 2003).

Ligand-receptor binding leads to activation of G proteins through exchange of GTP for

GDP, resulting in the release of α and ßγ complexes, which mediate downstream

actions. AT1 receptors interact with various heterotrimeric G proteins including Gαq/11,

Gαi, Gα12 and Gα13 which couple to distinct signaling cascades. AT1 receptors are

primarily found in the circumventricular organs and other regions of the hypothalamus

in the brain (Song et al., 1992; MacGregor et al., 1995), in adrenal glands (Montiel et

al., 1993), in the conducting system and atrial epicardium of the heart (Allen et al.,

1990; Sechi et al., 1992; Saavedra et al., 1993), in the vasculature, including the aorta,

pulmonary and mesenteric arteries, as well as in glomerulo-mesangial and interstitial

cells in the kidney (Zhuo et al., 1992). In the vasculature, AT1 receptors are expressed

mainly in smooth muscle cells (Touyz et al., 1999). In the myocardium, AT1 receptors

are present on fibroblasts and cardiomyocytes (Allen et al., 2000). More recently, AT1

receptors have been found expressed in the human prostate, suggesting a role played by

Ang II in cell growth and sympathetic activity in that organ and in relation to urinary

flow (Dinh et al., 2001).

To date, AT1 receptors have been shown to mediate most of the physiological actions

of Ang II and this subtype is predominant in the control of Ang II-induced vascular

functions. Ang II stimulation of AT1 receptors in blood vessels causes vasoconstriction

leading to an increase in peripheral vascular tone and systemic blood pressure

(Gustafsson and Holstein-Rathlou, 1999). In the heart, AT1 receptors mediate the

inotropic and chronotropic effects of Ang II on cardiomyocyte contractility (discussed

in Chapter IV). AT1 receptors are believed to mediate cell growth and proliferation in

cardiac myocytes and fibroblasts, as well as in vascular smooth muscle cells, and can

induce the expression and release of various endogenous growth factors, such as FGF,

TGF-β1 and PDGF (Day et al., 1999; Rosenkranz, 2004; Skaletz-Rorowski et al.,

2004). These trophic effects of Ang II are implicated in the long-term processes of

cardiac remodeling and development of hypertrophy, as well as in the pathophysiology

of hypertension. Indeed, transgenic mice over-expressing AT1 receptors in cardiac

Chapter I

49

myocytes develop cardiac hypertrophy with no changes in blood pressure, and die

prematurely of heart failure (discussed in the Introduction of Chapter III). Additional

indirect effects of Ang II on cardiac remodeling could come from the facilitation of

sympathetic transmission by enhancing the AT1-dependent release of noradrenalin from

peripheral sympathetic nerve terminals. Moreover, Ang II stimulates the release of

catecholamines from the adrenal medulla and aldosterone from the adrenal cortex

(Figure I-8). Ang II also exerts diverse AT1-mediated actions on the brain by

modulating drinking behavior and salt appetite, central control of blood pressure,

stimulation of pituitary hormone release and has effects on learning and memory

(Fitzsimons, 1998; Badaue-Passos et al., 2001; Seltzer et al., 2004; Saavedra, 2005).

Renal AT1 levels mediate regulation of sodium and water reabsorption from the

proximal tubules and inhibition of renin secretion from the cells in the macula densa

(Kurtz and Wagner, 1999; Imanishi et al., 2003).

7.2.7. The AT2 receptor

Ang II binds to the AT2 receptor with similar affinity as to the AT1 receptor (de

Gasparo et al., 1995). The AT2 receptor is also a 7TM domain receptor and shares only

34% sequence identity with the AT1 receptor (Mukoyama et al., 1993). It is still unclear

whether the AT2 receptor is coupled to a G protein, but it has been reported that an

inhibitory Gα is linked to the receptor signaling pathway (Feng et al., 2002). The AT2

receptor has been mapped to chromosome X, in both humans and rodents (Lazard et al.,

1994; Hein et al., 1995). The AT2 receptor is highly expressed during fetal

development but rapidly declines at birth (Everett et al., 1997; Samyn et al., 1998). In

the human heart, the AT2 receptor localizes mainly to fibroblasts in interstitial regions

and atria, whereas a lower degree of expression is seen in the surrounding myocardium

(Brink et al., 1996). In rabbit hearts, AT2 receptors are almost totally absent from

nervous, conductive and atrial tissues, but are expressed at low levels in ventricles

(Brink et al., 1995). AT2 receptors are markedly expressed in the adrenals of most

species including humans (Belloni et al., 1998; Tanabe et al., 1998). In the kidney of

adult rats, AT2 localized mainly in glomeruli, while in human the same expression is

seen in glomeruli, tubules, and renal blood vessels. In the human uterine myometrium,

AT2 receptors are highly abundant, but are downregulated during pregnancy, possibly

Chapter I

50

due to sex hormones (Mancina et al., 1996). In the ovary, AT2 receptors are localized in

follicular granulose cells. More importantly, the expression of the AT2 receptor is

(variably) upregulated in pathological conditions, such as heart failure, renal failure,

myocardial infarction, brain lesions, vascular injury and wound healing. Ang II, insulin,

IGF-1 and some cytokines have been shown to stimulate AT2 receptor gene expression.

Since AT2 receptors are highly abundant in fetal tissues, it is believed that they play an

important role in fetal development. However, AT2-receptor knock-out (KO) mice

appear to develop and grow normally (see the Introduction of Chapter III), suggesting

that the AT2 receptor is not that crucial for fetal development. In the adult, AT2 receptor

binding is generally found to antagonize AT1-mediated signaling effects, by inhibiting

cell growth and by inducing vasodilation (Fukada et al., 2005). In support of this

hypothesis, AbdAlla et al. (2001) have reported that the AT2 receptor may antagonize

AT1 receptor-mediated responses by direct binding to the AT1 receptor. Recent studies

have shown that the AT2 receptor is involved in the production of cGMP, NO and

prostaglandins (Hiyoshi et al., 2005; Kim et al., 2005), suggesting a role in vascular

function, including vasodilation and blood pressure regulation. The AT2 receptor has

been suggested to play a role in neointimal repair and formation after injury (Okumura

et al., 2005). Moreover AT2 activation has been shown to promote nerve generation and

neuronal differentiation in cells of neuronal origin, through an increase in NO

production (Unger, 2001). However, AT2 receptors have also been shown to mediate

anti-proliferative effects in cultured coronary endothelial cells and vascular smooth

muscle cells and to induce both regeneration and apoptosis in different neuronal cell

types in vitro (Gallinat et al., 1997; Lucius et al., 1998).

Thus, it could be speculated that a change in the expression ratio of AT1 and AT2

receptors in a given tissue or cell population could determine the type and the potency

of the intracellular signal transduced and its corresponding physiological response.

Chapter I

51

7.3. Cardiac components of the renin-angiotensin system

Some years ago, the existence of a regulated cardiac-specific RAS was postulated. It is

now established that all the elements of the RAS cascade are present in normal cardiac

tissue. It is now beyond doubt that synthesis of ACE mRNA and protein does occur in

the normal adult myocardium, where AT1 and AT2 receptors and the precursor Agt are

expressed, but generally at very low levels (Dostal and Baker; 1999). It is also agreed

that renin is not synthesized in situ but predominantly taken up from the circulation

(Dostal et al., 1994; von Lutterotti et al., 1994). Nevertheless, experimental studies

performed with infusion of radio-labeled Ang I suggested that more than 75% of

cardiac Ang II is synthesized at tissue sites (van Kats et al., 1998). This would imply

that local Ang II production is of key importance in the pathophysiology of the RAS in

the heart.

Despite the lack of evidence for substantial renin production at cardiac tissue sites

under normal conditions, gradual increases in cardiac Ang II levels have been reported

in experimental models and clinically during the development of heart failure (Wollert

et al., 1999; Serneri et al., 2001). Numerous studies have demonstrated the efficacy of

RAS blockade in the treatment of cardiac remodeling and heart failure, independently

of the reduction in systemic blood pressure (Bohm et al., 1996; Tingleff et al., 1996).

Furthermore, elevated cardiac Agt levels are observed in various animal models of

pressure and volume overload cardiac hypertrophy, where Ang II is considered to

contribute significantly to cardiac remodeling through its growth promoting properties

(Sadoshima and Izumo, 1993a; Modesti et al., 2001).

Chapter I

52

8. METABOLIC REGULATION OF CARDIAC

HYPERTROPHY

8.1. Metabolism in the normal heart

Normal cardiac function requires substantial supply of energy in the form of ATP. ATP

production results from metabolic processing of glucose, free fatty acids (FFA),

pyruvate and ketone bodies (Rodriguez and McNeill, 1992). During fetal development,

the major source of energy comes from the glycolytic process, with mitochondrial

acetyl CoA (produced from pyruvate) as the major source of reducing equivalents

generating energy. After birth and at maturity, mitochondria utilize fatty acids as major

source for energy production, degrading them in the fatty acid β-oxidation spiral. In

general, it is known that acute changes in cardiac work result in instant activation of the

metabolic processes in a highly coordinated fashion. For example, when the workload

of the heart is doubled, oxygen consumption rate doubles; simultaneously, there is a

sudden coordinated increase in the oxidation of glycogen, glucose and lactate

(Goodwin et al., 1998). Moreover, lactate is oxidized preferentially when it constitutes

the prevailing carbohydrate substrate, as during exercise (Goodwin and Taegtmeyer,

2000). In contrast, rates of FFA oxidation do not change in response to acute changes in

workload (Goodwin and Taegtmeyer, 2000).

8.1.1. Fatty acid metabolism

The adult and postnatal mammalian heart relies on long-chain fatty acids as the

principal source and substrate for ATP production (Neely et al., 1972). The transition

from glucose to fatty acid consumption occurs after birth at a time when the

mammalian diet is composed almost entirely of high-fat breast milk. This period is

accompanied by a dramatic increase in the cardiac expression of genes encoding

Chapter I

53

enzymes in the mitochondrial fatty acid beta-oxidation pathway (Lockwood and Bailey,

1970; Nagao et al., 1993). The high postnatal cardiac expression of most nuclear genes

encoding mitochondrial fatty acid β-oxidation enzymes is controlled in part by the

transcription factor PPARα (peroxisome proliferator-activated receptor α). PPARα is a

member of an extended nuclear hormone receptor subfamily which includes the

vitamin D receptor, steroid receptors and thyroid hormone receptor (Desvergne and

Wahli, 1999). In the heart, activation of PPARα increases expression of genes involved

in the cellular fatty acid utilization pathways (Figure I-10).

Free fatty acids (FFA) are supplied by lipolysis from endogenous triglycerides, or from

blood, where they are transported bound to albumin or incorporated in chylomicrons

and very low density proteins (Rodrigues and McNeill, 1992). The transport of FFA

from the blood into mitochondria is mediated partly through passive diffusion and

partly through carrier membrane proteins (Van Der Vusse et al., 2000). Among the

transporters already identified in the heart, there is the fatty acid translocase CD36

(Luiken et al., 1999) and plasma membrane associated fatty-acids transport proteins

(Schaffer and Lodish, 1994). Once in the cytoplasm, long-chain fatty acids are bound

and transported by heart-type fatty acid-binding protein (H-FABP or FABP 3) and are

rapidly esterified to acetyl CoA. Long-chain fatty acids are then transported into

mitochondria and translocated across the inner mitochondrial membrane where they

enter into the β-oxidation spiral. The latter generates the electrons that are ultimately

transferred to the electron transport chain where ATP is produced at the expense of

oxygen consumption (that is, oxidative phosphorylation).

8.1.2. Glucose metabolism

Cellular glucose uptake, vital for subsequent glucose utilization (Opie, 1968), is limited

by the rate of glucose transport across the plasma membrane of cardiomyocytes, and is

mediated by glucose transporter proteins named GLUTs. The insulin-sensitive tissues,

(heart, skeletal muscles and adipose tissue) express at least two differentially regulated

glucose transporters, namely GLUT1 and GLUT4 (Pessin and Bell, 1992). GLUT4

protein levels increase postnatally and are maximal during adulthood (Castello et al.,

Chapter I

54

1991; Vanucci et al., 2000). In contrast, GLUT1 proteins are highly expressed during

fetal growth but downregulate rapidly after birth (Castello et al., 1991). There are

different factors which influence glucose transport in the heart; an increase in pyruvate,

lactate and fatty acid substrate can lead to a 50% downregulation of glucose transport

in isolated rat cardiomyocytes (Fischer et al., 1997a). Moreover, the basal glucose

uptake is significantly reduced under the influence of insulin, FFA and ketones (Kahn,

1996; Fischer et al., 1997b). In the cytosol, glycolytic metabolism of glucose-6-

phosphate to pyruvate precedes the mitochondrial oxidation of pyruvate. The rate

limiting enzyme for pyruvate production (and for both glycolysis and glucose

oxidation) is phosphofructokinase. This enzyme is strongly inhibited by ATP and by a

reduction in pH. An increase in pH and in levels of adenosine diphosphate (ADP) and

phosphate (Pi) increases the affinity of phosphofructokinase for its substrate. The

enzyme pyruvate dehydrogenase determines the rates at which pyruvate can enter the

citric acid cycle for oxidation or, during anoxia, be degraded to lactate (Opie, 1968).

Only 10-30% of energy required in the heart can be supplied by anaerobic glycolysis

and only for 1-2 minutes.

8.2. Metabolic adaptation in cardiac hypertrophy and failure

During load-dependent cardiac hypertrophy and other severe forms of cardiac

dysfunction, myocardial metabolism shifts towards a reliance on glycolysis as the

primary pathway for energy production, constituting a re-induction of the fetal energy

metabolic program (Bishop and Altschuld, 1970; Takeyama et al., 1995). Allard et al.

(1994) showed that the hypertrophic heart possesses an increased glycolytic capacity

and the contribution of ATP production from fatty acid oxidation is markedly

decreased. This metabolic shift results in lower oxygen consumption cost per mole of

ATP generated. Myocardial energy utilization pathways also undergo alterations during

cardiac hypertrophy, as evidenced by the downregulation of genes coding for various

ATPases, like the NaKATPase and SERCA2, as well as a shift toward fetal isoforms of

the contractile proteins. Thus re-activation of the ‘fetal’ metabolic program in the

hypertrophic ventricle likely comprises an adaptive metabolic energy response that

could have long-term consequences resulting from diminished energy reserves, reduced

Chapter I

55

capacity to maintain myocyte lipid balance and altered ionic and calcium homeostasis,

ultimately leading to cardiac dysfunction. Interestingly, in hypertensive rabbits

developing pressure-overload cardiac hypertrophy, this shift occurs before there is any

change in cardiac mass (Taegtmeyer and Overturf, 1988), suggesting that metabolic

adaptation to changes in workload may precede cardiac hypertrophy.

Cardiac hypertrophy is accompanied by downregulation of various genes involved in

fatty acid metabolism. PPARα transcript levels are downregulated in hypertrophic

hearts (Young et al., 2001). Moreover, PPARα activity is decreased at the post-

transcriptional level during cardiac hypertrophy, possibly by a phosphorylation event.

MAPK pathways are an obvious candidate for mediating the deactivation of PPARα

given that they have been implicated in the hypertrophic growth (Schaub et al., 1997;

Barger et al., 2000) and are known to regulate nuclear receptors, including the PPAR

sub-family (Hu et al., 1996). Indeed these results indicate that the activity of the PPAR

gene regulation pathway is rapidly reduced by the action of the ERK pathway, but not

by the p38 and the JNK pathways. Interestingly, Ang II downregulates the expression

of PPARα and PPARγ leading to the activation of the pro-hypertrophic NFκb signaling

pathway (Tham et al., 2002). Moreover, PPARγ activators, such as the anti-diabetic

agents thiazolidinediones, inhibit Ang II-induced pressure-overload cardiac

hypertrophy in mice, and prevent the upregulation of skeletal α-actin and ANF genes in

neonatal cultured cardiomyocytes, thus suppressing hypertrophic growth (Asakawa et

al., 2002).

8.3. Cardiac metabolism in type 2 diabetes and insulin resistance states

8.3.1. Diabetes mellitus

Diabetes mellitus refers to a diverse group of metabolic diseases that are characterized

by high plasma glucose concentrations and that have been estimated to affect more than

108 people worldwide (King et al., 1998). Type 2 diabetes, previously called ‘adult

onset’ or ‘non-insulin-dependent’ diabetes mellitus (NIDDM), accounts for more than

Chapter I

56

90% of cases and, in several countries, affects up to 20% of adults (King et al., 1998).

Insulin resistance (see below) typically precedes the onset of type 2 diabetes and

predisposes individuals to various cardiovascular diseases, including hypertension,

hyperlipidemia, premature atherosclerosis, coronary artery diseases and heart failure

(Reaven, 1991; Anker et al., 1997; Grundy et al., 1999). Type 1 diabetes (or ‘juvenile-

onset’ or ‘insulin-dependent’ diabetes mellitus, IDDM) accounts for the other 10% and

it is characterized by the failure of insulin secretion by pancreatic β-cells due to

autoimmune destruction (Todd, 1999). In human heart muscle, both the myocardial

glucose uptake and glycolytic flux to lactate are decreased in type 1 diabetes.

Normalization of blood glucose with insulin supplementation restores the patterns of

lactate and ketone body kinetics (Avogaro et al., 1990).

Hyperglycemia defines both types of diabetes and results from an absolute insulin

deficiency in type 1 diabetes and tissue insulin resistance in type 2 diabetes (American

Diabetes Association, 1997). High circulating levels of glucose cause accelerated

micro- and macro-vascular diseases (such as ischemic heart diseases, stroke,

retinopathy, neuropathy and nephropathy) and increase morbidity and mortality in

diabetic patients (Klein, 1995). Diabetes is a strong independent cardiovascular risk

factor, and the likelihood of death from cardiovascular causes is two to five fold higher

in diabetics (Kannel and McGee, 1979; Stamler et al., 1993). Clinically, diabetes

mellitus is frequently associated with a diabetic cardiomyopathy which is not directly

attributable to microvascular disease, hypertension or obesity (see below) (Grundy et

al., 1999; Hayat et al., 2004).

8.3.2. Insulin resistance

Insulin resistance is a principal feature of type 2 diabetes and precedes the clinical

development of the disease by several years in patients. It is defined as a relative

inability of insulin to promote glucose transport into peripheral target tissues

(especially skeletal muscle) (Petersen and Shulman, 2002). In physiological conditions,

insulin promotes glucose uptake, which is rate limiting for cellular glucose metabolism

under most physiological settings (Manchester et al., 1994). In individuals with normal

glucose tolerance, a strong correlation exists between GLUT4 protein abundance in

Chapter I

57

muscles and the rates of glucose disposal during insulin clamp studies. This correlation

between peripheral insulin action and muscle GLUT4 is lost once diabetes mellitus is

established, suggesting that GLUT4 is a primary effector molecule for insulin-mediated

glucose disposal and that GLUT4 downregulation contributes significantly to the

development of insulin resistance in peripheral tissues.

In addition to diabetes, insulin resistance predisposes individuals to hypertension,

hyperlipidaemia, premature atherosclerosis, coronary artery diseases, muscle wasting,

left ventricular hypertrophy and chronic heart failure (Paolisso et al., 1991; Reaven,

1991; Anker et al., 1997). Risk factors for insulin resistance include inactivity, obesity,

ageing (Ferrannini et al., 1996), elevated levels of TNF-α (Hotamisligil and

Spiegelman, 1994) and increased oxidative stress (Paolisso and Giugliano, 1996).

Possible links between insulin resistance, hypertension and cardiovascular disease

include sodium retention, hyperaldosteronism, over-activity of the sympathetic nervous

system, and vascular smooth muscle proliferation and hypertrophy (DeFronzo, 1992).

8.3.3. Cardiac glucose transport in states of insulin resistance

Under conditions of increased workload or anoxia, impaired glucose uptake due to

insulin resistance could be a critical factor in the onset of metabolically-induced cardiac

remodeling (Belke et al., 2000; Semeniuk et al., 2002; Desrois et al., 2004).

As mentioned above, glucose is transported into cardiomyocytes via GLUT4 and

GLUT1. It is accepted that the effect of insulin on cardiac glucose disposal is mediated

through its action on GLUT4, which is the most abundant isoform expressed in the

adult heart (Abel, 2004; Desrois et al., 2004). In the heart GLUT4 translocates to the

plasma membrane in response to insulin (Egert et al., 1997), ischemia (Egert et al.,

1997), hypoxia (Sun et al., 1994), and high-frequency contraction (Till et al., 1997).

Decreased GLUT4 activity and expression is suggested as one of the factors

responsible for metabolic and contractile dysfunction in the diabetic heart, where

decreased glucose uptake is comprimized (Eckel and Reinauer, 1990; Garvey et al.,

1993; Desrois et al., 2004; Rosenblatt-Velin et al.,2004).

Chapter I

58

The mechanisms governing the differential regulation of GLUT4 in metabolic

pathologies and the relationship between altered glucose transporter expression and

cardiac function are incompletely understood. One approach to address this issue has

been to manipulate cardiac and body glucose transporter expression in transgenic mice.

Mice genetically engineered to overexpress the GLUT4 gene in insulin-sensitive tissues

(cardiac and skeletal muscle and adipose tissue) display enhanced insulin

responsiveness, cellular glucose uptake and peripheral glucose utilization, suggesting

that indeed GLUT4 is the primary effector molecule for insulin-mediated glucose

disposal. On the other hand, global or tissue-specific disruption of the GLUT4 leads to

insulin resistance, glucose intolerance and the development of mild-to-severe cardiac

hypertrophy. These transgenic and knock-out models for the GLUT4 protein are

presented and discussed in detail in the Introduction to Chapter III.

8.3.4. Cardiac fatty acid regulation in states of insulin resistance

In contrast to the metabolic shift observed during the development of most forms of

load-dependent cardiac hypertrophy, cardiac remodeling induced by insulin resistance

and hyperglycemia is associated with a primary increase in cardiac fatty acid

metabolism. Myocardial uptake of ketone bodies and FFA was shown to be increased

in diabetic patients compared with controls (Avogaro, 1990). Levels of fatty acid

binding protein were significantly elevated (+40-50%) in the hearts of experimental

diabetic rats (Engels et al., 1999). Protein levels of fatty acid translocase CD36, located

in the pericapillary myocardial tissues, were increased 2 to 4 fold (Pelsers et al., 1999).

Rat heart fatty acid-binding protein content was 34% higher in type 1 diabetic rats and

103% higher in type 2 diabetic rats compared with control (Glatz et al., 1994).

Moreover, the tissue triacylglycerol level in the heart is strikingly increased in various

models of diabetes (Paulson and Crass, 1982). In isolated diabetic rat hearts perfused

under diabetic metabolic conditions, elevated triacylglycerol levels were maintained,

lipolysis was reduced or unchanged and triacylglycerol synthesis was enhanced. When

the same hearts were perfused simulating normal metabolic conditions, lipolysis and β-

oxidation were markedly enhanced (Paulson and Crass, 1982). In states of diabetes and

insulin resistance, accumulation of fatty acids within the cytoplasm occurs (Rodrigues

and McNeill, 1992). This can cause various deleterious electrophysiological,

Chapter I

59

biochemical and mechanical effects in cardiac muscle cells: accumulation of fatty acids

and their toxic intermediates is associated with depression of contractility (Criddle et

al., 1990), combined with inhibition of the NaKATPase (Dhalla et al., 1991), NCX

(Ashavaid et al., 1985), and the SERCA2 (Adams et al., 1979). Fatty acids have been

demonstrated to directly interact with voltage-dependent Ca2+ channels (Spedding and

Mir, 1987).

8.4. Diabetic cardiomyopathy

Accumulating data from experimental and clinical studies have shown that diabetes

mellitus results in cardiac functional and structural changes, independent of

hypertension, coronary artery disease, or any other known cardiac disease. These

observations support the existence of a distinct diabetic cardiomyopathy. The

development of diabetic cardiomyopathy is likely to be multifactorial. Putative

mechanisms include metabolic disturbances, myocardial fibrosis, autonomic

dysfunction, and insulin resistance (reviewed in Fang et al., 2004).

8.4.1. Metabolic disturbances in diabetic cardiomyopathy

Increasing evidence suggests that altered substrate supply and utilization by cardiac

myocytes could be the primary injury in the pathogenesis and development of diabetic

cardiomyopathy (Rodrigues et al., 1998). As mentioned above, the major restriction to

glucose utilization in the diabetic heart is the slow rate of glucose transport across the

sarcolemmal membrane into the cardiomyocyte, probably due to the cellular depletion

of glucose transporters, such as the GLUT4. A second mechanism of reduced glucose

oxidation is via the inhibitory effect of fatty acid oxidation on the pyruvate

dehydrogenase complex due to high circulating FFA (Liedtke et al., 1988). This has the

net effect of reducing ATP availability and may be more detrimental in type II diabetes,

in which FFA levels tend to be higher.

Chapter I

60

8.4.2. Mechanical dysfunction in diabetic cardiomyopathy

Eventually, altered substrate supply, insulin resistance and hyperglycemia will lead to

defects in calcium and other ion transport, myocyte apoptosis and necrosis,

upregulation of the RAS and TGF-β1 activity, resulting in myocyte injury, myocardial

fibrosis and early diastolic dysfunction. This stage of diabetic cardiomyopathy is

mainly characterized by cardiomyocyte hypertrophy, fibroblast proliferation,

inflammatory response and myocardial fibrosis (these aspects are detailed in Chapter

IV). Prolonged changes in metabolism, mitochondrial stress and development of

myocardial fibrosis result in severe cardiac microvascular structural and functional

changes. Alterations in cardiac structure and function are marked and at this stage

diabetic cardiomyopathy is frequently associated with overt systolic dysfunction,

hypertension and early development of ischemic heart disease (Brands and Fitzgerald,

2002).

Chapter I

61

9. AIMS

In the following Chapters, the findings of an investigation of two different transgenic

mouse models of primary cardiac hypertrophy are presented: 1) a cardiac-specific

angiotensinogen-overexpressing transgenic mouse, the TG1306/1R; and 2) a glucose

transporter (GLUT4) knock-out mouse model, the GLUT4-KO. The objective of these

studies is to provide new insight into the mechanisms that are common and/or unique in

the development of cardiac hypertrophic states induced either by the cardiac-specific

overproduction of the bioactive peptide angiotensin II or a cardiomyocyte-specific

glucose metabolic deficiency. The general hypothesis addressed is that cardiac-specific

neuro-endocrine activation or metabolic imbalance are sufficient provocations (in the

absence of hemodynamic disturbance) to induce hypertrophic structural, functional and

molecular remodeling of the myocardium. More specific experimental hypotheses are

postulated in relation to the investigation detailed in the Chapters to follow.

Experimental investigations described include cellular and tissue morphometric

analysis on the hearts of these rodent models, combined with examination of

cardiomyocyte contractility and cardiac gene expression profiling by cDNA microarray

assay.

Chapter I

62

Figure I-1: Intracellular ionic homeostasis during cardiomyocyte contraction and relaxation

An overview of the mechanisms of intracellular calcium ionic homeostasis in cardiomyocytes during contraction (A. to E.) and relaxation (F. to I.). Refer to Section 5 of the present Chapter for more explanations. Briefly: after electrical depolarization, elevation of calcium depends upon calcium entry through voltage-dependent channels (VOC) in the plasma membrane, the sodium/calcium exchanger (NCX) and calcium release through sarcoplasmic reticulum (SR) ryanodine receptors (RyR), via the calcium-induced calcium release (CICR) mechanism. The mechanisms which remove calcium from the cytosol to the extracellular environment include ion pumps (calcium ATPases) and exchange (the NCX). The majority of calcium ions are pumped back into the SR through the SR calcium ATPase SERCA2.

[Ca2+]i

ATPase NaK-ATPase

SERCA2

VOC

[Ca2+]e

RyR-dependent calcium pools

[Na+]e

[Na+]i

[K+]e

[K+]i

[Ca2+]e

NCX

NCX

CICR CICR

Contraction

[Na+]e

[Na+]i

Depolarization

Repolarization

Relaxation

B.

C.

D. A.

E.

F.

G.H.I.

[Ca2+]i

ATPase NaK-ATPase

SERCA2

VOC

[Ca2+]e

RyR-dependent calcium pools

[Na+]e

[Na+]i

[K+]e

[K+]i

[Ca2+]e

NCX

NCX

CICR CICR

Contraction

[Na+]e

[Na+]i

Depolarization

Repolarization

Relaxation

B.

C.

D. A.

E.

F.

G.H.I.

o o

ooo

[Ca2+]i

ATPase NaK-ATPase

SERCA2

VOC

[Ca2+]e

RyR-dependent calcium pools

[Na+]e

[Na+]i

[K+]e

[K+]i

[Ca2+]e

NCX

NCX

CICR CICR

Contraction

[Na+]e

[Na+]i

Depolarization

Repolarization

Relaxation

B.

C.

D. A.

E.

F.

G.H.I.

[Ca2+]i

ATPase NaK-ATPase

SERCA2

VOC

[Ca2+]e

RyR-dependent calcium pools

[Na+]e

[Na+]i

[K+]e

[K+]i

[Ca2+]e

NCX

NCX

CICR CICR

Contraction

[Na+]e

[Na+]i

Depolarization

Repolarization

Relaxation

B.

C.

D. A.

E.

F.

G.H.I.

oo oo

oooooo

Chapter I

63

DSH

PKC

AxinAPC

GSK3

LRP

P

β-catenin β-cateninP

•Cyclin D1, D2, D3•c-MYC, c-JUN•MMP7•BMPs

RhoA/Rac JNK2/3 Apoptosis

LDLR

PP2A

UbiquitinationProteasome degradation

β-catenin

TCF

WntDkk

DSH

PKC

AxinAPC

GSK3APC

GSK3

LRP

P

β-catenin β-cateninP

β-cateninP

•Cyclin D1, D2, D3•c-MYC, c-JUN•MMP7•BMPs

RhoA/Rac JNK2/3 Apoptosis

LDLR

PP2A

UbiquitinationProteasome degradation

β-catenin

TCF

WntDkk

Figure I-2: Embryonic heart formation

Heart formation is cued by a combination of positive and negative signals from surrounding tissues. Inhibitory signals that block heart formation in anterior paraxial mesoderm include Wnt family members expressed in dorsal neural tube and anti-BMPs expressed in the axial tissues (i.e., Noggin in the notochord). Wnt signaling pathway, which is essential for setting up the entire body pattern during embryonic development involves glycogen synthase kinase-3 (GSK3). In the absence of Wnt signaling, GSK3 is active and phosphorylates beta-catenin resulting in its degradation by ubiquitin-mediated proteolysis. Activation of Wnt signaling inhibits GSK3, thereby preventing phosphorylation of beta-catenin, which is then able to move to the nucleus. There it associates with members of the LEF-1/TCF family of transcription factors, which activate the transcription of genes like cyclin-D1, Myc, and MMPs. The Wnt signaling pathway is blocked by a family of secreted proteins such as Dickkopf-1 (Dkk-1) sufficient for induction of heart formation in posterior mesoderm. BMP signaling can also be blocked by the BMP antagonists Noggin and Chordin, which are secreted from the notochord and cooperate with Wnts to prevent cardiogenesis. Receptors for BMPs are persistently expressed during cardiac development. Activin receptor-like kinase 3 (ALK3), a BMP receptor, is specifically required at mid-gestation for normal development of the trabeculae, compact myocardium, interventricular septum, and endocardial cushion. Cardiac muscle lacking ALK3 is specifically deficient in expressing TGFβ, an established paracrine mediator of cushion morphogenesis.

Chapter I

64

cFOS

Gβγ GαGDP

GαsGTP

GαiGTP

GαqGTP

Gα12GTP

PKAAC

ATP

cAMP

CRE

CREBP

P

RAS Erk1/2

A.

B.

SRE

ELKP

SRE

STATP

C.

NFAT-RE

NFAT

AP-1

cJUN

RhoAAKAP13

PKA GTPStress fiberformation

PLDD.

PLC

PIP2

IP3

DAG PKC

Ca2+ CaMCalcineurin

cFOS

Gβγ GαGDP

GαsGTP

GαiGTP

GαqGTP

Gα12GTP

PKAAC

ATP

cAMP

CRE

CREBP

P

RAS Erk1/2

A.

B.

SRE

ELKP

SRE

STATP

C.

NFAT-RE

NFAT

AP-1

cJUN

RhoAAKAP13

PKA GTPStress fiberformation

PLDD.

PLC

PIP2

IP3

DAG PKC

Ca2+ CaMCalcineurin

Figure I-3: Signaling through 7TM receptors and G-protein

A. In general, Gαs proteins couple to stimulation of adenylyl cyclase (AC) which increases cyclic adenosine monophosphate (cAMP) levels. cAMP is an activator of protein kinase A (PKA), a serine/threonine kinase that phosphorylate many substrates, including 7TM receptors, other kinases and transcription factors. B. Gαi/o couples to inhibition of AC as well as to activation of G-protein coupled inwardly rectifying potassium channels, increases c-Src tyrosine kinase activity and stimulates ERK/mitogen-activated protein kinase (MAPK) pathway. C. Gα12/13 proteins couple to the activation of the Rho guanine-nucleotide-exchange factors (proteins that facilitate the replacement of GDP with GTP) and stress-fiber formation. Gα12/13 proteins also modulate phospholipase D (PLD) activity, probably via RhoA and c-Src activity. D. Gαq/11 activates phospholipase C beta (PLCβ), which hydrolyses phosphatidylinositol 1,4-bisphosphate (PIP2) to inositol triphosphate (IP3) and diacylglycerol (DAG). IP3 binds to specific receptors on the endoplasmic reticulum (especially on endothelial and smooth muscles cells) inducing the release of calcium into the cytoplasm. DAG is a potent activator of the family of serine/threonine kinases, protein kinase C (PKC). Finally, the Gβγ subunit has been linked to modulation of inwardly rectifying potassium channels, GPCR phosphorylation and desensitization mediated by G protein-coupled receptor kinases, as well as activation of AC, PLCβ and phosphatidylinositol 3-kinase (PI3K).

Chapter I

65

EGF

Gβγ GαqGTP PLCγ

Angiotensin IIEndothelin I

DAG PKC

IP3

Adam12

EGF

RASRho

Altered myocardialcell morphology

MAPK pathway

c-FOSc-MYC c-JUNNFκb

Activation of immediate early genes

EG

F-R

EG

F-R

Cardiac/cardiomyocytehypertrophy

ROS

EGF

Gβγ GαqGTP PLCγ

Angiotensin IIEndothelin I

DAG PKC

IP3

Adam12

EGF

RASRho

Altered myocardialcell morphology

MAPK pathway

c-FOSc-MYC c-JUNNFκb

Activation of immediate early genes

EG

F-R

EG

F-R

Cardiac/cardiomyocytehypertrophy

ROS

Figure I-4: Signaling through EGF receptor transactivation

Cardiomyocyte hypertrophic stimuli can result from cross-talk between GPCR signaling and the EGF receptor pathway. Several GPCR ligands are known to stimulate cardiac hypertrophy, including factors that regulate blood pressure such as angiotensin II and endothelin-1. These factors stimulate phospholipase C (PLC) through Gαq activation, and the production of IP3 and diacylglycerol (DAG) second messengers. PKC is activated by DAG and interacts with the metalloproteinase ADAM12. ADAM12 cleaves the membrane-bound HB-EGF to release soluble EGF ligand that activates EGF receptor in myocardial cells. EGF receptor activation downstream through small G proteins (Ras, Rho) and the MAPK pathway ultimately leads to cardiac hypertrophy. Signals by GPCR ligands such as angiotensin II result in transcriptional translation of immediate early genes like fos and other genes involved in long-term remodeling of heart tissue and the physiological response to stress in the heart such as the atrial natriuretic factor. Factors such as the AKT kinase, reactive oxygen species (ROS) and NF-kB also are involved in signaling that leads to hypertrophy, although their role is not yet clear.

Chapter I

66

RACRAS

PAKMAP4K3/4GCK

MAP3K2/3/8

MEKK1

MAP2K5

ERK5

RAF

MEK1/2

ERK1/2

MAP3K4/12

MKK7 MEK4

JNK1 MAPK9

MAPK10

MAP4K1/5

ASK1 TAK1

MAP3K11

MEK3/4

MAPK11/12p38

ERK6

MAP

4K

MAP

3K

MAP

2K

MAP

K

Targ

ets

of M

APK

Other kinases Other kinases

c-MYCMEF2 CREB STAT1

ELK-1 c-FOS c-JUN ATF-2 SP1

SHC

GR

B2 SOS

Growth factors, UVtrophic factors, etc

Stress: osmotic shockγ-radiations

GPCRRTK

FASL, inflammatory cytokines, dead receptors, UV, etc.

TRAD

DTR

AF2

NIK

IKK

NFκbCaspasecascade

RACRAS

PAKMAP4K3/4GCK

MAP3K2/3/8

MEKK1

MAP2K5

ERK5

RAF

MEK1/2

ERK1/2

MAP3K4/12

MKK7 MEK4

JNK1 MAPK9

MAPK10

MAP4K1/5

ASK1 TAK1

MAP3K11

MEK3/4

MAPK11/12p38

ERK6

MAP

4K

MAP

3K

MAP

2K

MAP

K

Targ

ets

of M

APK

Other kinases Other kinases

c-MYCMEF2 CREB STAT1

ELK-1 c-FOS c-JUN ATF-2 SP1

SHC

GR

B2 SOS

Growth factors, UVtrophic factors, etc

Stress: osmotic shockγ-radiations

GPCRRTK

FASL, inflammatory cytokines, dead receptors, UV, etc.

TRAD

DTR

AF2

NIK

IKK

NFκbCaspasecascade

Figure I-5: The MAPK signaling cascade

The ever evolving mitogen-activated protein kinase (MAP kinase) pathways consist of several groupings and numerous related proteins which constitute interrelated signal transduction cascades activated by stimuli such as growth factors, stress, cytokines and inflammation. The four major groupings illustrated here are the Erk (purple), JNK or SAPK (blue), p38 (green) and the Big MAPK or ERK5 (light blue) cascades. Signals from cell surface receptors such as GPCRs and growth factor receptors are transduced, directly or via small G proteins such as Ras and Rac, to multiple tiers of protein kinases that amplify these signals and/or regulate each other. In some cascades the first activation tier involves the MAPKKKKs, MAP kinase kinase kinase kinases or MAP4K proteins. The next tier are the serine/threonine MAPKKKs, MAP kinase kinase kinase or MAP3Ks such as RAF, TAK, ASK, and MEKK1. This level has the greatest amount of cross-communication curently known. The serine/threonine/tyrosine MAPKKs, MAP Kinase kinases or MAP2Ks, such as the MKK and MEK kinases, are one step up from the MAP kinase cascade, phosphorylating and activating these kinases. The focal tier, the MAPKs or MAP kinases includes JNK1, p38, and ERKs, and are the kinases that give each cascade its name. The endpoints of these cascades, shown in the bottom tier, include the MAPK activated protein kinases (MAPKAPK) and some of the numerous transcription factors that regulate genes involved in apoptosis, inflammation, cell growth and differentiation.

Chapter I

67

RAS

RAF1MEK1ERK1

ERK2

PI3K GSK3β

GSK3β

AKT

p70s6K

p38

JNK1

NF-ATc NF-ATcP

P

AKT pathway

β-adrenergic agonistsAngiotensin IIEndothelin I

IGF-1 FGF2EGF LIF

Ca2+

CaM CaM

CAMK I

CAMK IV

IP3

Calcineurin

NF-ATcMEF2c

GATA4

ANFα-actinβ-myosinET-1

(?)

PKA

RAS

RAF1MEK1ERK1

ERK2

PI3K GSK3β

GSK3β

AKT

p70s6K

p38

JNK1

NF-ATc NF-ATcP

P

AKT pathway

β-adrenergic agonistsAngiotensin IIEndothelin I

IGF-1 FGF2EGF LIF

Ca2+

CaM CaM

CAMK I

CAMK IV

IP3

Calcineurin

NF-ATcMEF2c

GATA4GATA4

ANFα-actinβ-myosinET-1

(?)

PKA

Figure I-6: Hypertrophic growth through calcineurin and NF-ATc

A single transcriptional regulator initially associated with the activation of the T-cells (NFATc4) has been shown to link embryonic heart development (and genetic and environmental causes of congenital heart disorders) to acquired cardiac hypertrophy. Within the embryonic endocardium, specific inductive events appear to activate NF-ATc, which are localized to the nucleus only in endocardial cells that are adjacent to the interface with the cardiac jelly and myocardium, and that are thought to give the inductive stimulus to the valve primordia. Treatment with FK506, a specific calcineurin inhibitor, prevents nuclear localization of NF-ATc4. More recently, it has been demonstrated that activated CaMK stimulates calcineurin, which than acts through NF-ATc4 in association with GATA4, to induce cardiac hypertrophy. A model for the proposed role of calreticulin in the regulation of cardiac development requires a myogenic signal from extracellular space to activate the production of IP3 that results in the release of calcium ions from ER under the regulation of calreticulin (CRT). Increased intracellular calcium binds to calmodulin (CaM) and activates calcineurin. Calcineurin dephosphorylates NF-ATc4 that translocates to the nucleus. In the nucleus NF-AT forms complexes with the GATA-4 and other transcription factors leading to activation of transcription of genes (ANF, α-actin, β-myosin, TNFα, ET-1, Adss1 etc) essential for embryonic cardiac development and cardiac hypertrophy in the adult heart.

Chapter I

68

Rip

TRA

DD

TRA

F2

Bcl-2

Caspase 8

FaddCaspase 10

NIK

IKK

NFκb

Bid

Iκb

IκbNFκb

tBid

Cytochrome C

Apaf-1Caspase 9

Caspase 3

lap

Caspase 6

Dr4

/5

Dr3

Caspase 7

Parp

DNA repair

DNA fragmentationCell shrinkage

Apoptosis

Rip

TRA

DD

TRA

F2

Bcl-2

Caspase 8

FaddFaddCaspase 10

NIK

IKK

NFκb

BidBid

Iκb

IκbNFκb

tBidtBid

Cytochrome C

Apaf-1Apaf-1Caspase 9

Caspase 3

lap

Caspase 6

Dr4

/5

Dr3

Caspase 7

ParpParp

DNA repair

DNA fragmentationCell shrinkage

Apoptosis

Figure I-7: Mechanisms of caspase-induced apoptosis

Stimuli can arise from the nucleus as well as from the mitochondrion or the membrane. Ultimately, the stimuli converge on the process of activation of caspases (cysteine proteases) enzymes, whose many substrates are thought to account for terminal events of apoptosis. Apoptosis is specifically induced via signaling through a family of receptors known collectively as 'death receptors' including Fas, TNFR, DR3, -4 and -5. Death receptor ligands (e.g. ApoL, FasL, etc.) characteristically initiate signaling via receptor oligomerization, recruitment of specialized adaptor proteins and activation of caspase cascades. Apo3L (or FasL) recruits initiator caspase 8 via the adapter protein FADD. Caspase 8 then oligomerizes and is activated via autocatalysis. Activated caspase 8 stimulates apoptosis via two parallel cascades: it directly cleaves and activates caspase-3, and it cleaves Bid (a Bcl-2 family protein). Truncated Bid (tBid) translocates to mitochondria, inducing cytochrome C release, which sequentially activates caspases 9 and 3. DR-3L can deliver pro- or anti-apoptotic signals. DR-3 promote apoptosis via the adaptor proteins TRADD/FADD and the activation of caspase 8. Alternatively, apoptosis is inhibited via an adaptor protein complex including RIP which activates NF-kB and induces survival genes including IAP. Induction of apoptosis via Apo2L requires caspase activity, but the adaptor requirement is unclear.

Chapter I

69

Angiotensinogen

Angiotensin I

Angiotensin II

ACE

ACEinhibitors

Bradykinin

Nitric OxidePGE2

Inactivekinin

fragments

AT1 receptor

Smooth muscle proliferation

•Sodium resorption•Aldosterone release•Renal blood flow

•Direct vasoconstriction•Peripheral noradrenergic activity•Adrenal catecholamine release•CNS sympathetic activity

Vasoconstriction

Blood volume

Cardiac and vascular remodeling

AT2 receptor

Antiproliferative andvasodilatory effects

Renin

AT1blockers

Angiotensinogen

Angiotensin I

Angiotensin II

ACE

ACEinhibitors

Bradykinin

Nitric OxidePGE2

Inactivekinin

fragments

AT1 receptor

Smooth muscle proliferation

•Sodium resorption•Aldosterone release•Renal blood flow

•Direct vasoconstriction•Peripheral noradrenergic activity•Adrenal catecholamine release•CNS sympathetic activity

Vasoconstriction

Blood volume

Cardiac and vascular remodeling

AT2 receptor

Antiproliferative andvasodilatory effects

Renin

AT1blockers

Figure I-8: The renin-angiotensin system (RAS)

The most familiar and best studied effects of angiotensin II (Ang II) are vasoconstriction and stimulation of synthesis and secretion of aldosterone by the adrenal cortex. However, the octapeptide has numerous other effects. Some involve stimulation of the heart and the sympathetic nervous system, these complement the direct vasomotor effects and contribute to increase in blood pressure caused by Ang II. Others, such as stimulation of drinking and increased secretion of anti-diuretic hormone, complement the effects of aldosterone and contribute to retention of sodium and water.

Chapter I

70

Agt

Ang I

Ang II

Ang III Ang IV

Ang [2-10]

Ang [1-7]

Ang [1-9]

Ang [1-5]

Renin

Toni

nC

athe

psin

G

AMP

ACEChymaseCathepsin A

AMPD-Amp

AMP

ACE

ACE2

PEPNEP

ACENEP

ACE

ACE2 PEPPCP

AMPACE

Agt

Ang I

Ang II

Ang III Ang IV

Ang [2-10]

Ang [1-7]

Ang [1-9]

Ang [1-5]

Renin

Toni

nC

athe

psin

G

AMP

ACEChymaseCathepsin A

AMPD-Amp

AMP

ACE

ACE2

PEPNEP

ACENEP

ACE

ACE2 PEPPCP

AMPACE

Figure I-9: generation and degradation of angiotensin peptides

Schematic representation of the enzymatic pathways involved in the generation of angiotensin peptides. Abbreviations: ACE = angiotensin-converting enzyme; Ang = angiotensin; AMP = aminopeptidase; D-Amp = dipeptidyl-aminopeptidase; PCP = prolyl-carboxypeptidase; PEP = prolyl-endopeptidase; NEP = neutral-endopeptidase.

Chapter I

71

Diet, release from adipose

Fatty acids (PUFA, MUFA, Saturated, Conjugated)Intermediary metabolism

PPARα

PPRE ?RE

PPRE-regulated genes Non-PPRE-regulated genes

Acyl-CoA OxidaseCytochromes P450Fatty Acid Binding ProteinAcyl-CoA Binding ProteinAcyl-CoA Synthetase17β-HSD IVLipoprotein LipaseCarnitine PalmitoylTransferase 1LXRα

Inducible NO SynthaseCyclooxygenase 2ProlactinFAT/CD36C-MycIκBαTNFα

Diet, release from adipose

Fatty acids (PUFA, MUFA, Saturated, Conjugated)Intermediary metabolism

PPARα

PPRE ?RE

PPRE-regulated genes Non-PPRE-regulated genes

Acyl-CoA OxidaseCytochromes P450Fatty Acid Binding ProteinAcyl-CoA Binding ProteinAcyl-CoA Synthetase17β-HSD IVLipoprotein LipaseCarnitine PalmitoylTransferase 1LXRα

Inducible NO SynthaseCyclooxygenase 2ProlactinFAT/CD36C-MycIκBαTNFα

Figure I-10: PPARα-dependent metabolic genes transcription

Similar to other nuclear hormone receptors, PPAR elements act as a ligand activated transcription factors. Upon binding fatty acids or hypolipidemic drugs, PPARα interacts with RXR and regulates the expression of target genes via a proliferator-response element (PPRE)-regulated mechanism (classic mechanism) or via other PPRE-independent mechanisms. The majority of genes regulated by PPARs are involved in the metabolism and catabolism of fatty acids. PPARγ, another member of the PPAR family, is activated by prostaglandins, leukotrienes and anti-diabetic thiazolidinediones and affects the expression of genes involved in the storage of fatty acids. PPARb is only weakly activated by fatty acids, prostaglandins and leukotrienes and has no known physiologically relevant ligand.

CHAPTER II

General methods

Chapter II

73

1. EXPERIMENTAL MODELS

1.1. Ethics approval

All animals were handled in the manner specified by the Prevention of Cruelty to

Animals Act 1986 and NHMRC/CSIRO/ACC Australian Code of Practice for the Care

and Use of Animals for Scientific Purposes (1997). In addition, other conditions

specified by the University of Melbourne Animal Experimental Ethics Committee were

followed.

1.2. The transgenic angiotensinogen overexpressing mouse

(TG1306/1R)

The generation and the descriptive characteristics of the transgenic heterozygous

TG1306/1R mouse are reported in the Introduction of Chapter III. Transgenic

heterozygous (TG) and wild-type (WT) littermate control mice were obtained from the

breeding colony in the Biological Research Facility (Faculty of Medicine, University of

Melbourne), established from male heterozygous TG1306/1R founders imported from

Switzerland (Division of Hypertension, University of Lausanne Medical School). Mice

were subject to Australian Quarantine and Inspection Service (AQIS). Heterozygous

TG breeders were mated with C57BL6 strain females to generate litters composed of

half TG heterozygous and half WT progeny. Genotyping was done on DNA extracted

from tail biopsies by semi-quantitative PCR. Duplex genomic PCR reactions were

performed with specific primers detecting the transgene and the housekeeping gene

glyceraldehydes-3-phosphate dehydrogenase (GAPDH) as positive control. Primer

sequences were as follow: 5’-AAGCCCATCACCATCTTCCAGGAG-3’ (forward)

and 5’-AGCCCT TCCACAATGCCAAAG-3’ (reverse) for GAPDH (product size: 308

bp); 5’-ACAGCAGATCACGATTCTCCCG-3’ (forward) and 5’-

Chapter II

74

CAGGTCAGGATGCAGAAGATGG-3’ (reverse) for the angiotensinogen transgene

(product size: 397 bp). Animals received standard laboratory chow (Clarke King,

Australia) consisting of 20% protein, 75.5% carbohydrate and 4.5% fat. All animals

had free access to drinking water. The genetic background of the TG and WT mice was

estimated to be ~99% C57BL6.

Experiments as described in Chapters III, IV and V were conducted on mice aged 15-

20, 35-40 and 50-60 weeks. Longevity data were also collected for an additional group

of TG and WT mice over a period of 94 weeks (Chapter III). To study possible cardiac

remodeling caused by variable cardiac Ang II production levels, the creation of a small

colony of homozygous TG1306/1R mice harboring double the transgene complement

was also attempted.

1.3. The GLUT4 transporter knock-out mouse (GLUT4-KO)

Generation and characteristics of the GLUT4-KO mouse are reported in the

Introduction of Chapter III. Experimental procedures were performed on GLUT4-

Lox+/+Cre+/- ‘knock-out’ (LLC) mice and their genetic controls GLUT4-Lox+/+Cre-/-

(LL). The colony was maintained by breeding LLC with LL mice and identifying the

Cre-positive mice by RT-PCR. Primers to detect transgenic Cre-positive mice were as

follow: 5’-AGCCCGGAGTAGCAGTTGTAGC-3’ (forward) and 5’-

ATGTCCATCAGGTTCTTGCG-3’ (reverse) (product size: ~350 bp). Mice were

housed in The University of Melbourne Department of Medicine, Royal Melbourne

Hospital. Animals were fed standard mouse chow consisting of 20% protein, 75.5%

carbohydrate and 4.5% fat. All animals had free access to drinking water. The genetic

background of the LL and LLC mice was estimated to be ~60% C57BL6, ~35% 129Sv

and ~5% CBA.

Experiments as described in Chapters III, IV and V were conducted on mice aged 15-

20, 35-40 and 50-60 weeks.

Chapter II

75

2. CARDIAC WEIGHT INDEX

Mice were anaesthetized by intraperitoneal injection of pentobarbitone sodium

(Nembutal, 70mg/kg). Following excision and submersion in an ice-cold basic

physiological buffer (142mM NaCl, 5.7mM KCl, 1.5mM KH2PO4.H2O, 1.7mM

MgCl2.6H2O), hearts from 15-20, 35-40 and 50-60 week old mice were trimmed of

adipose tissue, blotted, and weighed. The degree of cardiac hypertrophy was

established by measuring the ratio between the blotted wet heart weight and the body

weight (mg/g), i.e. the cardiac weight index (CWI).

Chapter II

76

3. WHOLE HEART HISTOLOGY AND MORPHOMETRY

3.1. Heart preparation and staining

Hearts from 15-20 and 35-40 week old mice were fixed in 10% buffered formalin

(Sigma) at room temperature for 24 hours, dehydrated and embedded in paraffin blocks

at 60 oC. Transversely oriented sections (5 μm) were cut at the mid point of the heart

apex-to-base aspect ratio and mounted on glass slides (Menzel Superfrost), two

sections per slide. Two slides per heart were stained with a modified version of Van

Gieson’s stain to evaluate interstitial collagen content (Figure II-1). Two other slides

were stained with Haematoxylin-Eosin staining for macroscopic evaluation of cardiac

structure and analysis of the number of cell nuclei per mm2 of tissue (Figure II-2).

3.2. Histological imaging

For estimates of collagen content, five adjacent images per ventricle and per section

were captured electronically using a high resolution CCD ‘Spot’ camera coupled to a

standard light microscope with a 10x power bright field objective. Myocardial fibrosis

was semi-quantitatively assessed in the interstitial space of the left and right ventricles

by computer-assisted morphometry. Regions of interest were selected to exclude areas

of uneven illumination or staining. Color images were converted into gray scale and

thresholded to produce binary images with collagen staining areas depicted as white,

and background as black (Figure II-1). Binary images were densitometrically analyzed

using the software ImagePro Plus 4.0. For each heart and each ventricular wall, images

were summed and results expressed as the percentage of the total cross sectional area

covered by red-stained interstitial collagen over the total sectional area analyzed (%tsa).

Areas of perivascular collagen and large blood vessels were excluded from analysis.

This analysis area was estimated at 6-8 mm2 of cardiac tissue per heart. All %tsa counts

Chapter II

77

were pooled for each experimental group and the mean %tsa calculated. For the

evaluation of the number of cell nuclei per mm2 of tissue, similar procedures were

performed on Hematoxylin-Eosin stained sections (Figure II-2). For this analysis, the

number of purple-stained nuclei over the total sectional area was computed and results

expressed as the number of cell nuclei per mm2 of myocardial tissue. Finally, for

macroscopic evaluation of cardiac structure, 1x power bright field images stained with

Hematoxylin-Eosin were captured and evaluated geometrically for myocardial

hypertrophy and ventricular chamber dilation. Chamber morphology was evaluated by

measurement of left ventricle (LV) lumenal area (mm2), while LV free wall thickness

(mm) was digitally measured from wall transects positioned to avoid papillary muscle

protrusions (average of 5 measurements per heart).

Chapter II

78

4. ADULT CARDIOMYOCYTE CELL ISOLATION

Isolated ventricular cardiomyocytes from 15-20 and 35-40 week old mice were

prepared by collagenase dissociation procedures and cell dimensions (length and width)

measured using bright field microscopy. Hearts were excised and rinsed in calcium-free

ice-cold HEPES buffer (142mM NaCl, 5.7mM KCl, 1.5mM KH2PO4.H2O, 1.7mM

MgCl2.6H2O, 10mM HEPES free-acid, 11.7mM Glucose, 20mM Taurine, pH 7.4,

100% O2 gassed). All chemicals were purchased either from Sigma or Fluka (reagent

grade).

Isolated hearts were submerged in ice-cold HEPES buffer, cannulated with the aid of a

stereomicroscope and perfused retrogradely via the aorta for 2-3 minutes with the same

HEPES buffer supplemented with 0.5 mM EGTA at 36oC (Figure II-3). The time-delay

between cardiectomy and cannulation was under 3 minutes. After a second period of 6-

8 minutes of retrograde collagenase perfusion at 36oC standard HEPES buffer

supplemented with 1mg/ml of BSA fraction V (Sigma); 0.4mg/ml collagenase type II

and 0.1 mg/ml collagenase type IV (Worthington), the left ventricular wall was minced

and the cells dissociated by shaking in HEPES buffer at 36oC. Myocytes were filtered

on a nylon mesh filter (250 μm pore opening) and resuspended in fresh HEPES buffer

(36oC). Cell dimension measurements (length and width) and cell yield were evaluated

using an eyepiece graticule fitted to the ocular tube of an inverted microscope.

Dimensions of 100 healthy rod-shaped adult cardiomyocytes were measured for each

heart. Viable cell yield (i.e. rod-shaped cardiomyocytes) ranged between 50-85% of the

total cell number, with the TG and LLC yield usually at the lower end of this range

(typically 700,000-1,000,000 cardiomyocytes). To select calcium tolerant

cardiomyocytes for single cell contractility experiments (Section 5.1), the calcium

concentration of the buffered solution was progressively increased to 25, 50 and 300

μM.

Chapter II

79

5. ADULT CARDIOMYOCYTE CONTRACTILITY

5.1. Experimental conditions

Single cardiomyocyte contractility was evaluated using a rapid imaging system, as

previously described (Harris et al., 1987; Delbridge et al., 1989). Isolated adult mouse

cardiomyocytes were allowed to settle and lightly adhere to the glass base of a trans-

illuminated temperature- and flow-controlled superfusion chamber under an inverted

Leitz Diavert trinocular microscope. Cells were superfused with pre-oxygenated,

buffered solution containing 2mM of Ca2+ (142mM NaCl, 5.7mM KCl, 1.5mM

KH2PO4.H2O, 1.7mM MgCl2.6H2O, 10mM HEPES free-acid, 2mM CaCl, 11.7mM

Glucose, 20mM Taurine, pH 7.4, 100% O2 gassed). The chamber temperature was

maintained at 36.0±0.5oC and the superfusion flow at 1.80±0.01 ml/min.

Cardiomyocytes were stimulated to contract at different frequencies (1.5 to 5Hz) by

platinum field electrodes delivering 1 ms pulses at 30% supra-threshold. Measurement

of single cardiomyocyte contractility was undertaken using a rapid scanning camera

(Fairchild CCD 1200R) attached to the upper ocular of the microscope. The camera

incorporated a scanning array line composed of 512 photodiodes, each 13x13 μm. The

scanning line was positioned along the longitudinal axis of the cardiomyocyte to track

temporal fluctuation of cell boundaries during the contraction cycle (Figure II-4). The

length of isotonically contracting cardiomyocytes was recorded at a high temporal

resolution, every 1.088 ms for up to 557 ms per contraction cycle, with scanning

synchronized to begin with electrical stimulus delivery. The spatial resolution of the

tracking system reached 2.46 pixels/μm with a 32x objective.

Chapter II

80

5.2. Recording and analysis

The scanning device was coupled to a Windows NT workstation operating the custom-

made software ‘Cardiac2000’, originally programmed in Pascal for a UNIX-based

platform (Delbridge et al., 1989), but rewritten in Visual C language to suit a more

user-friendly Windows-based environment. For each contraction cycle ‘Cardiac2000’

automatically computed a range of normalized parameters for comparative evaluation

of contractile status and inotropic response: maximum cell shortening (%S), expressed

as a percentage of initial resting cell length (Lo); the time at which the cell commenced

shortening after stimulus application, i.e. the excitation-contraction coupling latency

(To); the time at %S (Tm); the time at which the cell length returned to Lo (Tf); the

maximal rate of cell shortening (MRS) and lengthening (MRL). The detection of an

alteration in cell length of greater than 0.005 of resting length was identified as onset of

shortening (at start of cycle) or return to resting length (at cycle completion). MRS and

MRL were determined by step differentiation in increments of one scan (i.e. 1.088 ms)

with a selected step size of four scans (i.e. 4.352 ms). Cell shortening and lengthening

data stored during recording were converted from pixel to μm values, scaled and

calibrated to produce a graphical representation of a ‘cell shortening profile’, as

illustrated in Figure II-4. Some measurements reported in Chapter IV were derived

from the above-mentioned parameters: Tf-To indicates duration of the whole contractile

cycle; Tm-To, duration of cell shortening; Tf -Tm, duration of cell lengthening.

Compared to other commercially available edge detection methods with lower time

resolution (~4 ms), the present system combined high time resolution analysis (1 ms)

with automation of parameter calculation of shortening behavior (‘Cardiac2000’). This

enabled the tracking of fine and dynamic alterations in critical parameters of E-C

coupling, especially under higher (and more physiologic) ranges of stimulus frequency

(>2-3 Hz). The high temporal resolution of the system, also allowed precision

recording of the E-C coupling latency (To), which is the period of time following

stimulation during which a cardiomyocyte transduces the electrical signal

into observable mechanical behavior.

Chapter II

81

6. PROTEIN EXTRACTION AND WESTERN BLOTTING

Hearts from 15-20 and 35-40 week old mice were homogenized in 20mM Tris, 1mM

EDTA, 0.25M Sucrose (pH 7.4) and protease inhibitors (Leupeptin and PMSF).

Homogenates were spun at 3,400 rpm for 10 minutes at 4oC and crude

extracts/supernatants decanted. Pellets were resuspended in the same Tris buffer and

spun twice at 8,500 rpm for 10 minutes. Crude extracts for each sample were combined

and spun at 160,000 g for 1h in a Beckman ultracentrifuge. Pellets containing the total

membrane-bound proteins were resuspended in Tris buffer and stored at -80oC pending

analysis. After quantification of protein concentration (BioRad Protein Assay Kit),

60 μg of protein was loaded and separated in 10% SDS-polyacrylamide gels and

transferred to Polyvinylidene Difluoride (PVDF) membranes (Sigma) by a semi-dry

transfer system (BioRad). Blots were blocked in TBS buffer and 5% non-fat milk to

prevent non-specific protein-binding sites on the blot. Total protein transfer efficacy

was visualized with Ponceau S stain (Salinovich and Montelaro, 1986). For

immunodetection of the protein of interest, membranes were incubated with diluted

primary polyclonal antibodies overnight (4oC). After washing, blots were incubated

with peroxidase-conjugated (HRP) secondary antibodies and bands were revealed using

a specific chemiluminescence detection system (ECL from SantaCruz Biotechnology)

on autoradiography film (Kodak Biomax MR Film). Bands were semi-quantitatively

analyzed by densitometry on a transilluminator attached to a CCD camera (MCID) and

a Windows-based workstation. Values were expressed as protein levels detected and

normalized by total protein transferred on PVDF membrane. Details of primary and

secondary antibodies used during the experiments are reported in Chapter III and IV.

Chapter II

82

7. TOTAL RNA EXTRACTION AND RT-PCR

7.1. RNA extraction

Expression of mRNAs of interest was assessed by semi-quantitative RT-PCR of

extracted hearts from 15-20, 35-40 or 50-60 week old mice. Total RNA was purified

from heart tissue by strong denaturants and phenol/chloroform purification steps. The

first step involved homogenizing ventricular mouse tissue samples using a Polytron

Homogenizer (Kinematica AG) in a buffer containing 3 M guanidine thiocyanate, 0.7

M trisodium citrate and 10% lauryl sarcosyl. Total RNA was purified by sequential

addition and mixing of 3M sodium acetate, citrate-saturated phenol (pH 4.3) and

chloroform. Steps were repeated twice on supernatant collected after spinning (14,000

rpm). Total RNA was precipitated in isopropanol, washed in 80% ethanol, re-

suspended in sterile diethylpyrocarbonate (DEPC)-treated water and allowed to

dissolve overnight at 4°C. Samples were DNase I treated to remove DNA contaminants

and subsequently analyzed by spectrophotometry, where measurements of absorbance

(A260 and A280 nm) were recorded. From these measurements, concentration (A260 nm)

and quality (A260/A280) of each RNA sample was estimated. As a further check of RNA

quality, each RNA sample was electrophoresed (3-5 μg RNA) under denaturing

conditions (1% agar gel in tris acetate EDTA buffer (1xTAE)). Results were visualized

for integrity of 21S and 18S rRNA bands using a UV transilluminator and CCD camera

(MCID).

7.2. Optimization of RT-PCR conditions

Total RNA was reverse transcribed (RT) into cDNA and amplified in a duplex PCR

reaction using specifically designed primers. Co-amplification of the mouse GAPDH

was used as internal ratiometric control. Primers were designed using nucleotide

Chapter II

83

sequences for the genes of interest stored in the data bank of the National Center of

Bio-Informatics NCBI (http://www.ncbi.nlm.nih.gov/Tools/index.html). Primers were

designed using the on-line software Primer 3 (http://www-genome.wi.mit.edu/cgi-

bin/primer/primer3_www.cgi). Primer sequences successfully optimized were as

follows:

sodium/calcium exchanger (NCX1.1; product size: 528 bp)

5’-TTTGAGGACACCTGTGGAGTGC-3’ (forward)

5’-ATCATCGTCACG TTCCCCAGCG-3’ (reverse)

sodium/hydrogen exchanger (NHE-1; product size: 216 bp)

5’-GTACTTCCTGAAGATGTGGAGC-3’ (forward)

5’-ATGATGAACTGG TCCTTGGGGG-3’ (reverse)

ryanodine receptor (RyR2; product size: 635 bp)

5’-GAATCAGTGAGTTACTGGGCATGG-3’ (forward)

5’-CTGGTCTCTGAG TTCTCCAAAAGC-3’ (reverse)

angiotensinogen precursor (Agt; product size: 270 bp)

5’-TATCCACTGACCCAGTTCTTGC-3’ (forward)

5’-GAGAAGTTGTTC TGGGCGTCAC-3’ (reverse)

connexin 43 (Cx43; product size: 220)

5’-GTTCAAGTATGGGATTGAAGAACACGGCAA-3’ (forward)

5’-TGGTTTTCTCCGTGGGACGTGAGAGGAAGC-3’ (reverse)

GAPDH (product size: 308 bp)

5’-AAGCCCATCACCATCTTCCAGGAG-3’ (forward)

5’-AGCCCTTCCACA ATGCCAAAG-3’ (reverse)

All primers were designed to amplify mouse- and rat-specific cDNAs.

For each gene of interest, an optimization of PCR cycling conditions was performed

involving 1) the amplification of all loci individually (single locus PCR) and in

combination with the GAPDH (duplex PCR); 2) the determination of the right

proportion of primer concentration for the gene of interest and GAPDH (typical ranges

tested were 5:1, 1:1 and 1:5 of gene of interest vs. GAPDH primers); 3) the

determination of annealing time and temperature; and 4) the adjustment of the number

of PCR cycles for each primer separately, and in duplex with GAPDH (Henegariu et

Chapter II

84

al., 1997). The right PCR cycle number was determined by a semi-quantitative analysis

of the kinetics of the reaction efficiency, i.e. the amount of cDNA produced after an

established number of cycles, and the determination of the amplification phase

(Freeman et al., 1999) (Figure II-5). PCR products were run on 1% agar gel in tris

borate EDTA buffer (0.5xTBE buffer) supplemented with ethidium bromide. Bands

were visualized using a UV transilluminator and captured using a CCD camera

(MCID). Bands were semi-quantitatively analyzed by densitometry and values

expressed as mRNA expression of the gene of interest, normalized to the GAPDH

mRNA band density (Figure II-5).

Chapter II

85

8. cDNA MICROARRAY ASSAYS

8.1. Principles of DNA microarray analysis

DNA microarray analysis has emerged in the last 3-4 years as a flexible new method

for analyzing large numbers of nucleic acid fragments in parallel. The convergence of

ideas and principles utilized in the fields of hybridization methods, fluorescence

microscopy and diagnostic assays, together with advancement in bioinformatics and

miniaturization of technology have all contributed to the emergence of DNA

microarray and microchip technologies (see Introduction of Chapter V).

A variety of technical solutions that have been developed for performing microarray

analysis, but essentially all implementations involve miniaturized hybridization assays

for studying thousands of nucleic acid fragments simultaneously. All microarray

systems share the following key components: 1) the array, which contains immobilized

nucleic acid sequences, or ‘probes’ (as oligonucleotides or cDNAs) and 2) one or more

labeled samples or ‘targets’ (usually mRNA populations converted into cDNA), which

are hybridized with the microarray (Figure II-6).

8.2. RNA and cDNA preparation for hybridization

Total RNA was extracted from 60 week old mouse ventricles (left and right ventricles

including the septum) using the procedures detailed above (Chapter II, Section 7.1).

The CyScribe cDNA Post-Labelling Kit (Amersham Pharmacia) was utilized for the

preparation of Cy3- and Cy5-labelled cDNAs for microarray hybridization according to

the manufacturer’s recommended protocol. A total of 60 μg total RNA per dye labeling

reaction was used for each ‘sample versus sample’ comparison, giving a total of 120 μg

Chapter II

86

total RNA per array (2 labeling reactions per array) (Figure II-7). RNA was reverse

transcribed (RT) into cDNA using oligo(dT) to avoid transcription of non-messenger

RNA populations (i.e. ribosomal, transfer and small nuclear RNAs). During the RT

step, amino allyl-dUTPs (AA-dUTPs) were incorporated. After removal of RNA

template and purification of the amine-modified cDNA, chemical labeling with NHS-

ester derivative of CyDye was performed resulting in the generation of fluorescently

labeled cDNAs. Removal of any unincorporated CyDye molecules was necessary in

order to minimize the non-specific slide hybridization background signal and to

improve the sensitivity of detection of low abundance CyDye-labeled cDNA species.

8.3. Pre-hybridization and hybridization of cDNA microarray slides

Microarrays consist of a collection of single-stranded nucleic acid sequences

immobilized onto a solid support so that each unique sequence forms a probe, or ‘spot’.

A glass slide acts as the solid support onto which up to tens of thousands of spots can

be arrayed in a total area of a few square centimeters.

In the present study, microarray slides were produced at the Australian Genome

Research Facility in Melbourne (Australia) by robotic printing

(http://www.agrf.org.au/). PCR amplified cDNA clones were arrayed at high density

onto Corning CMT-GAPS aminosilane-coated glass slides. The specific clone set

printed onto the slides was the NIA 15K mouse embryonic clone set (National Institute

of Aging, NIH - USA- http://lgsun.grc.nia.nih.gov/cDNA/15k.html). Each PCR-

amplified cDNA clone represented a spot on the slide and 15,247 unique clones were

spotted in duplicate on each slide. For post-experimental normalization of data, each

microarray slide also possessed a subset of control spots including positive controls,

intensity controls, ‘housekeeping’ genes, blocking controls and negative controls.

Immediately prior to hybridization, cDNA microarray slides were incubated in pre-

hybridization buffer containing BSA-fraction V (Sigma), washed and dried for

hybridization. This process ‘blocked’ or inactivated the free amine groups present on

Chapter II

87

the glass slides, preventing possible non-specific binding of Cy3- and Cy5-labelled

cDNA to the glass slide.

For each slide, a hybridization mixture was prepared (volume 60 μl) containing 0.6

mg/ml poly-A (Amersham Pharmacia), 0.8 mg/ml Salmon Sperm DNA (Gibco BRL),

0.4 μg/μl mouse Cot1 DNA (Gibco BRL) and two purified Cy5- and Cy3-labeled

cDNA samples to be compared. A 60x25 mm lifter-slip (Grale Scientific) was placed

over the marked array region on the pre-hybridized microarray slide and then the entire

hybridization mixture was pipetted underneath the slip and allowed to move by

capillary action across the surface of the array. The slide was then placed horizontally

in hybridization chambers (Corning), sealed and placed at the bottom of a 42 oC water

bath where they incubated for 16-20 hours in the dark. During the hybridization step,

the labeled fragments in the target form duplexes with their immobilized

complementary probes. The number of duplexes formed reflects the relative number of

each specific fragment in the target, as long as the amount of immobilized probe is in

excess and not limiting the kinetics of hybridization (Figure II-8).

8.4. Slide washing and scanning

Following incubation, the array slides were removed from chambers, rinsed and spun

dry in a plate centrifuge. The microarray slides were placed in slide mailers and kept in

a dark, desiccated environment prior to scanning. Each slide was imaged by a GenePix

4000B (Axon Instrument) laser scanner which acquired two gray-scaled 16-bit single

image TIFF files corresponding to the Cy5 and Cy3 fluorescence emission channels

(argon laser operating at ~650 nm and ~550 nm excitation wavelengths respectively).

The gray-scaled images were ‘pseudo colored’ (red intensities for Cy5, green

intensities for Cy3) and merged to give one single picture for dual color (red/green)

differential microarray analysis. By measuring the different fluorescent signals

associated with each spot, the relative abundance of specific sequence in each of the

samples can be determined. When equal dual color signal was obtained from a spot, the

spot appeared yellow, suggesting equal cDNA abundance for that clone in the 2

Chapter II

88

samples. Shades and tonalities of red and green denoted differences in relative

abundance in favor of one or the other sample (Figure II-8).

8.5. Statistical normalization of cDNA microarray data and

clone ranking

Microarrays generate large quantities of data even from a single experiment. As a

typical experiment will involve the use of several analyzed samples on replicate arrays,

the use of computerized data processing and bioinformatics is necessary in order to

handle the amount of data generated and to gain maximum statistical and biological

information from the experiment. This can be achieved by specialized software and

statistical algorithms that extract primary data to remove the influence of experimental

variation and artifacts, and manipulate the data so that biologically meaningful

conclusion can be made (Quackenbush, 2001; Hoffmann et al., 2002; Zapala et al.,

2002).

In the present study, the microarray analysis software packages ‘Spot’ (CSIRO

Mathematical and Information Sciences, Australia) and ‘GenePix’ (Axon Instruments,

USA) were used to analyze the 16-bit TIFF files. Spot and GenePix extracted

foreground and background intensities for individual spots on a microarray image,

producing a matrix containing Cy5 (red) and Cy3 (green) intensities for each spot on

the microarray.

An M-value was calculated to quantify differential expression between two RNA

samples by representing a log ratio of the hybridization intensities for both Cy dyes:

M = log2 (R/G) = log2 R - log2 G

Chapter II

89

where R and G were the background-corrected red (Cy5) and green (Cy3) intensities

for each spot respectively. On this scale, M=0 represents equal expression, M=1 a two-

fold differential change (200%) between RNA samples, and M=2 represents a four-fold

change (400%), etc.

The log-intensity (A) of each spot was determined:

A= (log2 R + log2 G)/2 = 0.5 log2 (R*G)

as a measurement of overall brightness of the spot. An ‘MA-plot’, which provides a

scatter plot of the M-values against the A-values for an array, was used to represent the

data and to detect any non-linear relationship between log intensities due to artifacts

present on the array or dye-specific bias (generally, green-bias at low intensity and red-

bias at high intensity) (Dudoit and Fridlyand., 2002) (Figure II-9).

Following print-tip group loess (local weighted regression) normalization (Figure II-10)

and exclusion of spots presenting visual artifacts and negative values for R or G, the

results were transferred to a spreadsheet (Microsoft Excel) and matched to the Gene

Array List file (GAL file - supplied with the purchase of the microarray slide), which

assigned to each spot a clone name according to the geographical location of each spot

on the microarray slide (Smyth et al., 2003).

Clones were ranked in order of magnitude of differential expression (i.e. normalized M-

values), for positive and negative values.

To compare results obtained from different microarray replicates, clones were further

ranked following a critical-value for the ranking statistic above which any value was

considered to be significant or ‘true’. That is, clones were ranked according to their

variation in M value between experiments (replicates) using a parametric empirical

Chapter II

90

Bayes approach (Lönnstedt and Britton, 2004). The latter used a B-statistic to estimate

differential expression:

B = Me / √[(a + s2) / n]

where Me was the mean of the M-values for any particular clone across a series of

replicate arrays; s2 was the variance of the M-values across the replicates for the clone

in question; a was a constant estimated from the mean and standard deviation of s2; n

was the number of replicate arrays. Values of B-statistic greater that zero corresponded

to a greater than 50-50 chance that the M value found for the clone was truly

representing a pattern of differential expression. The higher the B value, the greater the

chance that the result is significant.

Candidate clones that generated a B-value greater than 0, in addition to a M value

greater that 1 (2-fold upregulation) or lower than -1 (2-fold downregulation), were

verified for their sequence at the NIA web site

(http://lgsun.grc.nia.nih.gov/cDNA/PublicSeqVerify.html) and a gene name (if any)

assigned according to the best match in sequence alignment

(http://lgsun.grc.nia.nih.gov/15k/).

8.6. Gene Ontology classification of selected transcript candidates

Candidates that presented significant patterns of differential expression (M>1 or M<-1

AND B>0) were clustered according to the Gene Ontology (GO) Consortium

classification groups (http://www.geneontology.org/index.shtml): The goal of the GO

Consortium is to produce a controlled vocabulary that can be applied to all organisms,

even as knowledge of gene and protein roles in cells is accumulating and changing. GO

provides three structured networks of defined terms (molecular function, biological

process and cellular component) to describe gene product attributes. Clustering was

performed according to these three networks using the web-based software GFINDer

Chapter II

91

(http://www.medinfopoli.polimi.it/GFINDer/). GFINDer organizes lists of 'candidate'

genes (e.g., up- and downregulated genes from a microarray experiment) for biological

interpretation in the context of the Gene Ontology, Pathways, Protein Domains,

Diseases and Chromosomes (Masseroli et al., 2004). Statistical analysis and

classification of candidate genes is explained on the GFINDer website (under ‘Tutorial

Section’) and in Masseroli et al. (2004).

Chapter II

92

9. STATISTICS

With the exception of microarray data, all results are expressed as mean ± SEM. One-

way analysis of variance (ANOVA) was performed to evaluate single factor differences

between two experimental groups (usually the animal age or the genotype). Multiple-

way ANOVA was used to evaluate age and genotype differences between

experimental groups and to assess interactions of a single dependent continuous

variable with multiple independent nominal variables (that is, interactions between age,

genotype, frequency, etc.). Significant differences were identified at p<0.05. Legends

of Figures and Tables reporting specific data and results state the type of statistical test

utilized and the degree of significance.

Chapter II

93

100μm100μm

White surfaceWhite surface + Black surface

White surfaceWhite surface + Black surfaceΣ(20)

20x LV20x RV

= %tsa for 1 ventricle%

A.

B.

C.

Figure II-1: Histological analysis of myocardial collagen density

Sections stained using a modified Van Gieson’s method were viewed at 10x magnification with a bright field light microscope (A.). Ten segments per section were selected (5 for the LV, 5 for the RV) and four sections analyzed. Color images were converted to gray scale (B.) and thresholded, to create a binary mask (C.) which was densitometrically analyzed (total white pixels over the total surface area). Collagen was ‘white’ delineated against the black background. A totality of 40 fields were analyzed per mouse heart (20 for LV and 20 for RV). The summation of the analysis of the 20 segments per ventricle and per heart corresponded to one single measurement (%tsa). Measurements were averaged per ventricle and for all animals in each group. LV= left ventricle; RV= right ventricle.

Chapter II

94

100μm

Number of white areasWhite surface + Black surface

Number of white areasWhite surface + Black surfaceΣ(20)

20x LV20x RV

= Nuclei/mm2 for 1 ventriclemm2

A.

B.

C.

D.

Figure II-2: Histological analysis of myocardial nuclei count

Transverse heart sections stained using Hematoxylin-Eosin method (A.) were viewed at 10x magnification with a bright field light microscope (B.). Ten segments per section were selected (5 for the LV, 5 for the RV) and four sections analyzed. Color images were converted to gray scale (C.) and thresholded, to create a binary mask (D.) which was densitometrically analyzed (total number of white particles with a surface area > 10 pixels, over the total surface area). Nuclei were ‘white’ delineated against the ‘black’ background. A total of 40 fields were analyzed per mouse heart (20 for LV and 20 for RV). The summation of the analysis of the 20 segments per ventricle and per heart corresponded to one single measurement (number of nuclei per mm2, for a total field of 6-8 mm2). Measurements were averaged per ventricle and for all animals in each group. LV= left ventricle; RV= right ventricle.

Chapter II

95

a.

b.c.

d.e.

f.

A.

B.

a.

b.c.

d.e.

f.

A.

B.

Figure II-3: Depiction of cardiac perfusion and myocyte isolation apparatus

A. Schematic diagram of the mouse cell isolation superfusion system: a. Pre-heated (36 oC) water bath where HEPES and collagenase solutions are maintained at warm temperatures; b. Thermostat allowing adjustment of temperatures and re-circulation of pre-heated water; c. A connector allows the ‘switching’ between HEPES and collagenase solutions at any time; d. Peristaltic pump allowing the perfusion of the heart at a constant flow (1.8 ml/min); e. The perfusion is kept at warm temperature and reaches the heart at 36 oC; f. A pre-heated water jacket keeps the cannulated heart in a warm environment during the collagenase digestion phase. B. Details of a cannulated mouse heart (yellow arrow), mounted on perfusion apparatus.

Chapter II

96

‘Shortening’

Time (ms)

A.

100

90

B.

To Tm Tf

mrlmrs

ms

% Lo

%S

Cell le

ngth

(L o)

‘Shortening’ ‘Lengthening’

‘Shortening’

Time (ms)

A.

100

90

B.

To Tm Tf

mrlmrs

ms

% Lo

%S

Cell le

ngth

(L o)

‘Shortening’ ‘Lengthening’

Figure II-4: Depiction of cardiomyocyte contractile function evaluation

A. The photo sensor array, which was positioned along the cell length (red bar), detected cell boundaries and tracked cell shortening and lengthening during the contractile cycle. B. Data stored during recording were converted, scaled and calibrated to produce the graphical representation of a ‘cell shortening profile’. Parameters calculated included the maximum cell shortening (%S), expressed as a percentage of initial resting cell length (Lo); the time at which the cell commenced shortening after stimulus application, i.e. the excitation-contraction coupling latency (To); the time at %S (Tm); the time at which the cell length returned to Lo (Tf); the maximal rate of cell shortening (MRS) and lengthening (MRL). Tf -To indicates period of shortening and lengthening; Tm-To, period of shortening; Tf -Tm, period of lengthening.

Chapter II

97

0

200

400

600

800

15 18 21 24 27 30 33 36 39 42 45

AognGAPDH

GAPDH

NCX1.1

(Mean pixel intensity x pixel number) – (Mean pixel intensity x pixel number)

(Mean pixel intensity x pixel number) – (Mean pixel intensity x pixel number)

Background

B.

A.

Amplification phase

No cycles

Expr

essi

on le

vels

(arb

. uni

ts)

ROI

0

200

400

600

800

15 18 21 24 27 30 33 36 39 42 45

AognGAPDH

GAPDH

NCX1.1

(Mean pixel intensity x pixel number) – (Mean pixel intensity x pixel number)

(Mean pixel intensity x pixel number) – (Mean pixel intensity x pixel number)

Background

B.

A.

Amplification phase

No cycles

Expr

essi

on le

vels

(arb

. uni

ts)

ROI

NCX1.1GAPDH

Figure II-5: Illustration of RT-PCR amplification and gel analysis

A. Gel bands were densitometrically analyzed using the ‘Histogram’ Function of Photoshop (7.0). Product yield for the gene of interest was determined in arbitrary units by subtracting the background light signal from the band light signal. The light signal was the product of the mean pixel intensity and pixel number in the region of interest (ROI). The product yield for the gene of interest (here, the angiotensinogen gene Aogn) was normalized for GAPDH by computing the ratio of the values determined for each of the two bands of the duplex PCR.; B. Successful application of semi-quantitative PCR methodology to amplification of genes of interest required optimization of temperatures, magnesium concentrations, primer concentrations and number of amplification cycles. Optimal PCR cycle number was determined by a semi-quantitative analysis of the kinetic of the reaction efficiency, i.e. the amount of cDNA produced after an established number of cycles, and the determination of the exponential amplification phase (Freeman et al., 1999). In the example, after the determination of the appropriate ratio of primer concentration for the two genes in the PCR reaction (here, NCX1.1:GAPDH= 5:1), the two genes were amplified at different number of cycles, each point corresponding to the average of five reaction tubes. For this particular gene, a number of 28 cycles was finally chosen for the experiments. All points of the graphic were obtained during one single PCR experiment and for one single RNA pool made of three different mouse hearts.

Chapter II

98

Target

Probe

Hybrid.

Glass slide

Target

Probe

Hybrid.

Glass slide

Figure II-6: Illustration of microarray ‘target’ and ‘probe’ species

In the literature, there currently exist at least two nomenclature systems for referring to hybridization partners. Both use common terms ‘probes’ and ‘targets’. With symmetry akin to the hybridization reaction itself, each system mirrors the other. The strategy of the ‘standard’ microarray parallels that of a reverse dot-blot, in which the probe is immobilized on membrane. For this reason the tethered nucleic acid is described here as ‘probe’, and the free nucleic acid as ‘target’.

Chapter II

99

Total RNA Samples A (60μg) and B (60μg) (2 tubes) = 1 array

1. Annealing using oligo(dT)2. Elongation (and incorporation of AA-dUTP)3. Degradation of RNA template (alcaline treatment)

Amine-modified cDNAs A and B (2 tubes)

1. Purification2. Clean-up3. Concentration (Microcon 30)

Purified Amine-modified cDNAs A and B (2 tubes)

Over-night coupling with CyDye and Hydroxylamine step

Cy5-labelled cDNA A and Cy3-labelled cDNA B (2 tubes)

1. Purification2. Clean-up (of free unincorporated Cy5 and Cy3)3. Concentration (Microcon 30)

Purified Cy5- and Cy3-labelled cDNAs A + B (in 1 tube)

1. Washing and drying of microarray slides2. Laser scanning and first data analysis

Day 2

Day 3

RNA extraction from heart tissue Samples

Quantification (μg/μl)

Quality check RNA gel electrophoresis

OD measurement (260/280 nm)

Day 1

Over-night hybridization with BSA pre-treated microarray slide

Statistical normalization of cDNA microarray data

‘Hunting’ for differential gene expression patterns

Total RNA Samples A (60μg) and B (60μg) (2 tubes) = 1 array

1. Annealing using oligo(dT)2. Elongation (and incorporation of AA-dUTP)3. Degradation of RNA template (alcaline treatment)

Amine-modified cDNAs A and B (2 tubes)

1. Purification2. Clean-up3. Concentration (Microcon 30)

Purified Amine-modified cDNAs A and B (2 tubes)

Over-night coupling with CyDye and Hydroxylamine step

Cy5-labelled cDNA A and Cy3-labelled cDNA B (2 tubes)

1. Purification2. Clean-up (of free unincorporated Cy5 and Cy3)3. Concentration (Microcon 30)

Purified Cy5- and Cy3-labelled cDNAs A + B (in 1 tube)

1. Washing and drying of microarray slides2. Laser scanning and first data analysis

Day 2

Day 3

RNA extraction from heart tissue Samples

Quantification (μg/μl)

Quality check RNA gel electrophoresis

OD measurement (260/280 nm)

Day 1

Over-night hybridization with BSA pre-treated microarray slide

Statistical normalization of cDNA microarray data

‘Hunting’ for differential gene expression patterns

Figure II-7: Summarized microarray hybridization protocol

Schematic diagram representing the time-line of a microarray experiment. Usually, experimental procedures were run in 3 days, while statistical normalization and ‘candidate gene hunting’ generally took several weeks to be performed.

Chapter II

100

A. B.

C.

D.

A. B.

C.

D.

Figure II-8: Microarray hybridization and visualization depiction

A. Schematic illustrating an AGRF cDNA microarray sector made of 26x26 spots (338 clones printed in duplicate). Tonalities of green (B.) or red (C.) denoted differences in relative abundance in favor of the ‘green-colored’ or the ‘red-colored’ cDNA sample, respectively. A yellow color signal (D.) was obtained from a spot where cDNA abundance for that clone was the same in the 2 samples.

Chapter II

101

Figure II-9: Depiction of a MA plot after normalization

Colored spots indicate various positive and negative controls printed on the slides. For each array a ‘Microarray Sample Pool’ (MSP) is generated and the concentration adjusted to 500 ng/μl. This is used as a positive control and also to mark the top left and the top right of each pin group. A series of standard dilutions of the MSP are used for intensity-dependent normalizations. Other controls consist of ‘housekeeping genes’, ‘blocking controls’ (polyA, Cot1 and Salmon Sperm DNA), ‘negative controls’ (yeast genes; to assess true background) and spikes controls (to evaluate dose-dependent labeling efficacy). Details about control spots used by the AGRF are reported at the web site http://www.agrf.org.au/mic_controls.html. The example presented here does not correspond to a microarray slide used during the experiments. The example illustrates where control spots and subset of genes are supposed to line up according to their expected intensity or M-value (red arrows) and where they finally are after being used to normalize the microarray experiment.

Chapter II

102

A

B

Figure II-10: Side-by-side box plots of the M-values from a microarray experiment

Loess (local weighted regression) normalization was performed according to print-tip groups, which corresponded to 48 sectors of 26x26 spots on the AGRF microarray slides (here only 32 sectors are illustrated). Box plots can be useful for comparing M-values between various groups. They display graphically the so-called 5-number summary of a set of numbers, the three quartiles and the maximum and the minimum. The central box of the plot extends from the first to the third quartile and therefore encompasses the middle 50% the data. Each box plot represents a sector on the slide. A. The graphical illustration after normalization. B. The same graphical illustration before normalization. In this case, the mean intensity was normalized from a red-green bias due to differences between the labeling efficiencies and scanning properties of the Cy3 and Cy5 dyes. The lower red intensity shifted the M-values to negative values.

CHAPTER III

Diverse evolving cardiac and cardiomyocyte

phenotypes in Ang II-induced and insulin resistant

cardiac hypertrophy

Chapter III

104

1. INTRODUCTION

1.1. Cardiac adaptation in response to environmental changes

Cardiac remodeling can be defined as a physiologic and/or a pathologic response where

the heart dynamically changes in size, shape and function in response to a variety of

extra-cardiac and intra-cardiac stimuli (described in Chapter I). It is currently a matter

of debate whether remodeling in response to pathologic signaling is detrimental from

the outset or, whether it may be initially beneficial and leads to cardiac failure only

when prolonged or exacerbated. Notwithstanding, the degree of remodeling is

determined by various structural and genetic factors, including:

1. Collagen concentration or collagen isoform expression.

2. The balance between cell growth/proliferation and cell death, which would alter

the proportion of myocyte to non-myocyte cell types in the myocardium.

3. Differential gene expression of proteins involved in cell communication, cell

activation and metabolism.

1.2. In vivo models of Ang II-induced cardiac remodeling

1.2.1. Cardiac remodeling regression by Ang II suppression

As reviewed in Chapter I, the hypertrophic and pro-fibrotic actions of Ang II

significantly influence the degree of response of the myocardium in the remodeling

process. However, there are still open questions concerning the intrinsic capacity of

Ang II to produce cardiomyocyte growth and fibrosis. Ang II codependence on

additional stimuli such as mechanical stretch or the presence of other co- factors (e.g.

Chapter III

105

endothelin-1 or TGF-β) in growth induction has been indicated in vitro and in vivo (Ito

et al., 1994; Arai et al., 1995; Campbell and Katwa, 1997; Dostal, 2000; Wenzel et al.,

2001). Animal models of left ventricular cardiac hypertrophy including the SHR, the

TGR (mRen2)27, and isoproterenol-infused or aortic banded rats, respond to blockade

of the RAS with reduction in left ventricular weight and decreased myocardial fibrosis

(Brilla et al., 1991; Pahor et al., 1991; Nagano et al., 1992; Bohm et al., 1996).

However, the majority of the beneficial effects observed on the myocardium were also

confounded by systemic reduction of plasma Ang II levels and decrease of pressure

load. Whether intra-cardiac elevation of Ang II per se (in the absence of other

coincident stimuli) is sufficient to induce myocardial growth has been identified as a

research question of considerable interest.

1.2.2. Overexpressing or knocking out the intra-cardiac RAS

Genetically engineered mice carrying cardiac-specific gain- or loss-of function

transgenic complements of components of the RAS offer a unique opportunity to

address the question of whether Ang II-specific signaling is sufficient to induce

cardiomyocyte hypertrophy and myocardial remodeling or whether stretch stimulus and

complimentary growth factors are also necessary. Different transgenic and knock-out

mouse models have been created by incorporating or deleting angiotensin or

angiotensin receptor genes in the myocardium.

1.2.3. Overexpressing the AT receptors

To elucidate whether Ang II can activate a growth response in cardiomyocytes in vivo,

Hein et al., (1997) overexpressed the AT1a receptor in a cardiomyocyte-specific

manner, under the control of the αMHC promoter. The offspring displayed a massive

atrial enlargement due to myocyte proliferation, and died within the first weeks after

birth, possibly related to bradycardia and heart block. Cardiomyocyte hyperplasia in

AT1a-overexpressing mice suggests that Ang II-mediated activation of AT1a receptors is

sufficient to induce a growth response associated with altered electrical conduction.

Interestingly, Hoffmann et al. (2001) developed a transgenic rat overexpressing the

Chapter III

106

human AT1 receptor under the αMHC promoter in the myocardium. Transgenic rats

exhibited normal cardiac growth and heart function under basal conditions, but a

pronounced cardiac hypertrophy associated with enhanced contractile response was

observed after Ang II-dependent pressure- and volume-overload. Paradis et al. (2000)

overexpressed the human AT1 receptor in transgenic mice under the control of the

αMHC. These mice were normotensive but exhibited significant cardiac hypertrophy

and fibrosis, associated with expression of ANP. Human AT1-overexpressing mice died

prematurely (after ~3 months) of heart failure. This finding partially contrasted with the

findings of Hein et al. (1997), who observed early postnatal mortality and hyperplasia

in myocyte overexpressing the murine AT1a receptor. It is important to note that

whereas the mice used by Paradis et al. (2000) had only one renin gene, the animals

used by Hein et al. (1997) were outbred from a Swiss strain, and carried two renin

genes. Because mice with two renin genes have enhanced systemic and intra-cardiac

RAS activity (Tronik et al., 1987; Clark, 1997; Mazzolai et al., 1998), the more severe

phenotype observed by Hein et al. (1997) may reflect Ang II hyperstimulation of the

already increased cardiac AT1 receptor levels. Overexpression of AT1 receptors in

mouse cardiomyocytes shows that increased receptor number is sufficient to induce and

maintain cardiac and cardiomyocyte hypertrophy, independently of changes in

hemodynamic load. Strain and species differences in the systemic expression of renin

and Agt levels, as observed in rats and in two renin gene mice, could enhance or

suppress the Ang II-dependent remodeling processes on the heart, suggesting that Ang

II production levels are also important in the degree of cardiac and cardiomyocyte

remodeling.

In a different murine model, Masaki et al. (1998) overexpressed the AT2 receptor under

the cardiac-specific αMHC promoter. These mice survived to adulthood and showed no

obvious developmental or morphological changes of the myocardium. AT2 transgenic

mice showed in vivo and ex vivo decreased sensitivity to the pressor effects of Ang II.

The authors suggested that this effect was caused by an Ang II- and AT2-mediated,

catecholamine-independent, negative chronotropic effect. Interestingly, AT1 receptor-

dependent activation of the MAPK pathways was blunted in AT2 transgenic mice,

suggesting that AT2 receptors can oppose AT1-mediated trophic responses in

cardiomyocytes in vivo. In subsequent studies, Sugino et al. (2001) showed that

Chapter III

107

cardiac-specific overexpression of AT2 receptors in vivo is not associated with

enhanced apoptosis after 28 days of Ang II infusion. Kurisu et al. (2003) demonstrated

that AT2 receptor overexpression in vivo does not lead to any suppression effect on

cardiomyocyte hypertrophy in response to Ang II-mediated pressure overload, but

attenuates myocardial perivascular fibrosis by a kinin/NO-dependent mechanism.

These data would suggest that a shift in the AT1/AT2 expression ratio by AT2

overexpression can influence perivascular collagen production and deposition in the

heart, but has limited effects on stretch-mediated cardiac remodeling.

1.2.4. Knocking out the AT receptors

Several groups have generated mice with a targeted disruption of the AT1a or AT2

receptor genes. Although gene knock-out was not induced in a cardiac-specific manner,

these mice provide some information about the functional importance of AT receptor

sub-types in the development of cardiac remodeling in the in vivo context. AT1a

receptor deficiency induced systemic hypotension but did not produce overt signs of

cardiac and cardiomyocyte hypertrophy (Sugaya et al., 1995). This indicates that AT1-

mediated signaling is necessary for the regulation of resting blood pressure in mice.

Harada et al (1998a, 1998b) demonstrated that infusion of sub-pressor doses of Ang II

induced hypertrophy and fetal gene expression in wild-type but not in AT1a knock-out

hearts, suggesting that the AT1a receptor subtype is necessary in the promotion of

cardiac hypertrophy by exogenous Ang II. In contrast, acute pressure-overload induced

by transverse aortic banding induced, in wild-type and AT1a knock-out hearts, produced

similar activation of fetal genes and MAPK kinases, in association with marked cardiac

hypertrophy, fibrosis, SERCA2 downregulation and systolic dysfunction. These data

suggest that acute hypertrophy in response to pressure overload can occur in the heart

independently of AT1a signaling pathways. Thus it could be proposed that other Ang II-

and AT1-independent mechanisms of pressure overload-induced cardiac hypertrophy

coexist in the myocardium. More recently, Harada et al. (1999) demonstrated in the

same knock-out mice that the AT1a receptor subtype plays a pivotal role in the

progression of LV remodeling after myocardial infarction. Although the response of

wild-type and knock-out mice at 1 week after infarction was comparable (with

development of LV dilatation, LV dysfunction and cardiac fibrosis in the non-infarcted

Chapter III

108

area), the survival rate and the degree of remodeling was lower in AT1a knock-out mice

than wild-type 4 weeks after infarction. These results suggest that Ang II activation of

AT1a signaling plays a negative role in the progressive development of heart failure

following infarction. Finally, Oliverio et al. (1998) generated mice lacking both the

AT1a and AT1b receptor subtypes. AT1b-deficient mice did not present any specifically

abnormal phenotype. However, mice lacking both AT1a and AT1b subtypes presented

hypotension and virtually no systemic pressor response to infusions of Ang II, thus

complementing the study from Harada et al. (1998a, 1998b). This suggests that the

AT1a, but not the AT1b receptor subtype is necessary in the promotion of cardiac

hypertrophy and vascular pressor response by exogenous Ang II.

Ichiki et al. (1995) demonstrated that AT2 receptor inactivation increased blood

pressure and enhanced sensitivity to the pressor action of Ang II, thus proposing that

AT2 receptors mediate a depressor effect and antagonize the AT1-mediated pressor

action of Ang II. Interestingly, AT2 knock-out mice show subtle behavioral

disturbances, including an impaired drinking response to water deprivation. Li et al.

(2003) showed that intracerebroventricular injection of Ang II in AT2 knock out mice

increased systolic blood pressure more markedly than in wild-type littermates,

indicating that AT2 receptors play an important role in the central regulation of

systemic blood pressure. Water intake following intracerebroventricular injection was

partly inhibited in AT2 (and AT1a) knock out mice, suggesting that both receptor

subtypes act synergistically in the regulation of water intake induced by Ang II. In

more recent work, Brede et al. (2003) demonstrated in another AT2 knock out mouse

strain that Ang II failed to induce the expression of the anti-hypertrophic endothelial

nitric oxide synthase (eNOS) in cardiomyocytes after myocardial cryoinjury,

suggesting that eNOS is expressed in cardiomyocytes via an AT2-dependent

mechanism that ultimately confers an anti-hypertrophic effect. Finally, Gross et al.

(2004) showed in the same AT2 knock out mouse strain that treatment with N-nitro-L-

arginine methyl ester (L-NAME) induced an exacerbated hypertensive phenotype,

associated with enhanced cardiac hypertrophy, fibrosis, and greater expression of

natriuretic peptides (BNP). The end-diastolic pressure-volume relationship, indicative

of LV contractility was also decreased in their AT2 knock-out mice. These data indicate

that the AT2 receptor offers a protective effect in the development of L-NAME induced

cardiac hypertrophy.

Chapter III

109

The consensus conclusion which may be drawn considering all these studies is that

cardiomyocyte AT2 receptors do not cause hypertrophy and do not independently

induce apoptosis. There is evidence however that AT2 receptors expressed by cardiac

myocytes or fibroblasts or even other cell types may be essential for hypertrophy

induction via AT1 receptor subtype.

1.2.5. Knocking out the angiotensinogen gene

Tanimoko et al. (1994) generated angiotensinogen-deficient mice which lacked Agt in

the liver and kidney, resulting in complete suppression of plasma Ang I and Ang II. As

seen with the deletion of the AT1a receptor subtype, these mice exhibited a hypotensive

phenotype, suggesting that both Ang II and AT1a levels are important modulators of the

vascular (and central) blood pressure homeostasis. In addition, Agt knock-out mice

showed defects in the development of the kidneys as well as hyperplasia of vascular

smooth muscle cells associated with altered renin sequestration from the circulation

(Nagata et al., 1996). Sumida et al. (1998) demonstrated that Agt deficient mice

exhibited an ~100% increase in the density of AT1 receptors in the heart which was not

associated with AT2 receptor expression changes. However, no cardiac or

cardiomyocyte remodeling was demonstrated in these mice. Interestingly, Kang et al.

(2002) generated a new mouse model by breeding Agt deficient mice with hypertensive

transgenic animals expressing the rat Agt gene specifically in brain and liver. Crossbred

animals lacked detectable expression of Agt in the heart and in kidneys. Brain and liver

overexpression of Agt in these crossbred animals overcame the hypotension shown in

Agt knock-out mice and even caused hypertension similar in magnitude to the Agt

overexpressing mice. As a consequence of the lack of Agt in the heart, crossbred mice

presented with less marked pressure overload-induced cardiac hypertrophy and less

pronounced perivascular and interstitial fibrosis. These experiments showed that local

Agt synthesis, independently of pressure load or circulating Ang II levels, is important

in the development of cardiac remodeling.

Chapter III

110

1.2.6. Overexpressing Ang II fusion protein directly in the heart

Van Kats et al. (2001) described various transgenic mouse lines in which

overexpression of variable levels of an Ang II-producing fusion protein was associated

with two forms of cardiac remodeling. When the fusion protein was sequestered into

the cardiac tissue, the mouse was normotensive and exhibited increased fibrosis but not

hypertrophy. When the fusion protein levels were high enough to spill into the systemic

circulation, hypertension developed inducing pressure overload cardiac hypertrophy.

Thus, the issue of whether or not Ang II alone can induce cardiac hypertrophy

independently of hemodynamic changes remained unresolved.

1.2.7. Overexpressing the angiotensinogen gene in the heart

The final genetic approach which has involved manipulation of the cardiac RAS is the

angiotensinogen (Agt)-overexpression mouse model. As this model has been employed

in the studies presented in this and subsequent Chapters, a more detailed description of

the model generation is provided here.

In order to selectively increase the production of Agt in the heart, a transgene

composed of the cardiac-specific promoter of the αMHC gene coupled to a rat Agt

cDNA was microinjected in fertilized eggs from mice, carrying either a two renin or a

one renin genotype (Tronik et al., 1987; Clark, 1997). Two transgenic lines were

obtained: the first line, the TG153/1R, carried a low copy number of the transgene

under one renin gene background. This line was normotensive and the onset of

hypertrophy was delayed (not evident until the age of 20 weeks). The second line, the

TG101/2R, carried a high number of copies of the Agt transgene and was generated in

a two renin gene strain. This second line developed hypertension and cardiac

hypertrophy, which was already significant at the age of 8-9 weeks. The main

difference between the two transgenic lines was that plasma renin activity was elevated

in TG101/2R, but not in TG153/1R mice. This was attributed to combined effects of

increased plasma renin and circulatory spillover of cardiac Agt in TG101/2R. In the

Chapter III

111

TG153/1R the normal feedback mechanism was apparently able to control renin

secretion from the kidneys.

In order to determine if hypertension was due to the high number of Agt genes or to the

two-renin gene background, the TG101/2R line was backcrossed with 1 renin gene

C57BL6 mice. At the F2 generation, heterozygote mice carrying a high number of the

transgene but only one renin gene were obtained. This line, called TG1306/1R was

normotensive, but exhibited a cardiac hypertrophy which was detectable from at least

the age of 8-9 weeks. Heterozygote TG1306/1R breeders were subsequently mated

with normal C57BL6 females to generate litters comprising of half heterozygote

transgenic and half wild-type progeny (designated TG and WT respectively in this

Thesis).

In the TG1306/1R (TG), at approximately 15 weeks, the cardiac mass increase was

estimated at 15-20% compared to littermate WT, and the development of hypertrophy

could be completely suppressed by in vivo AT1 blocker (losartan) treatment. The

cardiac levels of Ang II at 8 and 12 weeks were increased twofold in TG when

compared to WT. Interestingly, the normotensive state in this model is associated with

suppression of systemic renin secretion and circulation, thus suggesting that a negative

feedback associated with increased cardiac levels of Ang II ensures normal blood

pressure (Mazzolai et al., 1998; Mazzolai et al., 2000). These data show that the

hypertrophy developed in the TG (i.e. the TG1306/1R) is not a compensatory response

to pressure load, providing strong evidence for a role of Ang II on cardiac growth,

independent of its hemodynamic effects. The TG1306/1R mouse line offers the

opportunity to study and characterize the Ang II-induced primary hypertrophic

phenotype in the absence of hypertension (Mazzolai et al., 1998).

As published by Mazzolai et al. (1998) and Clement et al. (2001) re-expression of fetal

genes such as alpha-skeletal smooth muscle actin and ANP is evident in this transgenic

model, indicating that elevated Ang II levels are associated with a modified gene

expression profile. Moreover, blockade of the AT1 receptor induced regression of

cardiac hypertrophy, confirming that the remodeling is mediated by an AT1-receptor-

mediated signaling pathway (Mazzolai et al., 2000). Finally, Pellieux et al. (2000)

showed that in the TG mice, the cardiotrophic action of Ang II is mediated by the

Chapter III

112

interaction of Ang II with the AT1 receptor and subsequent activation of the p38

MAPK pathway.

1.3. Transgenic and knock out models for the GLUT4

1.3.1. Overexpressing or knocking out the GLUT4

GLUT4, the insulin-responsive glucose transporter, plays an important role in

postprandial glucose disposal. Altered GLUT4 activity is considered to be one of the

factors responsible for decreased glucose uptake in muscles and adipose tissue in

obesity and diabetes, where cellular insulin-resistance is present. Transgenic mice

overexpressing the human GLUT4 protein in cardiac muscle have been created (Belke

et al, 2001). As has been observed with other transgenic mice overexpressing human

GLUT4 in muscles and other insulin-sensitive tissues, these animals showed enhanced

basal and insulin-stimulated glucose disposal (Belke et al., 2001). Interestingly, a

maneuver to increase GLUT4 protein levels in diabetic C57BL/KsJ-Db/Db mice by

crossing them with a global tissue GLUT4 transgenic alleviated insulin resistance. A

similar cross, but in this instance with the cardiac-specific GLUT4 transgenic, led to

restoration of normal cardiac function ex vivo (Belke et al., 2000). To assess the role of

GLUT4 expression on whole-body glucose homeostasis and the development of type 2

diabetes, knock-out mouse models have been created by globally or selectively

disrupting the murine GLUT4 gene in all tissues (homozygote GLUT4-null) (Katz et

al., 1995), in skeletal muscles (Muscle-G4KO) (Zisman et al., 2000), in heart (G4H-/-)

(Abel et al., 1999) and in adipocytes (Adipose-G4KO) (Abel et al., 2001).

Stenbit et al. (2000) could not detect any significant difference in basal glucose uptake

in homozygote GLUT4-null mice. Signs of hyperglycemia were evident only after

insulin stimulation. Also, marked cardiac hypertrophy and increased fibrosis, without

hypertension, was demonstrated in homozygote GLUT4-null mice. Interestingly,

Stenbit et al. (1997) reported a different phenotype in heterozygote GLUT4-null mice

Chapter III

113

carrying 50% more GLUT4 levels than the homozygote: these mice suffered from

marked hyperglycemia and hypertension (Stenbit et al., 1997). Cardiac-specific

deletion of the GLUT4 in GH-/- mice also induced cardiac hypertrophy, although the

degree of remodeling was moderate and no signs of fibrosis, nor changes in blood

pressure were observed (Abel et al., 1999). Furthermore, Abel et al. (1999) observed no

changes in basal cardiac function in isolated hearts from G4H-/- mice. Interestingly,

however, ischemic conditions and catecholamine stimulation revealed diastolic

dysfunction in these mice, suggesting that additive factors such as hypoxia or systemic

changes in neuro-hormonal activation, in addition to lower GLUT4 content are

necessary to develop overt mechanical dysfunction in GLUT4-deficient hearts. Muscle-

and adipose-specific deletion of the GLUT4 did not cause any sign of cardiac

hypertrophy, although a detailed investigation of the cardiac structure and function in

these mice was not carried out.

The complex and nuanced systemic effects of the various GLUT4 gene knock-in and

knock-out in mice described above have been previously reviewed by Katz et al.

(1996), Charron et al. (1999) and more recently by Abel (2004). These reviews

extensively compare the different mouse models reported in the literature and

demonstrate that the facilitative glucose transporter GLUT4 is a primary effector

molecule for insulin-mediated glucose disposal in skeletal muscles and adipose tissue.

Indeed, the studies reported in these reviews demonstrate the important role played by

the GLUT4 transporter in the development of systemic insulin-resistance and type 2

diabetes, but fail to provide a link between glucose transporter expression and the

development of cardiac hypertrophy as it is observed in diabetic cardiomyopathy.

1.3.2. The GLUT4-KO mouse model

A more recently developed genetic model of insulin resistance is a mouse in which the

Cre-Lox system was employed to modulate GLUT4 expression in most insulin-

sensitive tissues (i.e. the GLUT4-KO mouse model). As this is the model utilized in the

study of insulin resistance-induced cardiomyopathy in this Thesis, a detailed

description of the model generation is provided below.

Chapter III

114

Kaczmarczyk et al. (2003) designed a Cre/Lox construct to selectively disrupt the

murine GLUT4 gene in tissue expressing skeletal α-actin (ie. predominantly skeletal

muscle) to induce insulin resistance. A vector was designed so that after homologous

recombination and gene replacement, the prokaryotic enzyme Cre-recombinase would

remove a 5 kb region between the two LoxP sites, which spanned exons 7 to 11 of the

GLUT4 gene together with the PGK/neoR selection cassette. F1 (C57BL6 x CBA)

mice were created to express a specific transgene containing the human muscle-specific

skeletal α-actin gene promoter (Muscat and Kedes, 1987) in association with the Cre-

recombinase (Cre+/- mice). According to wild-type developmental expression of actin

isoforms, the α-skeletal actin promoter directed transgene expression in the muscles

prenatally (cardiac muscle) and postnatally (mainly in skeletal muscles), (Brennan and

Hardeman, 1993; Miniou et al., 1999; Suurmeijer et al., 2003). Kaczmarczyk (2003)

reported strong expression of the Cre-recombinase in skeletal muscles of newborn

Cre+/- mice, with some expression in the cardiac tissue, and none in adipose tissue, liver

and kidney. Heterozygote Cre+/- mice were then mated with homozygote mice carrying

the LoxP-PGK/neoR construct (GLUT4-Lox+/+) to produce double heterozygote

GLUT4-Lox+/-Cre+/- mice. These animals were mated with GLUT4-Lox+/+ mice to

produce Lox+/+Cre+/- ‘knock-out’ mice (termed ‘Lox-Lox-Cre’ mice, LLC) and their

genetic controls GLUT4-Lox+/+Cre-/- (or ‘Lox-Lox’ mice, LL). The early genetic

background of LL and LLC mice (constituting the GLUT4-KO line) was estimated

56.25% C57BL6, 37.5% 129Sv and 6.25% CBA at the time of the experiments.

Expression studies to assess the outcome of the genetic manipulations show that control

LL mice exhibit a downregulation of GLUT4 expression in cardiac and skeletal

muscles and also in white and brown adipose tissues to levels 15–30% of wild-type

C57BL6 control mice (see Section 3.2.4 and Kaczmarczyk et al., 2003). As was

subsequently observed in other knock-out mouse models, it was concluded that the

PGK/neoR selection cassette was interfering with the normal expression of the GLUT4

gene in LL and LLC (Nagy, 2000; She et al., 2000), thus creating a Cre-independent

GLUT4 ‘knock-down’ effect in both LL and LLC mice.

Chapter III

115

Fasting glucose is not elevated, but insulin levels are double in both LL and LLC mice

when compared to C57BL6, indicating insulin resistance but not hyperglycemia in the

unstimulated state in both mice (Kaczmarczyk et al., 2003). Plasma glucose levels are

higher in the LL and LLC mice compared with C57BL6 (i.e. WT equivalent) during a

hyperinsulinemic clamp, suggesting postprandial hyperglycemia. Although the basal

rate of glucose uptake and metabolic clearance rate of glucose are not different in the

three groups of mice, in the insulin-stimulated state metabolic clearance rate is lower in

LL and LLC, indicating a defect in insulin-mediated glucose uptake in both LL and

LLC mice. Surprisingly, in both skeletal muscles and adipose tissue the levels of

GLUT4 protein downregulation are similar in LL and LLC mice, suggesting that the

expression of the Cre-recombinase in LLC mice does not reduce further the levels of

GLUT4 protein in these tissues. Although systemic metabolic abnormalities are

equivalent in LL and LLC mice, the latter present a dramatic cardiac hypertrophy,

associated with a ~99% suppression of GLUT4 levels, apparently driven by the

expression of the Cre-recombinase in this tissue. It is important to note that both LL

and LLC mice are normotensive, confirming that cardiac hypertrophy in these mice is

not related to pressure overload. When compared to C57BL6, insulin-stimulated

glucose uptake is normal in LL hearts, but is markedly reduced in LLC hearts. This

suggests that there is a threshold level of GLUT4 below which insulin-stimulated

glucose uptake is impaired.

It is not clear why the LLC mice do not exhibit full GLUT4 deletion in skeletal muscle

as is evident in the heart and as was anticipated initially. It is known that when a

transgene is passed through the female germ line, it is completely silenced in some

offspring while in others expression is reduced (Kearns et al., 2000). This phenomenon

is caused by DNA methylation. Following maternal inheritance, epigenetic

modification of the transgene locus can be cumulative over successive generations

resulting in an irreversible methylation after three consecutive germline passages in

some strains (mainly Balb/c, C57BL6 and CBA) (Allen et al., 1990). Furthermore,

cellular mosaicism due to variable gene expression and partial penetrance have been

observed in lacZ transgene expression in preimplantation stage mouse embryos

(Kothary et al., 1992; Guy et al., 1997; Lau et al., 1999) where the extent of variation in

expression was influenced by the genetic background of the oocyte. Balb/c, C57BL6

Chapter III

116

and F1 (from C57BL6 x CBA) genetic backgrounds gave none or very little lacZ

activity. DNA methylation, transgene silencing and/or different degrees of transgene

penetrance, as well as differential developmental time of α-skeletal actin promoter

activation in skeletal muscles and the myocardium, could account for the GLUT4

variability observed in LLC mice.

Compared to the other GLUT4 knockout mouse models discussed above, the model

presented here (the GLUT4-KO) shows similarities with homozygote GLUT4-null

(Katz et al., 1995). Indeed, no difference in basal glucose uptake in homozygote

GLUT4-null mice could be detected (Stenbit et al., 2000). Signs of hyperglycemia were

evident only after insulin stimulation. Also, marked cardiac hypertrophy and increased

fibrosis, without hypertension, was demonstrated in homozygote GLUT4-null mice.

The present GLUT4-KO model however presents some additional interesting and

specific traits. For instance, previous data comparing LL and LLC littermates shows

that the development of the cardiac hypertrophic phenotype in GLUT4-KO is

independent of systemic disturbances such as insulin resistance and postprandial

hyperglycemia (Kaczmarczyk et al., 2003). This is in agreement with other data

reporting that diabetic cardiomyopathy is not directly attributable to the development of

systemic microvascular disease, hypertension or systemic metabolic shift (Zarich and

Nesto, 1989; Galderisi et al., 1991; Shehadeh and Regan, 1995). Thus the GLUT4-KO

mouse model is of particular value in the investigation of the development of insulin

resistant cardiomyopathy specifically due to impaired glucose uptake in the heart.

Diabetic cardiac abnormalities are reported to occur as early as the glucose intolerance

phase - that is hyperglycemia and hyperinsulinemia that follow insulin resistance

(Celentano et al., 1995).

Chapter III

117

1.4. In summary

Transgenic and knock out models are proven tools which have been of considerable

value in the investigation of cardiac and cardiomyocyte remodeling associated with the

activation of the intra-cardiac RAS or with the disruption of the glucose metabolism.

These animals offer the unique opportunity to study the functional impact of chronic

Ang II signaling and altered glucose homeostasis in hearts of normotensive models in

vivo, where trophic effects can be evaluated in the absence of confounding loading

effects. A detailed study of structural and functional alterations associated with insulin

resistance or Ang II overproduction on the intact heart and on the cardiomyocyte has

not been previously undertaken. In particular, a morphological, physiological and

molecular characterization of TG1306/1R and GLUT4-KO mice can address

fundamental questions relating to the role of the intra-cardiac RAS and insulin

resistance in initiating cardiac remodeling processes. The extent to which these

genetically defined states exhibit common phenotypic elements is of particular interest.

1.5. Aims

The present study was designed to comparatively evaluate the effects of experimentally

induced chronic in vivo cardiac exposure to elevated Ang II levels and to impaired

GLUT4-mediated glucose uptake on cardiac and cardiomyocyte structure and

morphology. Specifically using the TG1306/1R and GLUT4-KO mice, the

experimental aims were:

1. To characterize structural remodeling of the two major components of the heart

- the cellular fraction and the extracellular matrix - and to evaluate the

progression of this remodeling with ageing.

Chapter III

118

2. To assess the levels of expression of selected genes of interest involved in the

trophic and structural cardiac remodeling of these two genetic models, namely

the angiotensinogen gene (Agt), the gap junction connexin 43 (Cx43) and the

insulin-stimulated glucose transporter GLUT4.

The goal of this study was to clarify whether cardiac and cardiomyocyte remodeling

induced by chronic Ang II overproduction or impaired GLUT4-dependent glucose

transport in vivo could have a primary and long-term impact on myocyte and

myocardial morphology, independently of hemodynamic changes. It was hypothesized

that tissue and myocyte hypertrophy would be evident in both mouse models and it

would be associated at least with differential expression of the Agt in the TG1306/1R

mouse and differential GLUT4 expression in the GLUT4-KO mouse.

Chapter III

119

2. METHODS

2.1. Whole heart histology and morphometry

2.1.1. Procedures

Immediately post-excision, the degree of cardiac hypertrophy was established by

measuring the ratio between the blotted wet heart weight and the body weight (mg/g),

i.e. the cardiac weight index (CWI). Excised hearts were fixed in 10% buffered

formalin, dehydrated and embedded in paraffin blocks as described in Chapter II,

Section 3.1. Macroscopic evaluation of cardiac structure and microscopic estimations

of collagen content (% total surface area covered by collagen fibres, %tsa) and number

of nuclei (total number per mm2) were performed by densitometric analysis as

described in Chapter II, Section 3.2.

2.1.2. Experimental groups

To investigate the age-dependent progressive effects of remodeling on cardiac

morphology, hearts of 15-20 and 35-40 week old male transgenic and knock-out mice

were studied and compared with hearts of age-matched littermate control mice. Eight

experimental groups were defined according to animal age and genotype (in Tables III-

1 and III-2 the number of animals utilized for each experimental procedure is

represented in parentheses):

1. 15-20 week TG: hearts of 15-20 week male transgenic TG1306/1R mice.

2. 15-20 week WT: hearts of 15-20 week male wild-type littermates from

heterozygote breeding of TG1306/1R with C57BL6 mice.

3. 35-40 week TG: hearts 35-40 week male transgenic TG1306/1R mice.

Chapter III

120

4. 35-40 week WT: hearts of 35-40 week male wild-type littermates from

heterozygote breeding of TG1306/1R with C57BL6 mice.

5. 15-20 week LLC: hearts of 15-20 week male knockout Lox+/+ Cre +/- mice.

6. 15-20 week LL: hearts of 15-20 week male control Lox+/+ Cre -/- littermates

from heterozygote breeding of LL and LLC mice.

7. 35-40 week LLC: hearts of 35-40 week male knockout Lox+/+ Cre +/- mice.

8. 35-40 week LL: hearts of 35-40 week male control Lox+/+ Cre -/- littermates

from heterozygote breeding of LL and LLC mice.

2.2. Longevity data and homozygote TG1306/1R mice

As mentioned in Chapter II, longevity data were also collected for an additional group

of TG (n= 13) and WT (n= 12) mice over a period of 94 weeks. In addition, to study

possible cardiac remodeling caused by variable cardiac Ang II production levels, the

creation of a small colony of homozygous TG1306/1R mice harboring double the

transgene complement was also attempted. Colony mortality data were not available for

the LL and LLC. These animals were housed in a facility not accessible to the

candidate. Anecdotal reports are strongly suggestive of premature mortality in the LLC

mice when compared to LL littermates.

2.3. RNA and protein extraction and quantification

2.3.1. RNA extraction, RT-PCR and Western blotting

Total RNA was extracted from mouse hearts and expression of angiotensinogen (Agt)

and connexin 43 (Cx43) mRNAs were assessed by semi-quantitative RT-PCR on 15-20

and 35-40 week old ventricles using the procedures detailed in Chapter II, Section 7.

The procedures for protein extraction and Western blotting were also described in

Chapter II, Section 6. For immunodetection of the glucose transporter GLUT4,

Chapter III

121

membranes were incubated with diluted rabbit anti-mouse GLUT4 polyclonal

antibodies overnight (4oC). The antibodies were kindly provided by the laboratory of

Dr. Joseph Proietto, The Royal Melbourne Hospital. After washing, blots were

incubated with peroxidase-conjugated secondary anti-rabbit IgG antibodies at room

temperature (Santa Cruz Biotechnology). In order to investigate possible correlations

between the heart weight and the expression levels of GLUT4 in the heart, the blotted

wet heart weight and the body weight were also recorded and CWI determine prior to

any RNA or protein extraction procedures.

2.3.2. Experimental groups

To investigate the gene and protein expression described above, hearts of 15-20 and 35-

40 week old male transgenic and knock-out mice were studied and compared with age-

matched control littermates.

Eight experimental groups were defined according to animal age and genotype as

previously defined. For all groups ventricular tissues were harvested for mRNA and/or

protein extraction as indicated.

1. 15-20 week TG: n=10 hearts for mRNA extraction and n=15 hearts for protein

extraction.

2. 15-20 week WT: n=13 hearts for mRNA extraction and n=10 hearts for protein

extraction.

3. 35-40 week TG: n=10 hearts for mRNA extraction.

4. 35-40 week WT: n=10 hearts for mRNA extraction.

5. 15-20 week LLC: n=10 hearts for mRNA extraction and n=18 hearts for

protein extraction.

6. 15-20 week LL: n=10 hearts for mRNA extraction and n=18 hearts for protein

extraction.

7. 35-40 week LLC: n=10 hearts for mRNA extraction.

8. 35-40 week LL: n=10 hearts for mRNA extraction.

Chapter III

122

2.4. Statistical considerations and presentation of the results

For the estimation of collagen content and the number of nuclei per mm2, the mean

values were determined according to the procedure illustrated (Figures II-1 and II-2)

and discussed in Chapter II, Section 3.2. For each experimental group, the ‘n’ values

reported here reflect the total number of hearts analyzed per experimental group.

Results are expressed as mean ± SEM. One-way ANOVA was performed to evaluate

differences between two experimental groups with respect genotype or age. A 2-way

ANOVA approach was used to statistically evaluate possible interactions between

genotype and age. Significant differences were identified at p<0.05. Legends of Figures

and Tables reporting specific results state the type of statistical test utilized and the

level of significance.

Chapter III

123

3. RESULTS

3.1. Cardiac and cardiomyocyte remodeling in TG1306/1R mice

3.1.1. Survival and the hypertrophic phenotype in TG mice

Compared with WT littermate controls, TG mice exhibited a significant increase in

mortality (Figure III-1, Panel A) over the 94 week longitudinal study (WT 83%

survival vs. TG 46% survival). Post mortem analysis showed that premature mortality

in the TG group was predominantly associated with occurrence of a dilated cardiac

phenotype, whereas the survivor TG mice exhibited a concentric hypertrophic

phenotype (Figure III-1, Panels B).

To further explore how remodeling severity and lethality were associated with variable

cardiac Ang II production levels, a small number of homozygous TG1306/1R mice

harboring double the transgene complement was generated. These mice exhibited a

dilated cardiomyopathy and died prematurely after 7-10 days (Figure III-2, Panel A).

Cardiac remodeling in homozygotes involved dramatic dilation of the right ventricle as

well as tissue and cardiomyocyte disarray (Figure III-2, Panel A and B).

3.1.2. Myocardial and chamber remodeling in TG hearts

To further evaluate the effects of Ang II on cardiac structure prior to the overt

appearance of the dilated phenotype, the cardiac morphology of 15-20 and 35-40 week

TG mice and age-matched WT littermates was investigated (Table III-1). In these TG

mice a concentric growth pattern was consistently observed, with significant increase in

left ventricular (LV) wall thickness at both ages in the absence of significant chamber

dilation. Some chamber dilation was observed in some 35-40 TG hearts, but the mean

Chapter III

124

values were not significantly different from WT. Thus the dilated phenotype observed

in about half the aged TG mice involved in the survival study appears to represent a

transition process which occurs later in life, in association with increased mortality.

In both WT and TG groups, ageing was associated with increased heart weight and LV

chamber luminal area, but no alteration in myocardial wall thickness was observed

between 15-20 and 35-40 week animals. This suggests that structural remodeling after

the age of 15-20 weeks mainly involved chamber morphology rather than myocardial

thickening.

3.1.3. Cardiomyocyte hypertrophy but not fibrosis in TG hearts

Densitometric analysis of stained histological sections showed that, despite a

significant increase in cardiac mass at both ages, there was no evidence of an increase

in interstitial fibrosis in TG hearts when compared to age-matched WT, nor in

association with ageing (Table III-1). The number of cell nuclei per mm2 was also

consistently preserved between the 4 groups. A regional difference in collagen

expression and nuclei count was observed between the LV and the right ventricle (RV)

in both WT and TG hearts. There was a ~300% increase in collagen-staining content in

the RV independent of genotype or age influence. This was associated with an increase

in the number of nuclei per mm2 observed in the RV, suggestive of an increased

number of non-myocyte/fibroblast cell types linked with increased collagen production

in the RV when compared to the LV.

Therefore, the increased cardiac mass in TG hearts appears to result primarily from

cardiomyocyte hypertrophy rather than collagen deposition. Indeed, the dimensions of

adult LV cardiomyocytes were found to be greater in TG hearts. Mean cell length was

significantly increased in the cardiomyocyte population of TG mice at both ages. In the

older TG myocytes the length change was associated with a significant increase in cell

width (Table III-1).

Chapter III

125

3.1.4. Differential gene expression in TG hearts

As cardiac and cardiomyocyte remodeling in the TG is attributed to cardiac-specific

overexpression of Agt, the mRNA levels for this gene were investigated in the hearts of

15-20 and 35-40 week old TG and WT mice. Since connexins allow electrical and

metabolic coupling between cardiomyocytes, properties that are important for

coordinated action of the heart as well as tissue homeostasis, the mRNA levels of the

connexin Cx43 was also investigated in the same hearts. mRNA analysis showed a

~10-fold upregulation of the Agt gene in the hearts of 15-20 week old TG when

compared to age-matched WT (Figure III-3, Panel A). Ageing had relatively no

influence on the Agt gene expression in TG hearts, but significantly increased the

levels of Agt mRNA in the hearts of WT mice. Cardiac and cardiomyocyte remodeling

due to Agt overexpression was associated with a 2-fold downregulation of the Cx43

mRNA levels in 15-20 week old TG hearts. Ageing also caused a general

downregulation of the expression of the Cx43. This effect was most pronounced in WT

hearts (Figure III-3, Panel B).

3.1.5. Decreased GLUT4 protein content in TG hearts

Since previous studies from GLUT4- deficient mice (including the LLC and LL) have

shown that cardiac structural and functional integrity is dependent on cell glucose

metabolism (Kaczmarczyk et al., 2003), the expression of GLUT4 was investigated in

the hearts of 15-20 week TG and WT mice. Cardiac and cardiomyocyte remodeling in

TG hearts is associated with approximately a 25% decrease in levels of GLUT4 protein

(Figure III-4, Panel A). Interestingly, in both WT and TG hearts, a negative relationship

between CWI and the level of GLUT4 was observed (i.e. the higher the CWI, the lower

the intra-cardiac concentration of GLUT4, per μl of total protein extracted) (Figure III-

4, Panel B and C). The slope of this negative relationship was accentuated in the TG

group suggesting that higher degrees of cardiac hypertrophy (CWI) induced by Ang II

overproduction results in larger decline of GLUT4 levels in the heart. This is evidence

Chapter III

126

for an indirect effect of Ang II in the regulation of GLUT4 expression levels in the

heart.

3.2. Cardiac and cardiomyocyte remodeling in GLUT4-KO mice

3.2.1. Myocardial and chamber remodeling in GLUT4-KO

hearts

Cardiac morphology was investigated in 15-20 and 30-40 week old LLC and littermate

LL hearts to evaluate the effects of a direct suppression of GLUT4 expression (Table

III-2). In the LLC a concentric growth pattern was observed at 15-20 weeks, with

significant increase in LV wall thickness, in the absence of significant chamber dilation

(Figure III-5). However, chamber dilation was observed in 35-40 week LLC hearts

when compared to age-matched LL hearts. Thus, the dilated phenotype observed in 35-

40 week LLC mice appears to represent a transition process which occurs after 15-20

weeks of age and it is preceded by concentric growth.

Ageing was associated with an increase in LV chamber luminal area, but not

myocardial wall thickening, suggesting that myocardial hypertrophy and LV wall

thickening preceded chamber dilation in this mouse model. This effect was observed in

both LL and LLC groups. Although a full longitudinal study was not carried out,

compared with LL littermate controls, LLC mice have been reported to exhibit an

apparent increase in mortality at younger age and under anesthesia (Dr. Kaczmarczyk

S, personal communication).

3.2.2. Cardiomyocyte hypertrophy and fibrosis in LLC hearts

Densitometric analysis of histological sections showed that cardiac hypertrophy in LLC

hearts was associated with a ~6-fold increase in interstitial fibrosis in the LV and a ~2-

fold increase in the RV (Table III-2). The number of cell nuclei per mm2 in the LV of

Chapter III

127

LLC hearts was positively correlated with the degree of myocardial collagen content

(Figure III-6), suggesting that an increase in collagen production could be associated

with increased number of non-myocyte/fibroblast cells in LLC hearts. A regional

difference in collagen expression and nuclei count was observed between the LV and

the right ventricle (RV) in LL hearts. Figure III-7 shows that two types of fibrosis were

present in LLC hearts: interstitial fibrosis (Panel B) and focal/reparative fibrosis (Panel

C); 6 of the 15 LLC hearts studied exhibited various degrees of focal/reparative fibrosis

on the LV and RV wall, suggesting that infarctive scars can develop and possibly

influence the rate of survival in this mouse model.

The increased cardiac mass in LLC hearts appears to result from both cardiomyocyte

hypertrophy and collagen deposition. Indeed, the dimensions of LV cardiomyocytes

were found to be greater in LLC hearts compared to those of LL hearts. Mean cell

length and width were significantly increased in the cardiomyocyte population of LLC

mice at both ages. There were no age-dependent modifications of cardiomyocyte

dimensions in LL and LLC myocytes (Table III-2).

3.2.3. Differential gene expression in hearts from LLC mice

To assess whether the intra-cardiac RAS was activated in LLC and LL hearts, the Agt

mRNA levels were investigated in 15-20 and 35-40 week old LLC and LL mice

(Figure III-8). As measured in TG and WT mice, the levels of Cx43 mRNA were

evaluated to determine whether the fibrotic remodeling in this mouse strain was

associated with altered cell-to-cell uncoupling. The results show a ~2-fold upregulation

of the Agt mRNA in the hearts of LLC when compared to age-matched LL. Ageing

caused a significant increase in the levels of Agt mRNA in the hearts of LLC and LL

mice. Cardiac and cardiomyocyte remodeling in the LLC mice was associated with a 2-

fold downregulation of the Cx43 mRNA levels in 15-20 week old LLC hearts. Ageing

was linked with a downregulation of the expression of the Cx43 in LL hearts.

Chapter III

128

3.2.4. Decreased GLUT4 protein content in LL and LLC hearts

To establish the extent to which Cre expression could be correlated with reduced levels

of the GLUT4 transporter in association with cardiac hypertrophy, protein expression

levels were investigated in the hearts of 15-20 week LLC and LL mice. Cardiac and

cardiomyocyte remodeling in LLC hearts was associated with a ~99% deletion of the

GLUT4 protein levels in the heart (Figure III-9, Panel A).

Interestingly, the proportional decline of the GLUT4 protein levels with the increase in

CWI detected in WT and TG mice (see Section 3.1.5), was not observed in LLC and

LL hearts (Figure III-9, Panel B and C). This could suggest strain differences.

Alternatively, it would confirm the hypothesis of a direct interference of the PGK-neoR

cassette in the normal modulation of the GLUT4 in LL hearts, with an additional effect

of Cre expression in the LLC mice. Indeed, when compared to 15-20 week WT

(C57BL6), an ~85% reduction in GLUT4 levels was observed in the hearts of age-

matched LL mice. This was associated with a modest but significant ~15% increase in

CWI in LL when compared with WT (C57BL6) (p<0.05, 1-way ANOVA) (Figure III-

10, Panel A). In the LLC mice, the dramatic hypertrophy (approximately 80-90%

increase in CWI when compared to WT (C57BL6) was associated with a loss of ~99%

of GLUT4 protein levels.

To explore the relationship between GLUT4 protein levels and CWI, paired data were

analyzed. The plot of cardiac GLUT4 levels and CWI in WT (C57BL6), LL and LLC

mice showed that there was a threshold effect in this relationship. Suppression of

GLUT4 protein levels in the heart to below ~5% of WT (C57BL6) levels was linked to

the fibrotic hypertrophic phenotype. There was no overlap between the WT (C57BL6),

LL and LLC distributions of these two variables, as shown in Figure III-10, Panel B.

Chapter III

129

4. DISCUSSION

4.1. Cardiac and cardiomyocyte remodeling in TG1306/1R mice

4.1.1. Concentric hypertrophy and ventricular dilation

in TG mice

The present study demonstrates that chronic overexpression of cardiac Agt is sufficient

to induce the development of two contrasting hypertrophic phenotypes in aged TG

mice. Those TG animals which died during the 94 week observation period exhibited a

dilated hypertrophic phenotype, whilst their longer-surviving TG littermates were

characterized by a concentric hypertrophy (Figure III-1). These findings indicate that

the concentric hypertrophic state is associated with a degree of functional compensation

and survival, and that the dilated phenotype is linked with increased mortality. These

results also suggest that functional decompensation precedes overt myocardial dilation

and supports the hypothesis that both concentric and dilated hypertrophy share some

common cellular pathways of development (Sussman et al., 2000).

The observations of homozygous TG mice suggest that an increased expression of the

transgene, and therefore of Ang II production, exacerbates the severity of the cardiac

phenotype observed (Figure III-2). This is consistent with preliminary findings that Agt

mRNA levels in the hearts of homozygote mice are doubled when compared with

heterozygotes (Dr. Cefai D, unpublished real-time PCR data). It could be proposed that

type and severity of myocardial remodeling are time- and Ang II dose-dependent.

Chapter III

130

4.1.2. Cardiac and cardiomyocyte remodeling but not

fibrosis in TG

This study extends previous observations that chronic overexpression of cardiac Ang II

in TG mice is sufficient to induce blood pressure-independent cardiac hypertrophy

without signs of fibrosis (Clement et al., 2001). Although qualitative changes in

collagen content can not be excluded, the detailed morphological analysis suggests that

cardiac remodeling results almost entirely from cardiomyocyte hypertrophy rather than

from ECM deposition (Table III-1). As previously observed by Caspari et al. (1977)

and Medugorac (1980), collagen concentrations in the LV free wall were substantially

lower than in the RV wall, but no age- or genotype-specific differences were detected.

These findings in the TG contrast with previous studies indicating Ang II involvement

in fibroblast proliferation and collagen production (Brilla et al., 1995; Bouzegrhane and

Thibault, 2002) and provide unambiguous evidence that elevated intracardiac Ang II

levels alone do not necessarily promote fibrosis.

An apparently contradictory finding is observed in a normotensive transgenic mouse

model in which overexpression of an Ang II-producing fusion protein was associated

with increased fibrosis but not hypertrophy (van Kats et al., 2001). These seemingly

different responses to cardiac Ang II overproduction may reflect a difference in the

origin of the Ang II (in the TG1306/1R insertion of the substrate peptide gene rather

than a protein fusion product) or trait differences between mouse strains (one renin

gene C57BL6 vs. two renin genes FVB/N mice). Lack of hypertension and stretch

stimulus on the myocardium, more rapid ECM protein turnover (Senzaki et al., 1998;

Coker et al., 2001), or absence of secondary stimuli (e.g. TNF-α, EGF and TGF-β1)

activating the non-myocyte cell population (Peng et al., 2002; Hao et al., 2004;

Stawowy et al., 2004), could account for the absence of a fibrotic phenotype in the TG

hearts. Interestingly, Min et al. (2004) have recently reported that AT1 receptor

stimulation in skin fibroblasts can increase collagen production. This is in part due to

the inhibition of collagen degradation via the increase of a tissue inhibitor of

metalloproteinase (TIMP)-1 expression. In the same study the stimulation of the AT2

Chapter III

131

receptor was found to exert the opposite effect, in part by the activation of Src

homology 2-containing protein-tyrosine phosphatase (SHP)-1. From these observations

the hypothesis arises that differential regulation of AT1 and AT2 receptor levels in

myocardial fibroblasts could result in collagen production or collagen degradation.

Most interesting is the apparent similarity between the cardiac phenotype described

here and that observed in transgenic mice overexpressing the Gαq protein specifically in

the heart. The Gαq overexpressing mice are characterized by cardiac decompensation in

the absence of cardiac fibrosis and pressure overload (D’Angelo et al., 1997; Sakata et

al., 1998). Given that signaling pathways mediated by AT1 receptors in myocytes are

known to be linked to the Gαq class of G proteins (Wettschureck et al., 2001), it is

tempting to speculate that cardiac remodeling in the TG1306/1R model is caused by

specific activation of cardiomyocyte AT1 Gαq-coupled receptors. Previous studies on

TG1306/1R mice already demonstrated that the trophic actions of Ang II on the heart

are mediated by AT1 receptors (Mazzolai et al., 1998; Mazzolai et al., 2000). This

hypothesis, where fibroblast participation in the hypertrophic response is not

obligatory, requires further investigation.

4.1.3. Differential gene expression profiles in TG hearts

The present study provides evidence of high levels of Agt mRNA produced in the

hearts of TG mice (Figure III-3). This is consistent with the expected transgenic

overexpression of the rat Agt gene in these hearts. From these data it can be

extrapolated that there is at least 10-fold higher Agt production in the hearts of

transgenic animals when compared to WT.

Interestingly, there is an age-dependent increase in Agt mRNA production in the hearts

of WT mice. This provides evidence that ageing is associated with enhanced

overexpression of the intra-cardiac RAS. This increased production of Agt in the WT

hearts is associated with an age-dependent decrease in connexin Cx43 expression.

Gradual increases in cardiac Ang II levels have been reported experimentally and

clinically during the development of chronic heart failure (Wollert and Drexler, 1999;

Chapter III

132

Serneri et al., 2001). In addition, numerous experimental and clinical studies have

demonstrated reduced Cx43 expression and impaired intercellular communication with

increased arrhythmogenic activity in the development of heart failure (Peters, 1997;

Itoh et al., 2002; Kitamura et al., 2002; Kostin et al., 2004; Poelzing and Rosenbaum,

2004). The present work suggests that age-dependent accumulation of Ang II in the

heart could lead to molecular changes promoting structural remodeling and

abnormalities in cardiomyocyte coupling, ultimately leading to abnormal cell-to-cell

communication, metabolism and E-C coupling. This age-dependent functional

deterioration is accelerated in the TG mice due to the early activation of the intra-

cardiac RAS.

The finding in the TG that intra-cardiac Ang II-overproduction leads to a decrease in

Cx43 expression contrasts with previous studies indicating Ang II involvement in

overexpression and activation of the gap junction Cx43. It is generally reported that

compensatory growth after Ang II injection or infusion is associated with increased

connexin levels, increased number of gap junctions and enhanced intercellular coupling

in neonatal cardiomyocyte cultures (Dodge et al., 1998; Shyu et al., 2001; Dhein et al.,

2002; Polontchouk et al., 2002). However, the cardiac hypertrophic response to a

chronic stimulus is a dynamic continuum in which early shifts in gene expression

initiate adaptive structural and functional changes that ultimately lead to increasing

maladaptive responses and further evolving shifts in gene expression. It is likely that

Cx43 expression is elevated after an acute hypertrophic stimulus but is suppressed in

the chronic situation. Furthermore, Ang II stimulation of adult cardiomyocytes isolated

from cardiomyopathic hamsters was associated with decreased gap junctional

conductivity and cell-to-cell uncoupling. The response was suppressed by Losartan and

Enalapril (De Mello, 1996). The mechanisms leading to connexin downregulation in

chronic heart remodeling are poorly understood, but there is evidence that signaling is

mediated by JNK and possibly involves the homeodomain transcription factor Nkx2.5

(Petrich et al., 2002; Kasahara et al., 2003). Interestingly, Itoh et al. (2002) reported

that depressed mRNA expression levels for Cx43 in volume overload-induced heart

failure in rabbits were partially restored after treatment with an AT1 blocker (CV-

11974). This suggests that volume overload may downregulate Cx43 expression

through an AT1-dependent signaling pathway.

Chapter III

133

4.1.4. Decreased GLUT4 protein levels in hearts of TG mice

The present study demonstrates that cardiac and cardiomyocyte remodeling in TG

hearts is associated with a ~25% decrease in cardiac GLUT4 protein levels (Figure III-

4, Panel A). Interestingly, in both WT and TG hearts, a negative relationship between

CWI and the level of cardiac GLUT4 is also observed (Figure III-4, Panel B and C).

This suggests that accentuated hypertrophy (CWI) results in more marked decline of

GLUT4 levels in the heart. Although the data from homozygote TG would suggest so,

it is not possible to determine from the present study whether higher CWI is associated

with higher levels of cardiac Agt mRNA expression and therefore Ang II production

(because the hearts used for detection of Agt and GLUT4 expression levels were

derived from different animals). Further experiments are required to establish such a

direct association. If confirmed, it would directly link the severity of cardiac

remodeling and the expression levels of GLUT4 to the degree of activation of the intra-

cardiac RAS. Such a finding would suggest that Ang II, via modulation of GLUT4

expression is involved in the development of insulin resistance in the heart.

4.2. Cardiac and cardiomyocyte remodeling in LLC mice

4.2.1. Cardiomyocyte hypertrophy and fibrotic remodeling

in LLC

Myocardial fibrosis and myocyte hypertrophy are the most frequently proposed

mechanisms to explain cardiac changes observed in diabetic and pre-diabetic

cardiomyopathies. Clinical and experimental studies have shown that diabetes is

associated with an increase in collagen formation in the myocardium (Shimizu et al.,

1993; Mizushige et al., 2000), as well as altered expression of contractile proteins and

mechanical dysfunction (see Chapter IV). The insulin-resistant LLC mice present

interstitial and focal/reparative fibrosis (Figure III-7), associated with cardiomyocyte

Chapter III

134

hypertrophy (Table III-2). Fibrosis would contribute to increased myocardial stiffness

and electrical heterogeneity, thus increasing susceptibility of the heart to

arrhythmogenic activity. Electrical heterogeneity and cell-to-cell discontinuity could be

exacerbated by Cx43-mediated gap junctional impairment, as shown in LLC and

senescent LL hearts (Figure III-8). The origins of fibrosis in diabetic cardiomyopathy

are probably multifactorial. A favored hypothesis is that collagen accumulation in the

diabetic myocardium is due to impaired collagen degradation resulting from

glycosylation of the lysine residues on collagen (Avendano et al., 1999; Candido et al.,

2003; Liu et al., 2003). Glycosylation processes could also involve the apoptotic signal

p53, leading to cardiac Ang II production and Ang II-dependent myocyte cell death and

reparative scar/fibrosis (Fiordaliso et al., 2001), as observed in LLC hearts. This

hypothesis is supported by the present findings where LLC hearts overexpress Agt

mRNA and exhibit reparative fibrotic scars (Figure III-7).

Hyperglycemia also results in the production of ROS and oxidative stress that could

contribute to differential gene expression of molecules involved in p53-mediated cell

death signaling (Cai et al., 2002). Ang II-mediated pro-fibrotic and pro-infarctive

effects could also be exacerbated by impaired responsiveness to insulin and insulin-like

growth factor (IGF)-I stimulation. Indeed, both p53- and Ang II-mediated apoptosis are

normally reduced by IGF-I (Jost-Vu et al., 1992; Leri et al., 2000). IGF-I is decreased

in diabetes, and exogenous IGF-I treatment has been shown to ameliorate contractile

disturbances in diabetic animals (Kajstura et al., 2001). The effects of Ang II on ECM

remodeling may also be promoted by activation of fibroblasts releasing endothelin-1

and TGF-β1 (Lee et al., 1995). Ang II-dependent release of endothelin-1 could promote

mast cells infiltration and degranulation, leading to the release of pro-inflammatory and

pro-fibrotic cytokines and chymases within the myocardium and the myocardial

vasculature (Gilber et al., 2000; Gordon, 2000; Murray et al., 2004; Skrabal et al.,

2004). Such molecular and morphological remodeling processes could be prominent in

the age-dependent LV dilatation observed in the LLC hearts. Degranulation of mast

cells, activation of fibroblasts, qualitative changes in collagen isoforms and MMP

activity could lead to disorganization of the established fibrillar network, and cause

myocyte slippage. Finally, the present study shows a positive correlation between the

collagen content in the LV myocardium of LLC mice and the number of nuclei counted

Chapter III

135

per mm2 in the same tissue (Figure III-6). This result suggests that collagen production

and deposition in LLC hearts is associated with increased number of cells per mm2 of

tissue, possibly caused by proliferation of non-myocyte cell types.

4.2.2. Decreased cell-to-cell connectivity in LLC hearts

As reviewed in Chapter I, hyperglycemia and insulin resistance appear to be important

etiologic factors in the development of cardiovascular complications in diabetic

patients. In particular, degraded gap junctional intercellular communication plays an

important role in cardiovascular tissue homeostasis of diabetic patients and

experimental models. Impaired ventricular conduction in diabetic patients and in

streptozotocin (STZ)-injected rats is associated with enhanced cardiac arrhythmic risk,

which may depend on connexin downregulation and impaired cell-to-cell

communication (Robillon et al., 1999; Inoguchi et al., 2001; Eloff et al., 2003). For

instance, downregulation of gap junction Cx43 expression by high extracellular glucose

concentrations has been demonstrated in cultured aortic smooth muscle cells (Kuroki et

al., 1998), rat microvascular endothelial cells (Sato et al., 2002) and bovine retinal

endothelial cells (Fernandes et al., 2004). Diabetic cardiomyopathy in streptozotocin-

injected rats was shown to be associated with a decrease in Cx43 expression

(Okruhlicova et al., 2002). As reviewed by Fernandez-Real and Ricart (2003), insulin

resistance and atherosclerosis share similar pathophysiological mechanisms, mainly

due to the actions of TGF-β and proinflammatory cytokines, such as TNF-α, IL-6 or

leptin. Dysregulation of the inflammatory axis predicts the development of insulin

resistance and type 2 diabetes in patients. Interestingly, hypoxia, ROS, TGF-β1 and

various cytokines have been shown to mediate downregulation of intercellular

communication in vitro between confluent fibroblast, endothelial cells or epithelial

tumor cells and in vivo in the myocardium (Hu and Cotgreave, 1997; van Rijen et al.,

1998; Nishida et al., 2000; Janczewski et al., 2004). This suggests that the same pro-

fibrotic mechanisms that are activated in diabetic cardiomyopathy could also lead to

cell-to-cell uncoupling and promote conduction defects in the diabetic heart.

Chapter III

136

4.2.3. Decreased GLUT4 protein levels in hearts of LLC and LL

mice

The study of GLUT4 protein expression in LL and LLC mouse hearts suggests that in

active cardiac muscle, there is a threshold below which GLUT4 deficiency impairs

insulin-stimulated glucose uptake (Figure III-10). Indeed, Kaczmarczyk et al. (2003)

reported that insulin-stimulated cardiac glucose uptake was normal in LL mice with

~15% of wild-type GLUT4 levels, but was reduced in mice with near complete absence

of this transporter. Earlier mouse models of GLUT4 deficiency, whether global (Stenbit

et al. 2000), or cardiac-specific (Abel et al. 1999) also exhibited cardiac hypertrophy in

association with absence of GLUT4 expression in the heart. Adipose- and skeletal

muscle-specific GLUT4 knock-out mice presented normal heart weight-to-body weight

indices in association with normal cardiac GLUT4 expressions (Zisman et al., 2000;

Abel et al., 2001). Therefore, it can be speculated that cardiac-specific downregulation

of GLUT4 levels can promote cardiac and cardiomyocyte remodeling.

In heart-specific GLUT4-deficient mice (Abel et al. 1999), the extent of hypertrophy

was more modest than that described for the homozygote GLUT4-null model (Stenbit

et al. 2000). More importantly, heart-specific GLUT4-deficiency was not associated

with systemic metabolic disturbances, nor collagen accumulation. Thus, the authors

speculated that systemic metabolic abnormalities were necessary to induce dramatic

exacerbation of the hypertrophic phenotype and appearance of interstitial fibrosis. The

present study does not support such a conclusion. LLC animals had similar deficiency

of GLUT4 levels in peripheral tissues with a consequent similar level of

hyperinsulinemia and a similar reduction in metabolic clearance rate of glucose

compared to LL mice (Kaczmarczyk et al., 2003). However, LLC mice exhibited a very

different degree of cardiac and cardiomyocyte hypertrophy, in association with reactive

and reparative fibrosis, genetic activation of the intra-cardiac RAS and evidence of cell-

to-cell uncoupling. This remodeling appeared to be related to the severity of GLUT4

deficiency and the reduction in glucose supply to the heart. Thus it could be proposed

that deranged intra-cardiac glucose metabolism, in combination with other events, such

Chapter III

137

as tissue damage and inflammatory reaction is necessary to cause dramatic remodeling

of the ECM of the myocardium and to stimulate non-myocyte proliferation and

exacerbate cardiomyocyte hypertrophy.

5. IN SUMMARY

Data presented in this Chapter substantiate the conclusion that cardiac and

cardiomyocyte remodeling in the mouse heart is associated with the stimulation of the

intra-cardiac renin angiotensin system and a decrease in GLUT4 expression levels.

Chronic myocardial remodeling is likely linked to cell-to-cell uncoupling as indicated

by the downregulation of the gap junction connexin 43 (Cx43). The severity of

cardiomyocyte and tissue remodeling is determined by the extent of myocardial

collagen deposition and possibly by the degree of perturbation of the intra-cardiac

glucose metabolism. These comparative findings will be further considered in the

concluding Chapter VI, where the variable findings described here for the two

experimental models will be discussed, and comment will be made regarding study

limitations and future directions.

Chapter III

138

2773±45 (6)2701±58 (6)2831±63 (11)2758±53 (11)(nb/mm2)LV nuclei

2978±64 (6)2949±67 (6)2921±52 (11)2952±67 (11)(nb/mm2)RV nuclei

2.6±0.2 (6)2.3±0.2 (6)1.4±0.1 (11)1.4±0.1 (10)(mm2)LV lumen

1.5±0.05* (6)1.1±0.05 (6)1.4±0.05* (10)1.1±0.02 (10)(mm) LV free wall

(μm)

(μm)

(%tsa)

(%tsa)

(mg/g)

(mg)

31.3±0.5 (6)27.5±0.5 (6)29.0±1.0 (7)27.7±0.7 (7)Myoc. width

168.4±3.9* (6)157.1±1.8 (6)178.7±3.8* (7)150.9±2.5 (7)Myoc. length

3.5±0.2 (6)3.2±0.2 (6)3.3±0.3 (11)3.1±0.3 (11)RV collagen

1.2±0.2 (6)1.3±0.2 (6)1.1±0.3 (11)1.0±0.2 (11)LV collagen

5.5±0.2* (20)4.5±0.1 (20)5.4±0.1* (20)4.9±0.1 (20)CWI

200.5±9.4* (20)145.1±3.6 (20)152.6±6.7* (20)131.3±2.5 (20)Heart weight

TGWTTGWT

35-40 week15-20 weekParameter

2773±45 (6)2701±58 (6)2831±63 (11)2758±53 (11)(nb/mm2)LV nuclei

2978±64 (6)2949±67 (6)2921±52 (11)2952±67 (11)(nb/mm2)RV nuclei

2.6±0.2 (6)2.3±0.2 (6)1.4±0.1 (11)1.4±0.1 (10)(mm2)LV lumen

1.5±0.05* (6)1.1±0.05 (6)1.4±0.05* (10)1.1±0.02 (10)(mm) LV free wall

(μm)

(μm)

(%tsa)

(%tsa)

(mg/g)

(mg)

31.3±0.5 (6)27.5±0.5 (6)29.0±1.0 (7)27.7±0.7 (7)Myoc. width

168.4±3.9* (6)157.1±1.8 (6)178.7±3.8* (7)150.9±2.5 (7)Myoc. length

3.5±0.2 (6)3.2±0.2 (6)3.3±0.3 (11)3.1±0.3 (11)RV collagen

1.2±0.2 (6)1.3±0.2 (6)1.1±0.3 (11)1.0±0.2 (11)LV collagen

5.5±0.2* (20)4.5±0.1 (20)5.4±0.1* (20)4.9±0.1 (20)CWI

200.5±9.4* (20)145.1±3.6 (20)152.6±6.7* (20)131.3±2.5 (20)Heart weight

TGWTTGWT

35-40 week15-20 weekParameter

Table III-1: Age-dependent cardiac structural remodeling in WT and TG mice

The cardiac weight index (CWI), expressed in heart weight/body weight (mg/g); collagen density expressed as % of area occupied by collagen of total field analyzed (%tsa); nuclei density expressed as number of nuclei per mm2 (nb/mm2). The left ventricular (LV) lumen was measured (in mm2) to evaluate chamber dilation, while LV free wall thickness (mm) evaluated myocardial hypertrophy. RV= right ventricle. Numbers of animals are shown in parenthesis. * p<0.05, TG vs. age-matched WT (1-way ANOVA); ‡ p<0.05, age-dependent effect (measured by multiple-way ANOVA); † p<0.05, age-genotype interaction effect (value derived from analysis by multiple-way ANOVA).

Chapter III

139

3821±64* (7)2741±54 (7)3758±53* (8)2672±43 (8)(nb/mm2)LV nuclei

3644±57 (7)2936±43 (7)3538±54 (8)2838±42 (8)(nb/mm2)RV nuclei

3.2±0.4* (7)2.2±0.2 (7)1.5±0.2 (8)1.4±0.1 (8)(mm2)LV lumen

1.7±0.1* (7)1.3±0.1 (7)1.7±0.1* (8)1.2±0.1 (8)(mm) LV free wall

(μm)

(μm)

(%tsa)

(%tsa)

(mg/g)

(mg)

32.8±0.4* (7)28.3±0.5 (7)31.9±1.1* (7)29.0±1.3 (7)Myoc. width

210.6±4.2* (7)173.1±1.7 (7)214.7±4.6* (7)169.1±2.9 (7)Myoc. length

5.9±0.3* (7)3.2±0.2 (7)5.4±0.3* (8)2.6±0.3 (8)RV collagen

6.2±0.4* (7)1.4±0.1 (7)6.7±0.3* (8)1.3±0.1 (8)LV collagen

9.2±0.8* (20)5.9±0.6 (20)8.8±0.9* (21)5.8±0.5 (22)CWI

258.4±6.3* (21)162.3±4.2 (22)211.4±7.1* (21)137.3±2.9 (22)Heart weight

LLCLLLLCLL

35-40 week15-20 weekParameter

3821±64* (7)2741±54 (7)3758±53* (8)2672±43 (8)(nb/mm2)LV nuclei

3644±57 (7)2936±43 (7)3538±54 (8)2838±42 (8)(nb/mm2)RV nuclei

3.2±0.4* (7)2.2±0.2 (7)1.5±0.2 (8)1.4±0.1 (8)(mm2)LV lumen

1.7±0.1* (7)1.3±0.1 (7)1.7±0.1* (8)1.2±0.1 (8)(mm) LV free wall

(μm)

(μm)

(%tsa)

(%tsa)

(mg/g)

(mg)

32.8±0.4* (7)28.3±0.5 (7)31.9±1.1* (7)29.0±1.3 (7)Myoc. width

210.6±4.2* (7)173.1±1.7 (7)214.7±4.6* (7)169.1±2.9 (7)Myoc. length

5.9±0.3* (7)3.2±0.2 (7)5.4±0.3* (8)2.6±0.3 (8)RV collagen

6.2±0.4* (7)1.4±0.1 (7)6.7±0.3* (8)1.3±0.1 (8)LV collagen

9.2±0.8* (20)5.9±0.6 (20)8.8±0.9* (21)5.8±0.5 (22)CWI

258.4±6.3* (21)162.3±4.2 (22)211.4±7.1* (21)137.3±2.9 (22)Heart weight

LLCLLLLCLL

35-40 week15-20 weekParameter

Table III-2: Age-dependent cardiac structural remodeling in LL and LLC mice

The cardiac weight index (CWI), expressed in heart weight/body weight (mg/g); collagen density expressed as % of area occupied by collagen of total field analyzed (%tsa); nuclei density expressed as number of nuclei per mm2 (nb/mm2). The left ventricular (LV) lumen was measured (in mm2) to evaluate chamber dilation, while LV free wall thickness (mm) evaluated myocardial hypertrophy. RV= right ventricle. Number of animals shown in parenthesis. * p<0.05, LLC vs. age-matched LL (one-way ANOVA); ‡ p<0.05, age-dependent effect (multiple-way ANOVA).

Chapter III

140

100

80

60

40

20

%

20 40 60 80 100weeks

WT(10/12)

TG(6/13)

age

surv

ival

A.

100

80

60

40

20

%

20 40 60 80 100weeks

WT(10/12)

TG(6/13)

age

surv

ival

100

80

60

40

20

%

20 40 60 80 100weeks

WT(10/12)

TG(6/13)

age

surv

ival

A.

1. WT 2. TG – ‘concentric’

2.5 mm 2.5 mm2.5 mm

RVRV

RVLV LVLV

3. TG – ‘dilated’

B.1. WT 2. TG – ‘concentric’

2.5 mm 2.5 mm2.5 mm

RVRV

RVLV LVLV

3. TG – ‘dilated’

B.

Figure III-1: Decreased survival and dilated hypertrophy in TG hearts

A. Survival plot of heterozygous TG versus WT littermates. Survival expressed as % of the total number of animals per group surviving at observational time (in weeks). Number of surviving animals as fraction of total number shown in parenthesis. CWI of non-surviving TG= 11.3±1.1 mg/g (n= 7) and WT= 9.4±1.0 (n= 2); B. Transverse sections of hearts from 94 week WT and TG mice. Panel 1: WT; Panel 2 and 3: TG heterozygous exhibiting contrasting concentric and dilated phenotypes. LV= left ventricle; RV= right ventricle.

Chapter III

141

2. TG heterozygote 3. TG homozygote1. WT

150μm150μm150μm

1 mm1 mm1 mm

LV

RV

RV

RV

LV LV

A.

B.

2. TG heterozygote 3. TG homozygote1. WT

150μm150μm150μm

1 mm1 mm1 mm

LV

RV

RV

RV

LV LV

A.

B.

Figure III-2: Ang II expression levels and severity of remodeling

A. Transverse sections of hearts from 1 week WT and TG mice. Panel 1: WT; Panels 2 and 3: heterozygote and homozygote neonatal TG phenotypes. Remodeling severity reflected Ang II production levels in the heart: TG homozygotes developed a dilated cardiomyopathy and died prematurely (<10 days). B. Only cardiac remodeling in homozygotes involved tissue and cardiomyocyte disarray.

Chapter III

142

A.

Expr

e ssi

on le

vel

(Arb

. Uni

ts)

0

70

140

210

280

350

a young old15-20 week 35-40 week

Agt

0

100

200

300

a young old15-20 week 35-40 week

B. Cx43

Expr

essi

o n le

vel

(Arb

. Uni

ts )

TG (n=10)WT (n=10)

‡ ‡

‡ ‡

Figure III-3: Agt overexpression and downregulation of Cx43 in TG and WT hearts

A. Expression profiles of Agt mRNA and B. connexin Cx43 mRNA. In parenthesis are the number of animals per group. ‡ p<0.05, genotype-dependent effect (multiple-way ANOVA); † p<0.05, age-genotype interaction effect (value derived from analysis by multiple-way ANOVA).

Chapter III

143

Figure III-4: GLUT4 protein expression and negative correlation with CWI in TG and WT hearts

A. GLUT4 protein expression in WT and TG hearts. *p<0.05, TG vs. WT (one-way ANOVA). B. Plot of cardiac GLUT4 levels and CWI in 15-20 week WT mice. C. The same plot in TG hearts. A negative relationship between the CWI and GLUT4 protein levels was observed in both TG and WT hearts.

3

4

5

6

7

8

60 80 100 120 140 160 180

Y = 6.728 - .016 * X; R2 = .781

3

4

5

6

7

8

0 20 40 60 80 100 120 140 160

Y = 8.196 - .024 * X; R2 = .827

0

40

80

120

160

Expr

essi

on le

vel

(Arb

. Uni

ts)

WT TG

*

A.

B.

C.

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

GLUT4

WT

TG

3

4

5

6

7

8

60 80 100 120 140 160 180

Y = 6.728 - .016 * X; R2 = .781

3

4

5

6

7

8

0 20 40 60 80 100 120 140 160

Y = 8.196 - .024 * X; R2 = .827

0

40

80

120

160

Expr

essi

on le

vel

(Arb

. Uni

ts)

WT TG

*

A.

B.

C.

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

GLUT4

WT

TG

Chapter III

144

A.

B.

C.

LV

1 mm

RVLV

1 mm1 mm

RV

1 mm

LVRV

1 mm1 mm

LVRV

1 mm

LV RV

1 mm1 mm

LV RV

LLC

LL

WT

Figure III-5: Cardiac remodeling in LL and LLC hearts

Transverse sections of cardiac ventricles from 15 week old A. C57BL6 (i.e. WT), B. LL mouse and C. LLC mouse. Cardiac hypertrophy without dilatation was evident in LLC mice at 15-20 weeks. Senescence caused chamber dilatation in LLC mice (not shown on the picture). LV= left ventricle; RV= right ventricle.

Chapter III

145

Figure III-6: Fibrosis and non-myocyte proliferation in LLC hearts

Plot of LV myocardial collagen content and number of counted nuclei per mm2 of tissue in LLC mice. Each point represents a LLC heart (15-20 or 35-40 weeks). A positive relationship between the collagen content and the number of nuclei was observed in fibrotic LLC hearts. Similar positive relationships were observed for the collagen and nuclei counts in the right ventricle of LL, LLC and TG hearts (data not shown).

5

6

7

8

9

3000 3400 3800 4200

Y = -3.191 + .003 * X; R2 = .831

Number of nuclei per mm2

Col

lage

n le

vels

(%ts

a )

5

6

7

8

9

3000 3400 3800 4200

Y = -3.191 + .003 * X; R2 = .831

Number of nuclei per mm2

Col

lage

n le

vels

(%ts

a )

Chapter III

146

A.

B.

C.

100μm

100μm

100μm

LLC

LLC

LL

Figure III-7: Diffuse and focal/reparative fibrosis in LLC hearts

Left ventricular histological sections stained by Van Gieson’s method. A. No fibrosis evident remodeling in LL hearts. B. Diffuse fibrosis was present in LLC mouse ventricular tissue. C. Reparative fibrosis was observable in LLC hearts, especially at the level of the pericardium.

Chapter III

147

LLC (n=10)LL (n=10)

0

40

80

120

a young old15-20 week 35-40 week

A. Agt

Expr

e ssi

on l e

vel

(Arb

. Uni

ts)

0

75

150

225

a young old15-20 week 35-40 week

B. Cx43

Expr

e ssi

on le

vel

(Arb

. Uni

ts)

§

‡ ‡

‡ ‡

Figure III-8: Agt and Cx43 gene expression profiles in LLC and LL hearts

A. Expression profiles of Agt mRNA and B. connexin Cx43 mRNA. In parenthesis are the number of animals per group. ‡ p<0.05, genotype-dependent effect (multiple-way ANOVA); † p<0.05, age-genotype interaction effect (value derived from analysis by multiple-way ANOVA); § p<0.05, age-dependent effect (multiple-way ANOVA).

Chapter III

148

Figure III-9: GLUT4 protein expression and correlation with CWI in LLC and LL hearts

A. GLUT4 protein expression in LL and LLC hearts. *p<0.05, LLC vs. LL (one-way ANOVA). B. Plot of cardiac GLUT4 levels and CWI in 15-20 week LL mice. C. The same plot in LLC hearts. No correlation between the the CWI and GLUT4 protein levels was observed in LLC and LL hearts.

4

5

6

7

8

0 10 20 30 40 50 60

Y = 5.408 - 0.006 * X; R2 = 0.032

0

3

6

9

12

0.0 .5 1.0 1.5 2.0

Y = 8.223 - .025 * X; R2 = 1.441E-4

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

0

5

10

15

20

25

Expr

essi

on le

vel

(Arb

. Uni

ts)

LL LLC

A.

B.

C.

*

GLUT4

LL

LLC

4

5

6

7

8

0 10 20 30 40 50 60

Y = 5.408 - 0.006 * X; R2 = 0.032

0

3

6

9

12

0.0 .5 1.0 1.5 2.0

Y = 8.223 - .025 * X; R2 = 1.441E-4

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

CW

I (m

g/g)

GLUT4 levels (Arb. Units)

0

5

10

15

20

25

Expr

essi

on le

vel

(Arb

. Uni

ts)

LL LLC

A.

B.

C.

*

GLUT4

LL

LLC

Chapter III

149

Figure III-10: GLUT4 protein expression and correlation with CWI in LLC and LL vs. WT hearts

A. Western blot analysis of cardiac GLUT4 protein expression in C57BL6 (i.e. WT), LL and LLC mice. LL mice had ~85% reduction levels when compared to C57BL6. Cre expression caused further reduction in GLUT4 to very low levels. B. Plot of GLUT4 levels vs. the cardiac weight index (CWI) in C57BL6, LL and LLC hearts. A threshold level is observed below which cardiac hypertrophy is complicated by collagen deposition and impaired glucose transport (Kaczmarczyk et al., 2003). * p<0.05, LL vs. WT; † p<0.05, LLC vs. WT and LL (multiple-way ANOVA).

0

35

70

105

140

4

6

8

10

12

0 40 80 120 160

CW

I (m

g/g)

GLUT4 levels ( Arb. Units)

Expr

essio

n le

vel

(Arb

. Uni

ts)

C57BL6 LL LLC

A.

B.

GLUT4

C57BL6LLLLC

Threshold

*†

0

35

70

105

140

4

6

8

10

12

0 40 80 120 160

CW

I (m

g/g)

GLUT4 levels ( Arb. Units)

Expr

essio

n le

vel

(Arb

. Uni

ts)

C57BL6 LL LLC

A.

B.

GLUT4

C57BL6LLLLC

Threshold

*†

CHAPTER IV

Impaired cardiomyocyte contractility and

differential expression of calcium and proton

transporters in ageing models of Ang II-induced

and insulin resistant cardiac hypertrophy

Chapter IV

151

1. INTRODUCTION

1.1. E-C coupling with increased pacing frequency in the mouse

The modulation of cardiac and cardiomyocyte contractility by pacing frequency is a

fundamental property of cardiac muscle. In human and large mammals, an increase in

the steady-state frequency under physiological conditions generally enhances

contraction and relaxation and prolongs the contractile cycle (Freeman et al., 1987;

Gwathmey et al., 1990; Bers, 2001). A number of E-C coupling processes have been

shown to be frequency dependent. A positive force-frequency relationship could result

from increased inward calcium currents, increased SR calcium content, higher diastolic

intracellular calcium concentration and possible higher fractional SR calcium release

during contraction (Bers, 2001). In pathological conditions, this positive frequency

relationship may be diminished or converted into a negative relationship, as

demonstrated in human cardiac hypertrophy (Liu et al., 1993) and failure (Davies et al.,

1995; Brixius et al., 2001).

In relatively recent years, the mouse has become an important animal in which to study

cardiovascular physiology and pathology due to the development of murine transgenic

technology. Compared with larger mammals, the basic cardiac physiology would be

expected to be quite distinctive in this small animal with high heart rate in vivo. Indeed,

in studies performed in isolated tissue or with cardiomyocytes from mice, negative,

positive or even biphasic frequency relationships have been observed (Brooks and

Apstein, 1996; Hoit et al., 1997; Palakodeti et al., 1997; Gao et al., 1998; Kadambi et

al., 1999). Several E-C coupling features of the murine heart suggest that, while not

negative, the force-frequency relationship should be only moderately graded over

physiological heart rates (5-8 Hz) (reviewed in Bers, 2001). This was confirmed by

Georgakopoulos and Kass (2001) in open-chest mice and by Stull et al. (2002) in

mouse trabeculae. It has also been shown that ageing is associated with a blunted

positive force-frequency relationship in whole hearts and cardiomyocytes of mice (Lim

Chapter IV

152

et al., 1999; Lim et al., 2000). In these studies the blunted response was linked with

impaired cardiac and cardiomyocyte relaxation (or lusitropy) and a prolonged calcium

transient.

1.2. Modulation of cardiomyocyte contractility by Ang II

1.2.1. Acute modulation of myocyte contractility by Ang II

Modulation of cardiac and cardiomyocyte contractility is one of the many cellular

responses ascribed to Ang II in the heart (Koch-Weser, 1965; Kobayashi et al., 1978).

Ang II is considered to act acutely as a positive inotropic agent, although with modest

potency when compared to other peptides such as endothelin-1 (Sakurai et al., 2002).

This reflects the finding that Ang II receptor density is relatively low for adult

cardiomyocytes under normal physiological conditions (Touyz et al., 1996). A positive

contractile (or inotropic) effect of Ang II on cardiac tissue and cardiomyocytes is not

consistently observed in mammalian hearts. Positive inotropic responses have been

elicited in pigs in vivo (Broome et al., 2001) and in isolated hearts or cardiac tissue

preparations from dogs (Kobayashi et al., 1978), cats (Meulemans et al., 1990), rabbits

(Scott et al., 1992; Watanabe and Endoh, 1998), guinea pigs (Feolde et al., 1993),

hamsters and humans (Moravec et al., 1990). Furthermore, positive inotropic effects of

Ang II are documented in isolated adult cardiomyocytes in rats (Neyses and Vetter,

1989; Delbridge et al., 1995), cats (Petroff et al., 2000), dogs (Cheng et al., 1996),

rabbits (Barry et al., 1995) and hamsters (De Mello, 1998).

In contrast, a positive inotropic effect of Ang II could not be detected in a number of

other studies, including tissue preparations or isolated hearts from rats, rabbits and

human (Baker and Singer, 1988; Wikman-Coffelt et al., 1991; Marano et al., 1997) or

isolated adult cardiomyocytes from rats, guinea pigs, humans, dogs and mice (Ishihata

and Endoh, 1995; Lefroy et al., 1996; Sakurai et al., 2002). Finally, Sekine et al. (1999)

described different inotropic behaviors of isolated neonatal and adult mouse

Chapter IV

153

cardiomyocytes to Ang II, the latter producing a sustained positive inotropic response

in neonate and a maintained negative inotropic response in the adult.

The disparity in findings could be related to differences in pre-existing contractile

status (Li et al., 1994; Meissner et al., 1998), differences in type of tissue utilized

(trabeculae, papillary muscle, whole heart, ventricular or atrial tissue), variable

techniques (tissue strips or rings, isolated perfused heart, working heart,

echocardiography, etc), temperature variations and species- or age-related differences

(Ishihata and Endoh, 1995; De Mello, 1998; Sekine et al., 1999). Indeed it is possible

that the final effect of Ang II on myocardial inotropism results from the complex

combination of several of these factors as well as from the experimental duration of

Ang II exposure. This is particularly important, as duration of Ang II exposure could

elicit different ‘acute’ or ‘chronic’ inotropic responses following activation or

inhibition of Ang II-dependent intracellular signaling pathways in cardiomyocytes.

In contrast to ‘normal heart’ preparations and cardiomyocytes, it is consistently

reported that either the positive inotropic response to acute Ang II exposure is

diminished or the negative inotropic effect is enhanced in various pathological states.

Moravec et al. (1990) demonstrated that in the failing human ventricle and in

cardiomyopathic hamster hearts there was a significant decrease in inotropic response

to Ang II at the cellular level, while Ang II exacerbated contractile dysfunction in rat

papillary muscles after induction of pressure-overload cardiac hypertrophy (Meissner et

al., 1998). Furthermore, Skolnick et al. (1998) reported suppressed responsiveness to

Ang II on ventricular myocytes from infarcted rabbits, while Capasso et al. (1993)

showed negative inotropic effect of Ang II on papillary muscles from infarcted rats.

Ang II was reported to exacerbate contractile dysfunction at the tissue and cellular level

in pacing-induced heart failure in the dog (Cheng et al., 1996) and in isolated rabbit

hearts exposed to ischemic conditions (Mochizuki et al., 1992). Additionally, Xiao and

Allen (2003) demonstrated that AT1 receptor antagonists, applied during ischemia and

reperfusion in isolated rat hearts, protect against myocardial and mechanical damage.

Finally, chronic blockade of the renin-angiotensin system by ACE inhibitors and AT1

receptor blockers in a low-sodium environment converted a positive inotropic response

to a negative inotropic response in rat isolated adult cardiomyocytes (Trongtorsak et al.,

2003). These data suggest that cardiac remodeling in heart disease states may be

Chapter IV

154

associated with changes in the inotropic responsiveness to Ang II of cardiac tissues and

cardiomyocytes (Senzaki et al., 1998). The inotropic effect of acute exposure to Ang II

maybe significantly modified when endogenous levels of Ang II are chronically

elevated. Chronically high in vivo Ang II may suppress or even reverse the inotropic

response to exogenously applied Ang II. In support of this negative modulating effect

exerted by Ang II on cardiac function, progressive increases in cardiac Ang II

production have been reported experimentally and clinically in the course of cardiac

failure (Wollert and Drexler, 1999; Neri Serneri et al., 2001).

1.2.2. Mechanisms for differential inotropic effects of Ang II

on cardiac tissue

Several explanations are advanced to explain the mechanisms of the differential

inotropic effects of Ang II observed on cardiac tissue. Although Ang II could affect

cardiac performance via Ang II-sensitive inter- and intra-cardiac neurons (Horackova

and Armour, 1997), this could not explain the findings relating to isolated hearts and

cardiomyocytes, where the direct action of Ang II in the modulation of contractility

occurs through activation of AT-mediated intracellular signaling pathways. In the heart,

AT1 receptors coexist with the AT2 subtype on both myocyte and non-myocyte cell

populations. The relative proportion in which both subtypes are expressed varies

among cell types, tissues, developmental stages and species and is also affected by

ageing and the extent of cardiac remodeling in pathological responses. This variability

in expression of AT receptor sub-types could determine the positive or negative

inotropic actions of Ang II observed.

As detailed below, the precise mechanisms underlying Ang II-induced and AT-

mediated inotropic action are not well understood, but they are believed to be mediated

by activation of the AT1 receptors and linked with changes in intracellular calcium,

sodium and pH homeostasis in both myocytes and non-myocyte cells. Indeed, Ang II

has been shown to enhance the transcriptional regulation and the activity of key E-C

coupling transporters involved in sodium, calcium and hydrogen homeostasis, such as

the NHE, the NCX, the SERCA2, the RyR and the voltage-operated calcium channels.

Chapter IV

155

1.2.3. Regulation of the sodium/hydrogen exchanger (NHE) by Ang II

It is well recognized that Ang II modulates the activity of the NHE in a variety of

tissues, including the renal proximal tubule (Houillier et al., 1996; Becker et al., 2003),

vascular endothelial cells (Muscella et al., 1999), smooth muscle cells (Vallega et al.,

1988) and blood platelets (Argaman and Livne, 1988). As explained in Chapter I, the

NHE-1 is the most abundant isoform of the sodium/hydrogen exchanger found in the

heart. There is also extensive evidence that Ang II modulates NHE activity in cardiac

myocytes from experiments with isolated cells (Barry et al., 1995; Matsui et al., 1995)

and with intact myocardium (Grace et al., 1996; Camilion de Hurtado et al., 1998).

Furthermore, it was shown that the stimulation of the NHE by Ang II is impaired in

hypertrophied cardiomyocytes (Ito et al., 1997), in infarcted hearts (Skolnick et al.,

1998) and during ischemia and ischemic preconditioning (Xiao and Allen, 2003). It has

been suggested that the net inotropic effect of Ang II on cardiomyocyte performance in

various physiologic and pathologic conditions is partially dependent on NHE

stimulation and activity (Ikenouchi et al., 1994; Barry et al., 1995; Matsui et al., 1995;

Wang and Nygren, 2003; Xiao and Allen, 2003), possibly through activation of PKC-

dependent pathways (Grace et al., 1996; Cingolani et al., 1998; Maly et al., 2002).

Yet, very little is known about the direct modulation of NHE gene expression by Ang II

in the myocardium. NHE mRNA and protein levels were found to be over-expressed in

the rat LV remnant myocardium after infarction in association with an increase in the

activity of the exchanger. In the same study, the AT1 blocker valsartan, and the ACE

inhibitor ramipril, both inhibited upregulation and activity of the NHE post-infarct,

suggesting that the effects were Ang II- and AT1-mediated (Sandmann et al., 2001).

Chapter IV

156

1.2.4. Regulation of the sodium/calcium exchanger (NCX) by Ang II

Ang II has been shown to induce a PKC-mediated increase in the activity of the NCX

(Ballard and Schaffer, 1996). It is also suggested that Ang II can play a direct role in

the modulation of NCX gene expression in the myocardium. Krizanova et al. (1997)

reported that in vivo delivery of Ang I significantly increased the expression of NCX in

rat hearts via production of Ang II. In vivo captopril and candesartan treatments were

shown to reduce NCX mRNA expression in cardiac hypertrophy (Brooks et al., 2000)

and myocardial infarction (Hanatani et al., 1998). However, these in vivo findings

contrast with the in vitro finding of Ju et al. (1996) who reported decreased NCX

mRNA abundance in rat cultured neonatal cardiomyocytes following stimulation with

low or high doses of Ang II. How these conflicting in vivo and in vitro observations can

be reconciled is not clear. Ang II regulation of NCX expression in vivo in the neonatal

heart or in vitro may be different to that occurring in the adult heart or in pathologic

conditions.

Interestingly, recent work from Aiello et al. (2002) has suggested that the Ang II-

dependent modulation of the NCX is mediated by the autocrine action of endothelin-1

released from the myocytes after Ang II stimulation. This hypothesis could be extended

to a paracrine action of endothelin-1 released from fibroblasts acting on the

cardiomyocyte population.

1.2.5. Regulation of SR calcium ATPase (SERCA2) by Ang II

Among the calcium-regulating proteins, the calcium ATPase SERCA2 is generally

reported to be decreased in concentration during cardiac remodeling, independently of

the etiology or the model (Suko et al., 1970; Ito et al., 1974; Lim et al., 1999;

Terracciano et al., 2001; Zhong et al., 2001). Indeed, the diminished expression of

SERCA2 protein levels is considered to be one of the best markers of chronic cardiac

remodeling and failure (Kiss et al., 1995). Very little is known about the direct

modulation of cardiac SERCA2 activity and gene expression by Ang II. Ju et al. (1996)

Chapter IV

157

reported a decrease in SERCA2 mRNA abundance in cultured isolated neonatal rat

cardiomyocytes after treatment with low and high doses of Ang II. However, this effect

was seen neither in adult isolated cardiomyocytes, nor in rat ventricular samples after a

similar treatment, suggesting that angiotensin modulation of calcium signaling in

neonatal and adult myocytes may be different.

Other indirect evidence suggests that Ang II modulates the activity and expression of

the SERCA2 in pathological states. In rats, treatment of volume-overload cardiac

hypertrophy with an AT1 receptor antagonist (TCV-116) restored suppressed

myocardial SERCA2 mRNA levels to control levels (Hashida et al., 1999), while

captopril, an ACE inhibitor, attenuated the reduction of SERCA2 protein and mRNA

expression observed in heart failure after myocardial infarction (Shao et al., 1999).

Similar results were observed in cardiac hypertrophy in aortic-banded rats and guinea

pigs, where treatment with ACE inhibitors or AT1 blockers attenuated the depression of

SERCA2 activity and expression (Boateng et al., 1998; Liu et al., 1999; Takeishi et al.,

1999).

1.2.6. Regulation of the ryanodine receptors RyR by Ang II

In a number of models of experimental cardiac hypertrophy, in cyclically stretched

neonatal cardiomyocyte cultures, and in human heart failure, RyR binding site density

and/or the RyR mRNA expression are reported to be reduced (Brillantes et al., 1992;

Rannou et al., 1996; Cadre et al., 1998; Hashida et al., 1999; Hittinger et al., 1999;

Milnes and MacLeod, 2001). Other studies have reported unchanged (Assayag et al.,

1997) and increased (Arai et al., 1996) levels of RyR expression in senescent hearts and

in compensated cardiac hypertrophy respectively. Little is known, however, about the

direct modulation of RyR gene expression and activity by Ang II in the heart. Ju et al.

(1996) reported an AT1-dependent downregulation of RyR mRNA in isolated neonatal

rat cardiomyocytes after stimulation by Ang II, but there was no evidence of such an

AT1-mediated effect in vivo or with adult isolated cardiomyocytes. Guo et al. (2003)

showed that partial improvement in LV function by AT1 blockers and ACE inhibitors

in infarcted rats was associated with partial reversal of RyR (and SERCA2)

downregulation. Similarly, Hashida et al. (1999) demonstrated that treatment of

Chapter IV

158

volume-overloaded rat hearts with an Ang II receptor antagonist improved left

ventricular function in association with restored levels of RyR (and SERCA2) mRNA.

1.2.7. Regulation of the voltage-operated calcium channels by Ang II

Although it is accepted that Ang II can modulate L-type calcium currents in

cardiomyocytes, the mechanisms of Ang II action remain controversial. Kaibara et al.

(1994) proposed that Ang II can stimulate the L-type calcium channels in adult rabbit

ventricular myocytes through an AT1-mediated activation of the NHE. Aiello and

Cingolan, (2001) showed that this mechanism could be prevented by strong

intracellular calcium buffering and PKC inhibition. However these results were not

confirmed by Ichiyanagi et al. (2002) who reported NHE- and PKC-independent

modulation of L-type calcium currents by Ang II in single adult rabbit ventricular

myocytes using a perforated patch-clamp technique. Finally, De Mello (1998)

demonstrated that species-differences could account for differential modulation of L-

type calcium channels in ventricular myocytes by Ang II. De Mello (1998) also

described the modulation of the inward calcium current in single adult rat and hamster

myocytes by intracellular administration of Ang II and suggested that the peptide was

acting intracellularly, possibly through a PKC-independent pathway in rats. This raises

the possibility that Ang II can modulate the magnitude of calcium transients and

cardiomyocyte contractility through separate intracellular and extracellular pathways.

Interestingly, recent work from Ferron et al. (2003) suggested that an AT1-activated

MAPK kinase-1/2 (MEK1/2) pathway and a separate autocrine endothelin-1 pathway

can upregulate T-type calcium channel expression and density in hypertrophic isolated

cardiomyocytes from aortic-banded rats.

Chapter IV

159

1.2.8. Integrating the mechanisms of Ang II modulation of

cardiomyocyte inotropy

Taken together, these data would suggest the scheme summarized in Figures IV-1 and

IV-2, relating to Ang II modulation of cardiomyocyte contractility. The downstream

effect of acute Ang II exposure associated with protein kinase C (PKC) stimulation

appears to lead to:

a) increased inward calcium current, through activation of the voltage-operated

channels and the NCX (Figure IV-1, ‘Pathway’ A); this process would promote SR

calcium-induced calcium release, yielding a positive inotropic response associated

with a prolonged calcium transient;

b) activation of the NHE, leading to sodium load and consequently to an increase in

the activity of the NCX working in the reverse mode (calcium ‘in’) (Figure IV-1,

‘Pathway’ B);

c) activation of the NCX working in the forward mode (calcium ‘out’), promoting

arrhythmogenic activity as calcium efflux is associated with inward depolarizing

current (Figure IV-1, ‘Pathway’ C). However, sodium load could suppress the NCX

to operation in the calcium extrusion mode during relaxation. This would lead to

cytoplasmic and SR calcium overload and prolonged calcium transient and

relaxation time.

d) proton extrusion inducing alkalinization of the cytoplasm and positive inotropism

due to a change in the myofilament calcium sensitivity (Figure IV-1, ‘Pathway’ D)

(Ballard and Schaffer, 1996; Mattiazzi et al., 1997; Talukder and Endoh, 1997; De

Mello, 1998; Watanabe and Endoh, 1998; Petroff et al., 2000).

In a chronic state of Ang II stimulation, differential gene expression and

metabolic/mitochondrial remodeling could change the mechanical response of the cell

to Ang II. In a transition from acute to chronic exposure to Ang II, alkalinization

associated with increased NHE activity can stimulate gene and protein expression,

leading to cell growth and proliferation (Grinstein et al., 1989; Cingolani, 1999) (Figure

IV-2, A.). Alkalosis can also disrupt mitochondrial metabolism and reduce ATP

production (Figure IV-2, B.), decreasing the energy available for contraction, thus

Chapter IV

160

inducing a negative inotropic response. Through the mitochondrion, alkalinization can

promote apoptosis (Khaled et al., 2001), a phenomenon that could ultimately shift the

balance between cell growth and cell death (Figure IV-2, C.).

In chronic states, Ang II acts directly as a growth factor on cardiomyocytes, increasing

protein synthesis and also reprogramming gene expression. For instance, Ang II was

shown to induce the expression of immediate early genes (c-Fos and c-Jun) and

endothelin-1, a potent modulator of cardiomyocyte inotropy, through a PKC-mediated

activation of JNK (Figure IV-2, D. and E.). Autocrine and non myocyte-dependent

paracrine actions of endothelin-1 and other neuro-humoral factors such as

prostaglandins (Meulemans et al., 1990), could enhance and activate further

intracellular signaling pathways leading to a more sustained differential gene

expression and cardiomyocyte remodeling (Figure IV-2, E.). Ang II-dependent re-

expression of slow-twitch fetal isoforms of contractile proteins, such as α-skeletal actin

or β-myosin heavy chain, would be associated with a decreased cross-bridge cycling

rate and slower contraction kinetics (Lecarpentier et al., 1987; Kim et al., 1995;

Clement et al., 2001) (Figure IV-2, F.). In addition, differential gene expression of key

‘players’ in calcium homeostasis could shift calcium cycling from an intracellular

pathway toward an extracellular one (downregulation of SERCA2 and possible

upregulation of the NCX), depleting SR calcium stores and possibly decreasing the

amplitude of the calcium transients (Figure IV-2). The negative inotropic effects of

Ang II were previously associated either with no changes in the amplitude of the

calcium transient in mouse cardiomyocytes (Sakurai et al., 2002) or with decreased

calcium transient in isolated rat hearts (Wikman-Coffelt et al., 1991), in infarct-remnant

rabbit ventricular myocytes (Skolnick et al., 1998) and in hypertrophied rat papillary

muscles (Meissner et al., 1998).

Finally, Ang II may modulate positive and negative inotropic effects through oxidative

stress. Indeed, Ang II can activate phospholipase D (PLD) and NADPH oxidase,

leading to production of superoxide anions in cardiomyocytes and vascular smooth

muscle cells (Figure IV-1, E.) (Sadoshima and Izumo, 1993; Touyz and Schiffrin,

2001; Privratsky et al., 2003). The production of superoxide induces activation of the

NHE in adult rat ventricular myocytes (Snabaitis et al., 2002). When this activation is

Chapter IV

161

mediated by the MAPK, Erk the final inotropic response to Ang II is positive

(Figure IV-2, G), whereas an activation of the p38 MAPK would lead to negative

inotropism (Khaled et al., 2001), possibly secondary to a shift in gene expression and

mitochondrial damage (Figure IV-2, H).

These hypotheses suggest that Ang II can modulate different contractile responses

according to the duration of the stimulus and the pre-existing physiological (or

pathological) status of the myocardium. At the present time, the chronic effects of

elevated intra-cardiac levels of Ang II on cardiomyocyte function are not well

characterized. This is essentially due to the lack of appropriate models of Ang II-

induced heart failure where hypertension and mechanical stretch do not confound the

chronic effects of Ang II on cardiomyocyte function. For this reason, the use of

transgenic animals to address the question of whether upregulation of the cardiac RAS

alone is sufficient to induce cardiomyocyte dysfunction in the absence of confounding

hypertension is of significant interest.

1.3. Diabetic cardiomyopathy and cardiomyocyte contractility

1.3.1. Cardiac dysfunction related to diabetic cardiomyopathy

As discussed in Chapter I, cardiovascular disease is a leading cause of mortality and

morbidity in type 2 diabetes. Clinically, diabetic cardiomyopathy is frequently

associated with systolic and diastolic dysfunction which is not directly attributable to

microvascular disease, hypertension or obesity. Furthermore, in clinical studies,

detectable cardiac dysfunction is reported to occur as early as the glucose intolerance

phase - that is with the development of hyperglycemia and hyperinsulinemia following

association with insulin resistance (Celentano et al., 1995). In type-2 diabetic

cardiomyopathy, diastolic dysfunction usually precedes impaired systolic dysfunction

(Galderisi et al., 1991).

Chapter IV

162

The mechanisms of cardiac and cardiomyocyte dysfunction underlying diabetic

cardiomyopathy are relatively well characterized in animal models of type 1 diabetes.

This is due to a preponderance of studies conducted using chemically-induced models

of insulin-deficient diabetes, such as the STZ-treated diabetic rat. Another useful model

of type 1 diabetes is the spontaneously diabetic Bio-Breeding/Worcester (BB/W) rat, a

strain in which diabetes occurs spontaneously and closely resembles IDDM in humans

(Miller, 1983; Malhotra et al., 1985). Among the mouse models, the OVE26 transgenic

strain exhibits diabetic cardiomyopathy and complications associated with type 1

diabetes. This mouse over-expresses calmodulin in pancreatic β-cells, a feature that

causes type 1 diabetes in this strain. Duan et al. (2003) reported decreased

cardiomyocyte peak shortening, maximal velocity of shortening and lengthening,

prolonged time of peak contraction and reduced IGF-1 response in these mice.

In general, in type 1 diabetes abnormal cardiomyocyte E-C coupling includes

prolonged action potentials, reduced diastolic and peak systolic intracellular calcium

concentrations, slower rate of decay and prolongation of the calcium transient, negative

inotropy and prolonged time to peak shortening (Noda et al., 1993; Jourdon and

Feuvray, 1993; Yu et al., 1994; Lagadig-Gossmann et al., 1996). Kotsanas et al. (2000)

reported delayed onset of shortening and a depressed calcium-frequency relationship in

cardiomyocytes from STZ-diabetic rats. In addition Ren and Davidoff (1997) suggested

that long term diabetes (8 weeks) but not short-term (4-6 days) is necessary to cause

negative inotropy in STZ rat myocytes. These pathological E-C coupling characteristics

are generally accompanied by a shift in myosin isoenzymes (Hofmann et al., 1995; Kita

et al., 1996). Similar alterations in contractile function are observed in papillary

muscles of STZ-treated rats (Yu and McNeill, 1991; Ishikawa et al., 1999; Ren et al.,

1999). In the latter case a blunted inotropic response to increasing stimulus frequency

was also observed. These results were confirmed in isolated perfused hearts (Zhong et

al., 2001). Interestingly, Ren et al. (1999) demonstrated an increase in cardiac and

cardiomyocyte inotropy after administration of insulin in STZ-treated but not in control

rat hearts and cardiomyocytes.

Although the adult STZ-injected rat is primarily utilized as a model of type 1 diabetes,

some studies employ the STZ-treated rats as a model of type 2 diabetes with an insulin

Chapter IV

163

secretory defect. This state is obtained by an injection of STZ in neonatal rather than

adult rats. While this model cannot be considered as a true representation of the type 2

diabetic condition, a marked glucose intolerance due to insulin resistance that is

associated with normal plasma fasting and non-fasting glucose values is observed

(Schaffer et al., 1985; Schaffer and Wilson, 1993). Differences in myocardial

contractility in rats injected with STZ at both early ‘type 2’ and late ‘type 1’ stages in

development were studied by Banyasz et al. (1996). They reported decreased velocity

of contraction and relaxation in type 1, but not in ‘type 2’ isolated ventricular muscles

and a decreased sensitivity to catecholamines in both types of diabetes.

Type 2 diabetes is considerably more prevalent than type 1 diabetes, but understanding

the pathogenesis of cardiomyopathy in the type 2 condition is complicated by

numerous co-morbidities in both humans and animals (i.e. hypertension, obesity,

hyperinsulinemia, hyperglycemia and dislipidemia). For these reasons the study of

cardiac and cardiomyocyte function in type 2 diabetes has received far less research

attention. Sucrose-fed rats have been used to study the early stage of type 2 diabetes

(Pierce et al., 1989; Pagliassotti et al., 1996), which is characterized by whole-body

insulin resistance and hyperinsulinemia in humans. Dutta et al. (2001) reported slower

rate of shortening and re-lengthening, slower cytosolic calcium clearing and depressed

peak shortening (albeit inconsistent) in cardiomyocytes from sucrose-fed rats. These

changes in cardiomyocyte function were associated with prolonged calcium transients.

In genetic models of type 2 diabetes (i.e. non insulin-deficient) indirect evidence of

cardiomyocyte and cardiac dysfunction have been observed. The insulin-resistant and

diabetic Wistar Bonn/Kobori (WBN/Kob) rat showed impaired L-type calcium channel

response to β-adrenergic stimulation and a prolonged action potential due to depressed

transient outward potassium current (Ito) (Tsuchida et al., 1994; Tsuchida and

Watajima, 1997). Similarly, spontaneously diabetic KK mice showed an increase in the

action potential duration, possibly driven by a suppression of potassium current, such

as Ito and IK1 (Aomine and Yamato, 2000). Finally, very little is known about cardiac

and cardiomyocyte function in recently developed GLUT4 deficient mouse models.

Abel et al. (1999) reported that isovolumetric contractile performance is preserved in

isolated perfused hearts with selective cardiac deletion of GLUT4 protein (G4H-/-

Chapter IV

164

mice). Such hearts developed diastolic dysfunction only after low-flow ischemia ex

vivo and in response to catecholamines in vivo (Tian and Abel, 2001).

No studies on isolated cardiomyocyte contractility have been performed by these

investigators using previously described GLUT4 deficient models (Chapter III Section

1.3.1). Virtually all the information available relating to altered E-C coupling status in

diabetic cardiomyopathy (altered ion homeostasis in particular) has been gleaned from

type 1 models as is evident from this discussion and as considered below.

1.3.2. Altered pH homeostasis in diabetic cardiomyopathy

Pierce et al. (1990) demonstrated depressed cardiomyocyte NHE activity in

sarcolemmal membrane vesicles isolated from STZ-induced (type 1) diabetic rats, a

phenomenon that was paralleled by a reduction in activity of the NCX and sodium-

potassium ATPase (NaKATPase). Le Prigent et al. (1997) observed that diabetes

significantly decreased acid efflux through the NHE in ventricular myocytes from STZ

diabetic rats. This phenomenon was associated with decreased intracellular calcium and

sodium concentrations (Noda et al., 1992, Imahashi et al., 1998), but no changes in

intracellular pH, or proton buffering capacity were observed (Noda et al., 1992, Le

Prigent et al., 1997). Dyck and Lopaschuk (1998) suggested that such depressed NHE

activity could be compensatory to decreased rates of glycolysis and reduced production

of acid charges (H+). Inhibition of the NHE would lead to decreased intracellular

calcium and sodium concentrations, protecting diabetic hearts from acidosis-dependent

ischemic injuries and preserving ATP levels (Ramasamy and Schaefer, 1999; Allen and

Xiao, 2003). However, some studies reported no changes in NHE mRNA or protein

concentrations in STZ diabetic hearts (Dyck and Lopaschuk, 1998; Hileeto et al.,

2002).

Chapter IV

165

1.3.3. Altered calcium homeostasis in diabetic cardiomyopathy

It is generally agreed that calcium homeostasis is affected in diabetic cardiomyopathy.

Significant insulin-dependent depression in sodium-calcium exchange, resulting from

reduced activity of the NCX exchanger, was observed in cardiac sarcolemmal vesicles,

neonatal and adult cardiomyocytes isolated from STZ-injected rats (Pierce et al., 1990;

Chattou et al., 1999; Hattori et al., 2000; Choi et al., 2002). This reduced activity was

associated with a decrease in NCX protein levels (Hattori et al., 2000; Choi et al.,

2002), with and without a decrease in NCX mRNA expression (Hattori et al., 2000;

Schaffer et al., 1997). Golfman et al. (1998) observed reduced activity of the NCX in

sarcolemmal vesicles from alloxane-induced diabetic rats. The same authors also

observed a significant increase in NCX mRNA abundance at 3 weeks after induction of

diabetes. However, these mRNA levels returned to control by 5 weeks of diabetes.

Finally, Duan et al. (2003) demonstrated a reduced NCX activity and protein content in

cardiomyocytes isolated from OVE26 (type 1) diabetic mice.

Very little is known about cardiac NCX regulation in type 2 diabetes or in the pre-

diabetic insulin-resistant state. Dutta et al. (2002) observed unchanged NCX activity

and in mRNA expression in isolated rat ventricular myocytes cultured in a medium

containing high concentrations of glucose. There appears to be no information available

about how NCX activity and expression may be modified in mouse models of non-

insulin dependent diabetic or pre-diabetic states, such as the GLUT4 knock-out mouse

models described in Chapter III.

Altered SERCA protein and mRNA expression associated with contractile and SR

dysfunction have been observed in STZ diabetic rat hearts. Teshima et al. (2000)

reported decreased SERCA2 and RyR mRNA and protein levels in rats after 3 weeks of

STZ treatment. This downregulation of SERCA2 was associated with a downregulation

of RyR mRNA after 12 weeks of treatment. Zhong et al. (2001) reported a similar

decrease in cardiac SERCA2 protein content after 6 weeks, but not after 4 weeks, of

STZ treatment in rats. These general trends were confirmed by Kim et al. (2001) and

Choi et al. (2002), who reported decreased SERCA2, NCX and RyR protein levels in

Chapter IV

166

STZ-treated rats. This decrease in protein content was associated with impaired rates of

SR calcium release and sequestration. More recently, Duan et al. (2003) reported

similar trends of decreased SERCA2 protein and mRNA levels, associated with

decreased SERCA2 activity in type 1 diabetic OVE26 mice.

In isolated rat ventricular myocytes cultured in a medium containing high

concentrations of glucose, Dutta et al. (2002) failed to detect any alteration in SERCA2

protein or mRNA expression. It appears that no data are available which describe

SERCA2 or RyR protein and mRNA expression profiles in type 2 diabetic

cardiomyopathy.

1.4. Insulin resistant cardiomyopathy: evidence of activation of the

renin-angiotensin system

There is strong clinical evidence that non insulin-dependent type 2 diabetic patients

suffer from a cardiomyopathy and left ventricular dysfunction independent of co-

morbidities such as hypertension and coronary artery diseases (reviewed in Fang et al.,

2004). It is also well documented that stimulation of the local intra-cardiac RAS,

together with endothelin-1 production and enhanced NHE activity characterize the

diabetic heart (Sechi et al., 1994; Hileeto et al., 2002). In diabetic hearts, local effects

of Ang II appear to be modulated by PKC- and/or NADPH oxidase-dependent

pathways (Malhotra et al., 1997; Privratsky et al., 2003). Furthermore, IGF-1 attenuates

Ang II- and p53-mediated apoptosis and oxidative stress (Kajstura et al., 2001;

Privratsky et al., 2003) and AT1 receptor blockade protects the heart from the

development of cellular mechanical alterations typically associated with diabetes

(Raimondi et al., 2004). This evidence would suggest that many of the mechanical

disturbances observed in the diabetic heart at the tissue and cellular levels can be

accounted for by accumulating tissue Ang II levels, leading to cardiac dysfunction

through oxidative stress (Privratsky et al., 2003). Thus a comparative study of the

alterations in cardiomyocyte E-C coupling induced by the chronic exposure to Ang II,

Chapter IV

167

and by insulin resistance, in the absence of confounding hypertension, is of

considerable interest.

1.5. Aims

The present study was designed to comparatively evaluate the cardiomyocyte

functional effects of experimentally induced in vivo 1) chronic elevation of Ang II

levels and 2) impaired GLUT4-mediated glucose uptake. Specifically, using adult

myocytes isolated from the left ventricle of TG1306/1R and GLUT4-KO mice, the

experimental aims were:

1. To characterize myocyte intrinsic basal contractile function in their pathologies,

and to determine how contractility is modified by ageing and disease

progression.

2. To evaluate age-dependent alterations in the gene expression of the

sodium/calcium exchanger (NCX1.1) and sodium/hydrogen exchanger (NHE-

1), as well as the calcium release channel ryanodine receptor (RyR2). The

protein levels of the calcium ATPase SERCA2 were also evaluated at 15-20

weeks of age.

The goal of this study was to clarify whether cardiac and cardiomyocyte remodeling

induced by chronic Ang II overproduction or GLUT4 transporter suppression in vivo

could have a primary and long-term impact on myocyte function, independently of

hemodynamic changes. It was hypothesized that cardiomyocyte contractile dysfunction

would be evident in both mouse models and it would be associated with differential

gene (or protein) expression of key players in calcium, sodium and hydrogen

homeostasis.

Chapter IV

168

2. METHODS

2.1. Cardiomyocyte contractility

2.1.1. Myocyte preparation and recording

Myocytes were prepared from mouse hearts by the procedures detailed in Chapter II,

Section 5. Measurement of single cardiomyocyte contractility was performed by using

a rapid imaging system, incorporating a photodiode array line scan camera coupled to a

digitizing-equipped workstation, as described in Chapter II, Section 5. A range of

normalized parameters were automatically computed for each contraction cycle and

recorded for comparative evaluation of inotropic status as described earlier (Figure II-

4).

2.1.2. Experimental groups

To investigate the age-dependent effects of hypertrophic remodeling on cardiomyocyte

function, 15-20 and 35-40 week old male transgenic and knock-out mice were studied

and compared with age-matched control littermates. Experimental groups were defined

according to animal age and genotype. Isolated left ventricular cardiomyocytes from

the following 8 groups were studied:

1. 15-20 week TG: n=15 (from 5 different animals).

2. 15-20 week WT: n=17 (from 5 different animals).

3. 35-40 week TG: n=15 (from 5 different animals).

4. 35-40 week WT: n=16 (from 5 different animals).

5. 15-20 week LLC: n=16 (from 5 different animals).

6. 15-20 week LL: n=15 (from 5 different animals).

Chapter IV

169

7. 35-40 week LLC: n=16 (from 5 different animals).

8. 35-40 week LL: n=15 (from 5 different animals).

2.1.3. Recording protocols and data analysis for cardiomyocyte

contractility

Measurements of cardiomyocyte contractility were carried out at 36.0±0.5oC, with a

superfusate flow rate of 1.8 ml/min, extracellular Ca2+ 2mM and pH 7.4 (Delbridge et

al., 1989) (details of buffer are reported in Chapter II, Section 5). Cardiomyocytes were

initially stimulated to contract at 3Hz and cell performance was allowed to achieve

steady state before the test recording period commenced (5 minutes allowed).

Cardiomyocyte contractility was then monitored for 7-8 minutes following the protocol

illustrated in Figure IV-3. The protocol was designed to evaluate cardiomyocyte

contractile performance throughout an ascending ramp of four different frequencies

(1.5Hz – 3Hz – 4Hz – 5Hz) before returning to the basal rate at 3Hz. Thus, the pre- and

post-ramp performance levels could be compared to assess myocyte functional stability

during the test period.

Thirty contractile cycles were averaged at the end of each frequency step within the

ramp for statistical evaluation. To standardize the recording period relative to frequency

transition, 100 contractile cycles for 1.5Hz and 3Hz and 50 contractile cycles for 4Hz

and 5Hz were specified for the frequency ramp protocol. In addition, to assess myocyte

stability during the test period 100 contractile cycles at 3Hz before the frequency ramp

and at the end of the frequency ramp were averaged for comparison. Mean values were

determined for each experimental group using the average values computed for each

myocyte at each frequency.

Myocytes were included in analysis if the following criteria were satisfied: 1) rod-

shaped morphology with clear striation patterns; 2) calcium tolerant (see Chapter II,

Section 4); 3) quiescent in the absence of electrical stimulation; 4) stable mechanical

behavior at 3Hz during the run-in period. Furthermore, because cell size could

potentially introduce additional physiological and mechanical variables to the

functional study, myocytes were selected to ensure experimental groups were

Chapter IV

170

comprised of equivalently sized cells. Thus, myocyte dimensions were not different for

the eight experimental groups tested (mean length: 129.0 ± 13.0 μm; mean width: 29.0

± 3.0 μm), although it is important to note as demonstrated by data presented in

Chapter III (Table III-1 and III-2) that there were significant size differences in the

general cell populations comprising these groups.

2.2. RNA and protein extraction and quantification

2.2.1. Procedures

Total RNA was extracted from mouse hearts and expression of NCX1, NHE-1 and

RyR2 mRNA was assessed by semi-quantitative RT-PCR on 15-20 and 35-40 week old

ventricles using the procedures detailed in Chapter II, Section 7. The procedures for

protein extraction and Western blotting for evaluation of SERCA2 expression are

described in Chapter II, Section 6. For immunodetection of SERCA2, membranes were

incubated with diluted goat anti-mouse SERCA2 polyclonal antibodies (Santa Cruz

Biotechnology, product C-20) overnight (4oC). After washing, blots were incubated

with peroxidase-conjugated secondary anti-goat IgG antibodies at room temperature

(Santa Cruz Biotechnology).

2.2.2 Experimental groups

For the gene and protein expression investigations described above, 15-20 and 35-40

week old male transgenic mice were studied and compared with age-matched control

littermates.

Eight experimental groups were defined according to animal age and genotype. Left

ventricular tissues from the following 8 groups were studied:

Chapter IV

171

1. 15-20 week TG: n=10 hearts for mRNA extraction and n=10 hearts for protein

extraction.

2. 15-20 week WT: n=10 hearts for mRNA extraction and n=10 hearts for protein

extraction.

3. 35-40 week TG: n=10 hearts for mRNA extraction.

4. 35-40 week WT: n=10 hearts for mRNA extraction.

5. 15-20 week LLC: n=10 hearts for mRNA extraction and n=10 hearts for

protein extraction.

6. 15-20 week LL: n=10 hearts for mRNA extraction and n=10 hearts for protein

extraction.

7. 35-40 week LLC: n=10 hearts for mRNA extraction.

8. 35-40 week LL: n=10 hearts for mRNA extraction.

2.3. Statistical considerations and presentation of the results

As previously described, for myocyte contractility experiments mean group values

were determined using the average parameter values computed for each myocyte at

each frequency. For each experimental group, the ‘n’ values reported in the Results

Section reflect the total number of myocytes recorded. To ensure that the intra-heart

and inter-heart variability in myocyte performance was consistent between each

experimental group, an equal number of myocytes per heart were incorporated in each

group. Results are expressed as mean ± SEM. A multiple-way ANOVA with repeated

measures was used to statistically evaluate the contraction frequency staircase and to

evaluate possible interactions between the genotype and the age for each contraction

parameter. A subset of data comparing age-matched transgenic (or knock-out) mice and

their wild-type littermates were analyzed by one-way ANOVA and are presented for

the pacing frequency of 5 Hz, which is the pacing value closest to physiological range

in the mouse.

For the gene and protein expression data, a multiple-way ANOVA was used to

statistically evaluate the expression shifts and to identify possible interactions between

genotype and age.

Chapter IV

172

For all data, significant differences were identified at p<0.05. Legends of Figures and

Tables reporting specific results state the type of statistical test utilized and the degree

of significance.

Chapter IV

173

3. RESULTS

3.1. Stability of contractile performance

To ensure that only myocytes which exhibited stable contractile function were included

in the experimental groups, the performance of cells before and after applying the ramp

protocol was evaluated. Comparison of the contractile parameters prior to the

frequency ramp and at the end of the experiment reveals that no differences were

observed for these values in any of the TG1306/1R or GLUT4-KO experimental groups

(Figures IV-4 and 5). This confirms that the cells remain functional and responsive in

their contractile performance through the duration of the experiment and that changes

in pacing elicit physiologic homeostatic responses. Furthermore, there were no

statistical changes in cardiomyocyte length (Lo) over the experiment duration,

indicating that changes in frequency were not associated with a change in diastolic

status in contracting cardiomyocytes (data not shown).

3.2. Impaired cardiomyocyte contractility in Ang II-overproducing

transgenic mice

3.2.1. Decreased inotropy, lusitropy and prolonged cycle time at 5 Hz in TG

Examination of the specific alterations in contractile cycle parameters at near

physiologic range (5Hz) showed that at both ages studied, TG myocytes exhibited

significant functional impairment. Both rates of shortening and lengthening (MRS and

MRL) were reduced in the TG myocytes relative to WT controls at 5 Hz (Table IV-1).

The maximal rates of shortening were reduced by 15% and 35% in 15-20 and 35-40

week old TG myocyte groups respectively, while the maximal rate of lengthening was

Chapter IV

174

reduced by ~30% at both ages. This is indicative of a reduction in contraction

(inotropy) and relaxation (lusitropy) performance in the TG cardiomyocytes,

independent of the age. The slowest contraction kinetics were observed in the group of

older TG myocytes.

Genotype-dependent prolongation effects on contraction cycle timing events were also

evident at both ages: the latency (To), the shortening (Tm -To) and lengthening (Tf -Tm)

periods and total cycle (Tf -To) times were significantly longer in TG myocytes when

compared to age-matched WT (Table IV-1). All contraction cycle timing parameters in

15-20 week old TG myocytes were approximately 15-25% longer than WT. The same

parameters were 35-40% longer than WT in 35-40 week old TG myocytes. The total

cycle time (Tf) including the latency was also significantly prolonged in TG

cardiomyocytes at both ages. These data suggest that the timing of events in the

contractile cycle is significantly modified in TG cells, independent of age. The most

prolonged period of E-C coupling latency (To) was observed in the group of older TG

cells (+40%).

A particularly surprising finding was that, despite these significant alterations in rate

and timing parameters, the maximal shortening (%S) attained by TG myocytes was not

diminished relative to age-matched WT myocytes.

3.2.2. Age- and frequency-dependent alterations in cardiomyocyte function

A more extensive comparison of TG and WT myocyte function was undertaken by

analysis of data obtained at both ages for all frequencies using a multiple-way ANOVA

approach (Table IV-3; Figures IV-6 and IV-7). This extended analysis demonstrated

that all contractile parameters, except %S, exhibited significant genotype specific

shifts, as identified above in relation to the 5Hz data. A negative frequency relationship

was observed for all shortening and lengthening performance parameters (%S, MRS

and MRL) in WT and TG myocytes at both ages (Figure IV-6). In 15-20 week WT

myocytes, the overall shift from 1.5 to 5 Hz produced an approximate 45% reduction in

the magnitude of the maximum cell shortening (%S) and a reduction of 30% in the

magnitude of the rates of shortening and lengthening (MRS and MRL). A similar trend

Chapter IV

175

was observed in the other experimental groups, suggesting that the frequency-

dependent performance modulation was similar for the WT and TG groups, at the 2

developmental stages.

A negative frequency ‘staircase’ was also observed for all contraction cycle timing

parameters at both ages, except for the latency (To), which exhibited a modest but

positive frequency relationship (Figure IV-7). That is, E-C coupling delay increased at

higher frequency by ~15% for all groups. A similar trend was observed for all

experimental groups, thus no significant frequency-dependent shift in the magnitude of

contractile cycle timing was observed between the WT and TG groups.

Myocytes from older animals, both WT and TG, exhibited reduced maximal shortening

(%S) and rate of shortening (MRS) (Figures IV-6, Panel A and B), together with an

abbreviated duration of total cycle time (Tf -To) and lengthening period (Tf-Tm) relative

to their younger counterparts (Figure IV-7, Panels B and D). These findings

demonstrate that in these mouse myocytes, ageing per se regardless of genetic type, is

associated with decreased shortening activity and abbreviated contractile cycle and

lengthening periods. Interestingly, the E-C coupling latency (To), the period of

shortening (Tm-To) and the rate of lengthening (MRL) exhibited significant genotype-

specific alterations with ageing (i.e. showed a statistical age-genotype interaction).

Thus, in the WT, MRL showed an age-dependent reduction which was not observed in

the TG, where this parameter was already relatively depressed even in young animals

(Figure IV-6, Panel C). Similarly, while the shortening period (Tm-To) was decreased

with age in the WT, minimal alteration was evident with ageing in the TG myocytes.

The large increase in excitation-contraction coupling latency (To) seen in the older TG

myocytes was not apparent in the myocytes from the older WT.

3.2.3. Differential gene and protein expression profiles

The expression of calcium and proton handling transporters in the hearts of 15-20 and

35-40 week old TG and WT mice was investigated. Protein analysis showed an ~5-fold

downregulation of the SR calcium pump SERCA2 in the hearts of 15-20 week old TG

when compared with age-matched WT (Figure IV-8). This was in association with

Chapter IV

176

upregulation of both the NCX1.1 and the NHE-1 mRNA levels in the TG myocytes.

The difference in expression between TG and WT of the NCX1.1 was less marked at

15-20 weeks. Ageing was also associated with a general downregulation of expression

of the NHE-1 in both TG and WT. No significant differences were observed in the

expression of the ryanodine receptor RyR2 mRNA between WT or TG in younger or

older animals.

3.3. Impaired cardiomyocyte contractility in GLUT4-deficient mice

3.3.1. Decreased inotropy, lusitropy and prolonged cycle timing at 5 Hz

Analysis of the specific changes in contractile cycle parameters at 5Hz showed that at

both ages LLC cardiomyocytes exhibited significant impairment of contractile

performance and cycle timing. Maximal rates of shortening and lengthening (MRS and

MRL) were 20-25% reduced in the myocytes of both 15-20 and 35-40 week LLC

relative to LL at 5Hz (Table IV-2). The maximum shortening (%S) attained by LLC

myocytes was 25-35% smaller when compared to age-matched LL myocytes. This is

indicative of a reduction in inotropic and lusitropic performance in the LLC

cardiomyocytes, independent of age.

Genotype-dependent prolongation effects on contraction cycle timing were evident at

both ages for the latency (To) and cycle time (Tf). Tf was prolonged approximately 10-

15% in LLC. This result was mainly influenced by the delay of To in LLC

cardiomyocytes, as the total cycle time (Tf -To) and the duration of lengthening (Tf -Tm)

were not significantly changed in LLC myocytes when compared to age-matched LL at

5 Hz (Table IV-2). Interestingly, the shortening time (Tm-To) was significantly

extended (+15%) in LLC myocytes at 35-40 weeks when compared with age-matched

LL. The largest delay in E-C coupling (To) was observed in the group of younger LLC

cells, where latency was approximately 50% longer than LL. These data suggest that

the timing of events in the contractile cycle is relatively disturbed in LLC cells when

compared to LL at 5Hz.

Chapter IV

177

3.3.2. Age- and frequency-dependent alterations in cardiomyocyte function

Analysis of the data using the more robust multiple-way ANOVA approach

demonstrated that all contractile parameters exhibited genotypic-dependent shifts

(Table IV-4). In particular, when the pooled frequency data were considered, a

significant reduction in some cycle time parameters for LLC myocytes was detected -

an effect not evident when the 5Hz dataset alone was evaluated by one-way ANOVA.

This is the case for the total cycle time (Tf -To), the duration of lengthening (Tf -Tm), as

well as the shortening time (Tm -To) in younger LLC (Figure IV-10). A negative

frequency relationship was observed for all shortening and lengthening performance

parameters (%S, MRS and MRL) in LL and LLC myocytes at both ages. In 15-20

week LL myocytes, the overall shift from 1.5 to 5 Hz produced an approximate 30%

reduction in the magnitude of the maximum cell shortening (%S) and a reduction of

15% in the magnitude of the rates of shortening and lengthening (MRS and MRL)

(Figure IV-9). A similar trend was observed in all the experimental groups, suggesting

that the frequency-dependent performance modulation was similar for the LL and LLC

groups at the two developmental stages. A negative frequency ‘staircase’ was also

observed for all contraction cycle timing parameters at both ages, except for the latency

(To), which exhibited a positive frequency relationship (Figure IV-10). A similar trend

was observed for all experimental groups, thus no significant frequency-dependent shift

in the magnitude of contractile cycle timing was observed between the LL and LLC

groups.

Both LL and LLC myocytes from older animals exhibited increased maximal rate of

lengthening (MRL). This finding demonstrates that in these myocytes, ageing per se

regardless of genetic type, is associated with accelerated cardiomyocyte relaxation

performance. Genotype-specific alterations with ageing were observed for the time of

latency (To) and the shortening and lengthening periods (Tm -To, Tf -Tm) between LL

and LLC myocytes (Figures IV-10). In the LL, the myocytes shortening period (Tm-To)

showed an age-dependent reduction of greater magnitude when compared to LLC

myocytes. The opposite was observed for the lengthening period (Tf -Tm), where the

age-dependent reduction was greater in LLC myocytes. The large delay in E-C

Chapter IV

178

coupling latency (To) seen in the younger LLC myocytes was less marked in the older

LLC. The WT showed a similar pattern of age-related latency shift, but the differential

age effect was modest.

A particularly surprising finding was that, despite genotype-dependent alterations in

maximal shortening and rate of shortening (%S and MRS) in LLC myocytes, these

parameters did not exhibit significant age dependency in this myocyte population

(Figure IV-9).

3.3.3. Differential gene and protein expression profiles

Protein analysis showed a ~2-fold downregulation of the sarcoplasmic reticulum (SR)

calcium pump SERCA2 in the hearts of 15-20 week old LLC mice, in association with

downregulation of the sodium-calcium exchanger NCX1.1 and an upregulation of the

sodium-hydrogen exchanger NHE-1 mRNAs (Figure IV-11). Ageing caused a general

downregulation of the expression of the NHE-1. Finally, a significant downregulation

was observed in the expression of the ryanodine receptor RyR2 mRNA between LL

and LLC mice at both ages.

3.3.4. After-contraction in LLC cardiomyocytes at 1.5 Hz.

At the pacing frequency of 1.5Hz it was not unusual to observe unstable contractile

behavior associated with non-stimulated after-contractions in 35-40 week old LLC

cardiomyocytes (Figure IV-12). These after-contractions were generated in the early

post relaxation phase. Relative to stimulated twitches, they were reduced in amplitude

and were not evident at higher frequencies (3Hz to 5Hz).

Chapter IV

179

4. DISCUSSION

4.1. Decreased contractile performance in TG cardiomyocytes

The present study demonstrates that long-term elevated expression of cardiac Ang II

has a detrimental effect on E-C coupling at the cardiomyocyte level, even when there is

no elevation of afterload. This ‘functional remodeling’ is marked by a decrease in the

maximum rates of shortening and lengthening (MRS and MRL) in cardiomyocytes

from 15-20 and 35-40 week old TG (Table IV-1 and Figure IV-6). Assuming that

differential gene expression of the transgene and consequent cardiac-specific

overproduction of Ang II should occur at birth, if not in utero, this phenomenon could

be related to early changes in Ang II-induced gene and protein expression, leading

progressively to homeostatic disturbance in regulation of calcium and other ions and

encroachment of the contractile reserve. Slower rates of shortening and lengthening, as

observed in TG cardiomyocytes, may reflect a shift in fetal myosin and actin

isoenzymes, which has been observed in other models of cardiac hypertrophy in mice

(Dorn et al., 1994). Indeed, it has been previously established in this model that Ang II

chronically stimulates re-expression of fetal low-twitch alpha-skeletal and smooth

muscle actin in 20 week old TG mice (Clement et al., 2001). Changes in myosin

isoforms from α to β in the ventricle of small animals, such as the mouse, could reduce

ATPase activity and maximum shortening velocity, concomitant with increased

economy of isometric force development and decreased ATP utilization per gram of

tension (Hasenfuss et al., 1991).

A slower contraction rate observed in TG cardiomyocytes could also reflect differential

myofilament responsiveness to calcium ions (myofilament sensitivity). Indeed, new

evidence suggests that p38 MAPK activation mediates negative inotropic effect in

cardiac myocytes by decreasing myofilament response to calcium, independently of

alterations in calcium and pH homeostasis or troponin I phosphorylation (Liao et al.,

Chapter IV

180

2002; Chen et al., 2003). Interestingly, Pellieux et al. (2000) demonstrated p38 MAPK

activation in the same transgenic animals with Ang II-induced cardiac hypertrophy.

The age-dependent reduction in relaxation performance (MRL) observed in the 35-40

week old WT population was already evident in TG myocytes at 15-20 weeks (Figure

IV-4, Panel C). Thus, a significant genotype-specific alteration with ageing was

observed for MRL (Table IV-3). This suggests that TG cardiomyocytes have already

exhausted part of their relaxation performance reserve at a younger age, predisposing

them to more profound mechanical abnormalities with disease progression. These

findings indicate that chronic overproduction of Ang II in the heart causes pronounced

cardiomyocyte contractile dysfunction linked with early onset of relaxation dysfunction

- a prominent sign of heart failure. These data also complement the study of Huggins et

al. (2003) who demonstrated reduced peak developed pressure and maximum rate of

pressure development in 30-40 week old ex vivo Langendorff-perfused hearts from the

same TG mouse line. Furthermore, in a related collaborative study, myocardial

remodeling in 50-60 week TG mice was shown to be associated with decreased in vivo

contractility and relaxation, and preceded by more subtle sign of relaxation delay

observed in hearts of 15-20 week TG (Domenighetti et al., 2005). Taken together, these

data would suggest that the hypodynamic performance of TG hearts in vivo reflects

fundamental abnormalities in cardiomyocyte E-C coupling.

Additionally, these data emphasize the important role played by ageing in the

development of inotropic dysfunction. Indeed, the age is the only determinant of the

reduced maximal peak of shortening (%S) and rate of shortening (MRS) observed in

35-40 week old cardiomyocytes (Table IV-3). Thus, it could be suggested that age-

dependent factors are associated with deterioration of cardiomyocyte kinetics consistent

with the development of systolic dysfunction. With the data presented in Chapter III,

showing increased Agt mRNA expression levels in 35-40 week old WT, the results

presented here suggest that age-dependent increased production of Ang II in the heart

could amplify late maturity abnormalities in cardiomyocyte E-C coupling resulting in

hypodynamic performance of myocardial tissue in vivo and heart failure. These age-

dependent mechanisms are accelerated and exacerbated in the TG mice due to the early

activation of the intra-cardiac RAS.

Chapter IV

181

In support of this interpretation, Wei et al. (1984) documented that ageing hearts can

exhibit prolonged LV relaxation periods and decreased contractility. This phenomenon

was also reported in studies using isolated adult mouse cardiomyocytes and in whole

heart preparations (Lim et al., 1999; Lim et al., 2000), where a decreased rate of

cytosolic calcium removal was suggested to be a causal mechanism. In addition,

gradual increases in cardiac Ang II levels were reported in experimental models and

clinically during the development of chronic heart failure where systolic dysfunction is

evident (Wollert and Drexler, 1999; Serneri et al., 2001).

This study is in partial agreement with the work in vitro done by Sakurai et al. (2002)

who demonstrated a decreased inotropy in isolated adult mouse cardiomyocytes in the

presence of various concentrations of Ang II. These authors observed a decrease in the

maximal cell shortening (%S) in cardiomyocytes in the presence of Ang II, a result

which is not confirmed by the present study. However it is important to note that, due

to methodological limitations, Sakurai et al. (2002) stimulated their cardiomyocytes at

one single and very low frequency rate (0.25 Hz), which could explain some of the

discrepancies observed between that study and the present investigation. As stated in

the Introduction of this Chapter, other differences could be related to the nature of the

cardiomyocyte and stimulus utilized, that is normal adult cardiomyocytes from

C57BL6 mice stimulated with Ang II (acute stimulus) versus Agt over-producing

cardiomyocytes (chronic stimulus). Therefore it seems likely that in the TG

overexpression model, compensatory mechanisms associated with Ang II-dependent

cardiac and cardiomyocyte remodeling are recruited to sustain maximum shortening, in

the context of slowed contraction kinetics shortening and lengthening rates.

Finally, the results reported in this study accord with findings in the Gαq-

overexpressing transgenic mouse, where cardiomyocyte dysfunction (decreased rates of

shortening and lengthening) is associated with downregulation of SERCA2 and

prolonged duration of calcium transients (Yatani et al., 1999). Given that signaling

pathways mediated by AT1 receptors in myocytes are known to be linked to the Gαq

class of G proteins (Wettschureck et al., 2001), it is tempting to speculate that chronic

cardiomyocyte dysfunction in the TG mice is caused by specific Ang II-dependent

chronic activation of cardiomyocyte AT1 Gαq-coupled receptors. This hypothesis,

Chapter IV

182

where fibroblast participation in the hypertrophic response is not obligatory, requires

further investigation.

4.2 Increased twitch duration in TG cardiomyocytes

This study also demonstrates that chronic over-production of cardiac Ang II is

associated with an increase in cardiomyocyte cycle duration at both ages (Figure IV-7).

This extended twitch duration in TG myocytes is due to delayed E-C coupling latency

(To) and prolonged duration of the shortening and lengthening periods (Tm-To; Tf-Tm)

summing to extend the whole contraction cycle period (Tf-To).

Prolongation of the lengthening period (Tf-Tm) in TG myocytes (Figure IV-7) may

reflect a decrease in the rate of calcium removal from the cytoplasm. This could be

caused by a slower calcium re-uptake into the SR by the SERCA2, which normally

accounts for the removal of ~90% of the intracellular calcium concentrations in the

rodent (Bassani et al., 1994). Indeed, a characteristic of the failing and hypertrophic

heart is a reduction in myocardial SERCA2 protein expression (De La Bastie et al.,

1990; Schotten et al., 1999). This could explain reduced calcium re-uptake in

pathological conditions. Such reduction in SERCA2 could be partially or totally

compensated by an upregulation of the NCX, enhancing efflux of calcium ions to the

extracellular medium (Terracciano et al., 2001). These hypotheses are partially

supported by the results of the present study where a decrease in SERCA2 is observed

already at the age of 15-20 weeks in TG hearts (Figure IV-8). Downregulation of

SERCA2 would suppress the re-uptake of calcium into the SR and delay myocyte

relaxation. The concomitant upregulation of the NCX1.1 (Figure IV-8) in younger

cardiomyocytes could suggest an early compensatory mechanism that would ultimately

shift the calcium cycling from an intracellular pathway toward an extracellular one,

thus reducing the levels of releasable SR calcium and chronically eroding systolic

functional reserve. This hypothesis is strengthened by a related collaborative study,

where myocardial dysfunction in TG mice was shown to be associated with decreased

systolic calcium levels and increased calcium transient duration (Domenighetti et al.,

2005). In addition, NCX overexpression in adult rabbit ventricular myocytes was

Chapter IV

183

shown to cause depressed contractility related to decreased SR calcium stores and low

diastolic calcium levels (Ranu et al., 2002).

These data emphasize the important role played by ageing in the differential regulation

of twitch duration in both WT and TG myocytes. Cardiomyocytes from older WT and

TG hearts exhibit an abbreviated whole contraction period (Tf-To) and abbreviated

duration in lengthening period (Tf-Tm). The presence of a shorter contractile cycle

period, associated with decreased maximum shortening in older myocytes would

suggest a blunted calcium transient, possibly associated with a shorter action potential

duration. A shorter contractile cycle period would also mean longer beat-to-beat

‘diastolic’ interval (for a given frequency) which could enhance recovery of voltage

operated L-type calcium channels and therefore facilitate calcium-induced calcium

release from the sarcoplasmic reticulum. This could be an important antiarrhythmic

compensatory mechanism which would counteract age-dependent decrease in rates of

shortening and lengthening. The intrinsic decrease in the cycle duration in older

myocytes may also reflect a compensatory response to offset the effects of age related

Ang II-dependent and -independent desensitization to β-adrenergic stimulation (Ai et

al., 1998; Meissner et al., 1998; Lim et al., 1999; Barki-Harrington et al., 2003), or a

change in the preload in the situation in vivo.

Prolonged contractile cycle period (Tf -To) in younger TG cardiomyocytes is a prelude

to the later emergence of increased latency (To), a phenomenon which is not observed

in WT myocytes. This is a significant genotype-specific alteration associated with

ageing (Table IV-3). The explanation for a delayed onset of shortening (To) in the older

TG cardiomyocytes is rather speculative, but may be partly due to an increase in

cellular stiffness, decreased SR calcium release (Marks, 2001) or altered action

potential threshold conditions due to increased pH-dependent cardiac background

potassium currents (Backx and Marban, 1993; Lopes et al., 2000). Increased

refractoriness of voltage operated L-type calcium channels could also lead to delayed

E-C coupling latency (Antoons et al., 2002). Additionally, a delayed onset of

shortening (To) in the older TG cardiomyocytes could be explained by a decreased

activity of the NCX working in the reverse mode. Indeed, it could be further speculated

that an upregulation of the NHE-1 and NCX exchangers, as observed at the mRNA

Chapter IV

184

level in younger TG cardiomyocytes, could lead to cardiomyocyte sodium loading and

consequently enhanced inward calcium current driven by the NCX working in the

reverse mode. In a chronic state such as that which occurs with elevated Ang II in the

TG myocardium, these same mechanisms could work to maintain an appropriate E-C

coupling and action potential duration in an environment where cardiomyocyte

molecular remodeling is associated with sodium load, alkalinization and a shift of the

calcium cycling from an intracellular pathway toward an extracellular one (decrease in

SERCA2 protein expression). However, the age-dependent decrease of the NCX1.1 and

NHE-1 would cause a substantial decrease in NCX-dependent inward calcium current,

delaying the triggering of calcium release from the SR and delaying contraction.

Finally, cardiomyocytes from older WT and TG exhibit an abbreviated shortening

period (Tm-To). There is a significant genotype-specific alteration associated with

ageing for this parameter, with an approximate 25% age-dependent reduction in WT

myocytes but only ~10% reduction in TG myocytes (Figure IV-7, Table IV-3). Age-

dependent reduction in shortening duration is consistent with downregulation of the

NHE-1 in both WT and TG myocytes (Figure IV-8). Decreasing in proton export

would reduce myofilament calcium sensitivity and shortening duration. As NHE-1 is

upregulated in older TG myocytes relative to WT, the age-dependent effect on pH

would be more limited in TG myocytes, diminishing the magnitude of reduction for

this time parameter in older TG cardiomyocytes.

4.3. Decreased inotropy and lusitropy in LLC cardiomyocytes

The present study also demonstrates that the insulin resistant cardiomyopathy driven by

the deletion of the insulin-dependent GLUT4 transporter in LLC mice is associated

with impaired cardiomyocyte inotropy (MRS and %S) and lusitropy (MRL). These

findings indicate that chronic suppression of GLUT4-mediated glucose transport in

cardiomyocytes causes pronounced contractile dysfunction. These data also

complement the study of Huggins et al. (2004) who demonstrated reduction in peak

developed pressure and maximum rate of pressure development in ex vivo Langendorff-

perfused hearts from the same knock-out mouse line. As already suggested for the TG

Chapter IV

185

myocytes, the slower MRS and MRL observed in LLC cardiomyocytes may reflect a

shift in myosin and actin isoenzymes, probably associated with decreased actomyosin

ATPase activity and depletion of high energy phosphates. A shift in actomyosin

isoforms was previously shown to occur in experimental diabetic cardiomyopathy

(Garber and Neely, 1983; Pierce et al., 1989). Slower contraction rates observed in

LLC cardiomyocytes, where intracardiac glucose transport is impaired, could reflect

differential gene expression associated with ischemic conditions, and depressed

mitochondrial activity.

This study supports the results obtained by Dutta et al. (2001), who reported reduced

rates of shortening and lengthening and depressed peak maximal shortening (albeit

inconsistent) in cardiomyocytes isolated from sucrose-fed rats. No direct comparison of

the present results can be made with the contractile function measured by Abel et al.

(1999) in cardiac-specific GLUT4 knockout mouse (G4H-/-) hearts, as the phenotype

and cardiac function observed in that model (G4H-/-) are rather different to what is

described here for the LLC. Abel et al. (1999) observed no changes in basal cardiac

function in isolated hearts from G4H-/- mice. Interestingly however, ischemic

conditions and catecholamine stimulation revealed diastolic dysfunction in these mice,

suggesting that additive factors such as systemic changes in neuro-hormonal activation,

and possibly hypoxia are all necessary to develop overt mechanical dysfunction in

GLUT4-deficient hearts. Finally, the alterations in LLC cardiomyocyte contractility

reported here are somewhat similar to those observed in type 1 diabetic experimental

models, where negative inotropy and prolonged time to peak shortening are generally

noted (Noda et al., 1993; Jourdon and Feuvray, 1993; Yu et al., 1994; Lagadic-

Gossmann et al., 1996).

Surprisingly, ageing has no effect on cardiomyocyte inotropy (%S and MRS) in LL

and LLC myocytes. However ageing was associated with an increase in the rate of

lengthening (MRL) in the same myocytes (Table IV-4; Figure IV-9). This would

suggest that in this mouse strain ageing does not impair cardiomyocyte inotropy and

surprisingly improves relaxation dynamics. This age-dependent increase in lusitropy is

difficult to explain. An increase in rate of lengthening could represent an important

compensatory mechanism in this mouse model to shorten the contractile cycle and

preserve peak shortening (%S). It is also important to remember that LL mice are not

Chapter IV

186

exempt from systemic metabolic disturbances (insulin-signaling desensitization,

impaired glucose disposal) and cardiac remodeling (+15% in CWI) when compared to

C57BL6 controls (i.e. WT) (see Chapter III, Section 1.3.2), which could partly explain

why age-dependent changes in contractile parameters between LL and LLC myocytes

present a similar trend.

4.4. Increased twitch duration in LLC cardiomyocytes

This study demonstrates that there is an increased contractile cycle duration in LLC

cardiomyocytes when compared with LL. The increase in time parameters in LLC

myocytes is due to prolonged duration of E-C coupling latency (To) and increased

duration of shortening and lengthening periods (Tm-To; Tf-Tm) summating in an

extended contraction period (Tf-To). A delayed To in type 1 diabetes was previously

reported by Kotsanas et al. (2000) in cardiomyocytes isolated from STZ-treated

diabetic rats. Decreased SR calcium release due to RyR downregulation or increased

refractoriness of voltage operated calcium channels could explain the delayed onset of

contraction in LLC myocytes and the prolonged time of contraction. This interpretation

is consistent with the downregulation of the RyR2 receptors in LLC myocardium

(Figure IV-11, Panel B.) and also with decreased voltage-operated L-type calcium

current density observed in 15 to 35 week old LLC myocytes (Danes, 2004). A reduced

cytosolic calcium influx could account for the delay in E-C coupling and myocyte

shortening in LLC myocytes.

As discussed in relation to the TG myocyte data, a prolongation of the lengthening

period (Tf-Tm) in LLC may reflect a decrease in the rate of calcium removal from the

cytoplasm. This could be caused by slower calcium re-uptake into the SR by SERCA2,

combined with reduced calcium exit by the NCX. Both NCX and SERCA2 are

downregulated in LLC hearts (Figure IV-11, Panel A & C). Interestingly, embryos

from NCX knock-out heart tubes (at day 9.5) show E-C impairment associated with

increased diastolic calcium levels, decreased systolic calcium but unchanged SERCA2

expression levels (Reuter et al., 2003), suggesting a major developmental role played

Chapter IV

187

by the NCX in the role and longitudinal regulation of contractile cycle duration and E-

C coupling in heart myocytes.

The age-dependent decrease in the contraction period (Tf-To) observed in both LL and

LLC myocytes is associated with different trends in the shortening and lengthening

periods in each myocyte type. LLC cardiomyocytes from older animals show a more

pronounced age-dependent decrease of the lengthening period (Tf-Tm), while the LL

myocytes exhibit a more pronounced decrease of the shortening period (Tm-To) with

ageing (Figure IV-10). The significant genotype-specific alteration associated with

ageing for the shortening period (Tm-To), evidenced by an approximate 25% age-

dependent reduction in LL and only about 15% in LLC myocytes, is consistent with

previous observations made for the WT and TG myocytes (Figure IV-7). The

phenomenon was attributed to an age-dependent decrease in pH due to NHE-1

downregulation which is postulated to decrease myofilament calcium sensitivity and

reduce shortening duration. Since NHE-1 is still upregulated in older LLC myocytes

relative to LL, the age-dependent reduction in pH would be limited in LLC, thus

diminishing the magnitude of reduction for this time parameter in old LLC

cardiomyocytes. An explanation for the genotype-specific alteration associated with

ageing for the lengthening period is not immediately apparent and requires further

investigation.

4.5. Negative contraction-frequency relationship in mouse

cardiomyocytes

Rodent species such as the mouse and the rat have a short action potential, high

reliance on SR calcium cycling and high intracellular sodium concentrations. These

models generally exhibit a negative contraction-frequency relationship (Bers, 2001).

This negative relationship, widely documented in the rat (Orchard and Lakatta, 1986;

Capogrossi et al., 1986), may be related to exhaustion of the contractile and SR calcium

reserves, shortened SR filling times and/or oxygenation limitations at higher pacing

rates. In the perfused mouse heart and in isolated mouse myocardial preparations

Chapter IV

188

(trabeculae or intact mouse ventricles), biphasic frequency responses have been

reported with a modest positive contraction-frequency relationship at lower frequencies

(0.2-2Hz) which converts to a modest negative slope at higher frequencies (Gao et al.,

1998; Lim et al., 1999; Stull et al., 2002). Taking into account various aspects of E-C

coupling in the mouse heart, Georgakopoulos and Kass (2001) proposed that while not

negative, the frequency relationship should be quite modest over physiological heart

rates and stimulation frequencies.

This study demonstrates a negative contraction-frequency relationship in

cardiomyocytes from all experimental groups. The present work aimed to study the

response of isotonically contracting cardiomyocytes to frequency changes spanning

from 1.5 Hz to 5Hz, the latter being the closest pacing value to physiological range

tested here. It is interesting to note that at higher frequencies (i.e. 4 Hz and 5 Hz), no

statistical differences were observed for the inotropic and lusitropic values (%S, MRS

and MRL) in the 8 experimental groups. This would suggest that the negative

frequency relationship observed at lower frequency values could reach a plateau at

5 Hz (or higher) and possibly convert into a positive staircase at even higher pacing

values. Recent work with P12 cardiomyocytes and cardiomyocytes isolated from

C57BL6 mice support this idea, since a shift between negative and positive staircases

were observed around a specific stimulatory frequency, which corresponded to 6 Hz for

P12 cardiomyocytes and 2 Hz for C57BL6 myocytes (Tiemann et al., 2003).

The decrease in inotropic and lusitropic values observed with increased pacing

frequency in cardiomyocytes from TG1306/1R and GLUT4-KO mice could be related

to a reduction of inward calcium flux (mainly driven by the L-type calcium channels

and the NCX in the reverse mode) or a reduced fractional release of calcium from the

SR. Competition between the SERCA2 and NCX for calcium extrusion from the

cytosol may also change as a function of frequency. As reviewed by Bers (2001) the

SERCA2 becomes increasingly dominant over the NCX exchanger in transporting the

calcium from the cytosol at higher frequencies. This phenomenon is probably less

important in the mouse and the rat, where calcium removal from the cytosol is ~90%

driven by the SERCA2. For all the experimental groups reported in the present study,

differences in cardiomyocyte contractile parameters were proportionally consistent over

the range of pacing frequencies evaluated, indicating that decreases in SERCA2 protein

Chapter IV

189

levels and/or shifts in NCX gene expression do not directly affect the SR calcium

loading capability of myocytes under different genotypic backgrounds or at different

ages during a frequency staircase.

In agreement with previous observations (Lim et al., 2000; Antoons et al., 2002), the

present study reports a negative frequency relationship also for the twitch duration, that

is, shorter shortening and lengthening cycle periods at higher frequencies, in

combination with an increase in time of latency (To). Interestingly, Antoons et al.

(2002) reported a loss of trigger for calcium release from the SR at higher frequencies

in the mouse cardiomyocyte, which could explain in part the increase in To observed

here. In addition, a slowed recovery from inactivation of the voltage-operated channels

at higher frequencies could lead to a diminished inward calcium current via the voltage-

operated L-type calcium channels.

4.6. After-contraction, matters of calcium overload?

The NCX can have an adverse effect on contractility and arrhythmogenesis under

pathophysiological conditions. For instance, SR calcium overload is characterized by

waves of spontaneous SR calcium release that propagate through the cytoplasm

generating inward NCX currents (calcium out, sodium in) and delayed after-

depolarizations (Schlotthauer and Bers, 2000; Mackenzie et al., 2002; Pogwizd and

Bers, 2004). This mechanism is thought to underlie arrhythmias and after-contractions

associated with myocardial ischemia-reperfusion (Ch’en et al., 1998). Delayed after-

contractions observed in LLC myocytes at 1.5Hz could be generated by spontaneous

SR calcium release caused by calcium overloading conditions. At a low frequency rate

(1.5Hz), increased and prolonged systolic calcium transient, coupled with SERCA2 and

NCX1 downregulation, could create a state of calcium overload under low SR calcium

reserves in LLC myocytes. This would elicit spontaneous SR calcium release and

resting calcium levels to rise significantly. This state of calcium overload could trigger

ectopic beats and cause arrhythmias as observed in LLC cardiomyocytes at 1.5Hz.

Chapter IV

190

5. IN SUMMARY

The present Chapter reports impaired contractile function of isolated cardiomyocytes

and differential expression of key transporters involved in the calcium and pH

homeostasis in TG1306/1R and GLUT4-KO mice. A summary of the age-dependent

and genotype-dependent trends and variations of the contractile parameters for the 2

mouse models are reported in Tables IV-3 and IV-4. These data are further considered,

in the broader comparative context in conjunction with other morphological and

molecular data, in Chapter VI.

Chapter IV

191

98.7±2.7*79.3±4.2(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 5.7±0.35.9±0.5%S

85.1±2.7*67.9±3.8Tf - To

53.8±1.9*41.9±2.5Tf - Tm

31.3±1.6*26.0±1.5Tm - To

13.7±1.1*11.5±0.8To

2.9±0.1*4.2±0.3MRL

3.6±0.2*4.1±0.2MRS

1517n (myocytes) =

TGWT

15-20 weekParameter

98.7±2.7*79.3±4.2(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 5.7±0.35.9±0.5%S

85.1±2.7*67.9±3.8Tf - To

53.8±1.9*41.9±2.5Tf - Tm

31.3±1.6*26.0±1.5Tm - To

13.7±1.1*11.5±0.8To

2.9±0.1*4.2±0.3MRL

3.6±0.2*4.1±0.2MRS

1517n (myocytes) =

TGWT

15-20 weekParameterA.

99.6±4.4*67.9±3.3(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 4.2±0.55.0±0.4%S

80±3.5*56.1±2.5Tf - To

52.9±3.0*36.2±2.2Tf - Tm

27.1±1.3*19.9±0.7Tm - To

19.6±2.0*11.8±1.0To

2.6±0.3*3.5±0.3MRL

2.6±0.2*3.8±0.3MRS

1516n (myocytes) =

TGWT

35-40 weekParameter

99.6±4.4*67.9±3.3(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 4.2±0.55.0±0.4%S

80±3.5*56.1±2.5Tf - To

52.9±3.0*36.2±2.2Tf - Tm

27.1±1.3*19.9±0.7Tm - To

19.6±2.0*11.8±1.0To

2.6±0.3*3.5±0.3MRL

2.6±0.2*3.8±0.3MRS

1516n (myocytes) =

TGWT

35-40 weekParameterB.

Table IV-1: Isotonic shortening of adult cardiomyocytes from WT and TG mice (5 Hz)

A. 15-20 week old WT and TG; B. 35-40 week old WT and TG. Parameters calculated included the maximum cell shortening (%S), expressed as a percentage of initial resting cell length (Lo); the time at which the cell commenced shortening after stimulus application, i.e. the excitation-contraction coupling latency (To); the time at %S (Tm); the time at which the cell length returned to Lo (Tf); the maximal rate of cell shortening (MRS) and lengthening (MRL). Tf -To indicates duration of the whole contractile cycle; Tm-To, duration of shortening; Tf -Tm, duration of lengthening. * p<0.05, TG vs. age-matched WT (one-way ANOVA)

Chapter IV

192

98.7±2.7*79.3±4.2(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 5.7±0.35.9±0.5%S

85.1±2.7*67.9±3.8Tf - To

53.8±1.9*41.9±2.5Tf - Tm

31.3±1.6*26.0±1.5Tm - To

13.7±1.1*11.5±0.8To

2.9±0.1*4.2±0.3MRL

3.6±0.2*4.1±0.2MRS

1517n (myocytes) =

LLCLL

15-20 weekParameter

98.7±2.7*79.3±4.2(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 5.7±0.35.9±0.5%S

85.1±2.7*67.9±3.8Tf - To

53.8±1.9*41.9±2.5Tf - Tm

31.3±1.6*26.0±1.5Tm - To

13.7±1.1*11.5±0.8To

2.9±0.1*4.2±0.3MRL

3.6±0.2*4.1±0.2MRS

1517n (myocytes) =

LLCLL

15-20 weekParameterA.

4.8±0.7*6.6±0.3

75.0±7.870.9±2.0

51.2±5.245.0±1.5

23.7±2.725.9±1.4

24.6±2.3*12.1±0.6

2.6±0.3*3.9±0.1

3.2±0.4*4.4±0.2

1615

99.6±6.4*83.0±2.4

99.6±4.4*67.9±3.3(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 4.2±0.55.0±0.4%S

80±3.5*56.1±2.5Tf - To

52.9±3.0*36.2±2.2Tf - Tm

27.1±1.3*19.9±0.7Tm - To

19.6±2.0*11.8±1.0To

2.6±0.3*3.5±0.3MRL

2.6±0.2*3.8±0.3MRS

1516n (myocytes) =

LLCLL

35-40 weekParameter

99.6±4.4*67.9±3.3(ms)Tf

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) 4.2±0.55.0±0.4%S

80±3.5*56.1±2.5Tf - To

52.9±3.0*36.2±2.2Tf - Tm

27.1±1.3*19.9±0.7Tm - To

19.6±2.0*11.8±1.0To

2.6±0.3*3.5±0.3MRL

2.6±0.2*3.8±0.3MRS

1516n (myocytes) =

LLCLL

35-40 weekParameterB.

5.0±0.4*6.2±0.6

63.3±3.259.7±3.3

38.8±2.139.5±2.7

24.5±1.2*20.2±1.0

15.4±1.3*11.3±1.1

3.4±0.4*4.5±0.5

3.8±0.5*4.8±0.5

1615

78.7±4.5*71.0±3.0

Table IV-2: Isotonic shortening of adult cardiomyocytes from LL and LLC mice (5 Hz)

A. 15-20 week old LL and LLC; B. 35-40 week old LL and LLC. Parameters calculated included the maximum cell shortening (%S), expressed as a percentage of initial resting cell length (Lo); the time at which the cell commenced shortening after stimulus application, i.e. the excitation-contraction coupling latency (To); the time at %S (Tm); the time at which the cell length returned to Lo (Tf); the maximal rate of cell shortening (MRS) and lengthening (MRL). Tf -To indicates duration of the whole contractile cycle; Tm-To, duration of shortening; Tf -Tm, duration of lengthening. * p<0.05, LLC vs. age-matched LL (one-way ANOVA)

Chapter IV

193

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) ‡†%S

‡†*Tf - To

‡†*Tf - Tm

¶‡†*Tm - To

¶‡†*To

¶‡†*MRL

‡†*MRS

Age-Genot.FrequencyAgeGenotypeParameter

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) ‡†%S

‡†*Tf - To

‡†*Tf - Tm

¶‡†*Tm - To

¶‡†*To

¶‡†*MRL

‡†*MRS

Age-Genot.FrequencyAgeGenotypeParameter

Table IV-3: Summary of the multiple-way ANOVA analysis for the TG and WT myocytes

Significant genotype-, age-, frequency-dependent and interaction (age-genotype) findings for the contractile parameters of WT and TG cardiomyocytes at 15-20 and 35-40 weeks.

Chapter IV

194

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) ‡%S

‡†*Tf - To

‡†*Tf - Tm

¶‡†*Tm - To

¶‡†*To

‡†*MRL

‡*MRS

Age-Genot.FrequencyAgeGenotypeParameter

*

(ms)

(ms)

(ms)

(ms)

(Lo/s)

(Lo/s)

(L/Lo) ‡%S

‡†*Tf - To

‡†*Tf - Tm

¶‡†*Tm - To

¶‡†*To

‡†*MRL

‡*MRS

Age-Genot.FrequencyAgeGenotypeParameter

*

Table IV-4: Summary of the multiple-way ANOVA analysis for the LLC and LL myocytes

Significant genotype-, age-, frequency-dependent and interaction (age-genotype) findings for the contractile parameters of LL and LLC cardiomyocytes at 15-20 and 35-40 weeks.

Chapter IV

195

Figure IV-1: Effects of acute Ang II modulation of cardiomyocyte contractility

Refer to text in Section 1.2.8 of this Chapter for a detailed explanation of the different pathways. A. Ang II can activate the sodium-calcium exchanger (NCX) working in reverse mode (calcium ‘in’) via a PKC-dependent signal; B. Ang II can activate the sodium-hydrogen exchanger (NHE) via a PKC-dependent signaling pathway; C. Acutely, Ang II can promote calcium efflux via the NCX; D. Activation of the NHE result in cytosolic alkalinization and sodium load, influencing the myofilament calcium sensitivity and the activity of the NCX working in the reverse mode E. Ang II can modulate the positive inotropic effect through activation of the phospholipase D (PLD) and the NADPH oxidase, leading to production of reactive oxygen species (ROS). These superoxide anions induce the activation of the NHE via an ERK/MAPK-dependent pathway.

VOC

[Na+]i

[Ca2+]e

NCX

NCX

[H+]i

NHE-1

‘alkalosis’

[Ca2+]i

A.A.

[Na+]e

B.B.

‘inwardcalcium current’

‘positive inotropiceffect’

SR

[Na+]e

[Ca2+]i

C.C.

[Ca2+]i

[Na+]i

Ang II

AT1

myofilamentsensitivity

‘calciumoverload’

SERCA2

D.D.

PKC

NADPH oxidase

‘ROS’

‘Erk MAPK’ E.E.

‘calcium in’

‘calcium out’

oo

oo

oo

Chapter IV

196

Figure IV-2: Effects of chronic Ang II modulation of cardiomyocyte contractility

Refer to text in Section 1.2.8 of this Chapter for a detailed explanation of the different pathways. A. Alkalosis promotes growth and proliferation; B. Alkalosis also reduces ATP production and C. can promote apoptosis; D. Ang II-induced expression of endothelin-1 can (E.) enhance intracellular signaling and sustain differential gene expression and remodeling; F. re-expression of fetal isoforms of contractile proteins can result in slower contraction kinetics; G. NADPH oxidase-induced production of reactive oxygen species (ROS) can activate the NHE via an ERK MAPK pathway; H. mitochondria-induced production of ROS, can activate the p38 MAPK and JNK leading to differential gene expression.

‘growthproliferation’

(?) VOC

[Na+]i

[Ca2+]e

NCX (?)

NCX (?)

[H+]i

(?) NHE-1

‘alkalosis’

[Ca2+]i

A.A.

[Na+]e

B.B.

‘inwardcalcium current’

‘negative inotropiceffect’

SR

[Na+]e

[Ca2+]i

C.C.

[Ca2+]i

Ang II

AT1

SERCA2

D.D.

PKC

NADPH oxidase

‘ROS’

Endothelin-1ETA

α-SkAβ-MHC

ATP Mitoch.

‘gene expression’

‘ROS’

‘apoptosis’

NucleusE.E.

G.G.

F.F.

H.H.

p38 and JNK

o

o

o

‘growthproliferation’

(?) VOC

[Na+]i

[Ca2+]e

NCX (?)

NCX (?)

[H+]i

(?) NHE-1

‘alkalosis’

[Ca2+]i

A.A.

[Na+]e

B.B.

‘inwardcalcium current’

‘negative inotropiceffect’

SR

[Na+]e

[Ca2+]i

C.C.

[Ca2+]i

Ang II

AT1

SERCA2

D.D.

PKC

NADPH oxidase

‘ROS’

Endothelin-1ETA

α-SkAβ-MHC

ATP Mitoch.

‘gene expression’

‘ROS’

‘apoptosis’

NucleusE.E.

G.G.

F.F.

H.H.

p38 and JNK

oo

oo

oo

Chapter IV

197

0

2

4

6

8

1 0

1 2

1 4

1 6

1 8

%S

3Hz (basal) 1.5Hz 3Hz 4Hz 5HzFrequency:

100 30 30 30 30 100No cycles averaged:

Time (min): 2:00 4:25 7:00

A.

B.

3Hz (basal)

0

2

4

6

8

1 0

1 2

1 4

1 6

1 8

%S

3Hz (basal) 1.5Hz 3Hz 4Hz 5HzFrequency:

100 30 30 30 30 100No cycles averaged:

Time (min): 2:00 4:25 7:00

A.

B.

3Hz (basal)

Figure IV-3: Protocol summary for the contractility experiments

A. Protocol summary for contractility experiments with frequency variation and description of variable times and frequencies of parameter measurement. B. Typical shortening profile (%S) for a cell contracting according to the protocol illustrated above and number of contractile cycles averaged (with their location) at each frequency for analysis.

Chapter IV

198

Figure IV-4: Evaluation of %S and Tf prior to and after the frequency ramp

Evaluation of %S (A, B, C, D) and Tf (E, F, G, H) prior to the frequency ramp (‘pre’) and at the end of the experimental protocol (‘post’) in 15-20 and 35-40 week WT and TG myocytes. No statistical differences were observed.

%S TfA

B

C

D

E

F

G

H

15-2

0 w

. WT

15-2

0 w

. TG

35-4

0 w

. WT

35-4

0 w

. TG

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

%S TfA

B

C

D

E

F

G

H

15-2

0 w

. WT

15-2

0 w

. TG

35-4

0 w

. WT

35-4

0 w

. TG

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

f%S TfA

B

C

D

E

F

G

H

15-2

0 w

. WT

15-2

0 w

. TG

35-4

0 w

. WT

35-4

0 w

. TG

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

%S TfA

B

C

D

E

F

G

H

15-2

0 w

. WT

15-2

0 w

. TG

35-4

0 w

. WT

35-4

0 w

. TG

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

2

4

6

8

WT pre WT post

0

2

4

6

8

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

0

20

40

60

80

100

120

140

WT pre WT post

0

20

40

60

80

100

120

140

TG pre TG post

ff

Chapter IV

199

Figure IV-5: Evaluation of %S and Tf prior to and after the frequency ramp

Evaluation of %S (A, B, C, D) and Tf (E, F, G, H) prior to the frequency ramp (‘pre’) and at the end of the experimental protocol (‘post’) in 15-20 and 35-40 week LL and LLC myocytes. No statistical differences were observed.

%S Tf

A

B

C

D

E

F

G

H

15-2

0 w

. LL

15-2

0 w

. LLC

35-4

0 w

. LL

35-4

0 w

. LLC

0

2

4

6

8

LL Pre LL Post

0

2

4

6

8

LLC Pre LLC Post

0

2

4

6

8

LL Pre LL Post

0

2

4

6

8

LLC Pre LLC Post

0

40

80

120

LL Pre LL Post

0

40

80

120

LLC Pre LLC Post

0

40

80

120

LL Pre LL Post

0

40

80

120

LLC Pre LLC Post

f%S Tf

A

B

C

D

E

F

G

H

15-2

0 w

. LL

15-2

0 w

. LLC

35-4

0 w

. LL

35-4

0 w

. LLC

0

2

4

6

8

LL Pre LL Post

0

2

4

6

8

LLC Pre LLC Post

0

2

4

6

8

LL Pre LL Post

0

2

4

6

8

LLC Pre LLC Post

0

40

80

120

LL Pre LL Post

0

40

80

120

LLC Pre LLC Post

0

40

80

120

LL Pre LL Post

0

40

80

120

LLC Pre LLC Post

ff

Chapter IV

200

3

6

9

12

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3.5

5

6.5

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

Hz 1.5 Hz 3 Hz 4 Hz 5

A.

B.

C.

Normalized %S

Normalized MRS

Normalized MRL

1.5 3 4 5

Lo/s

Lo/

s%

Lo

(Hz)

(Hz)

(Hz)

1.5 3 4 5

1.5 3 4 5 3

6

9

12

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3.5

5

6.5

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

Hz 1.5 Hz 3 Hz 4 Hz 5

A.

B.

C.

Normalized %S

Normalized MRS

Normalized MRL

1.5 3 4 5

Lo/s

Lo/

s%

Lo

(Hz)

(Hz)

(Hz)

1.5 3 4 5

1.5 3 4 5

*

*†

A. Normalized %S

B. Normalized MRS

C. Normalized MRL

35-40 week, WT

15-20 week, WT

35-40 week, TG

15-20 week, TG

Figure IV-6: Pacing responses of isotonically shortening adult myocytes from WT and TG

Inotropic and lusitropic parameters. See legend of Table IV-1 for description of abbreviations. * p<0.05, TG vs. WT cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

Chapter IV

201

35-40 week, WT

15-20 week, WT

35-40 week, TG

15-20 week, TG

Figure IV-7: Pacing responses of isotonically shortening adult myocytes from WT and TG

Cycle time parameters. See legend of Table IV-1 for description of abbreviations. * p<0.05, TG vs. WT cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

8

12

16

20

24

Hz 1.5 Hz 3 Hz 4 Hz 540

80

120

160

Hz 1.5 Hz 3 Hz 4 Hz 5

10

20

30

40

50

Hz 1.5 Hz 3 Hz 4 Hz 530

50

70

90

110

Hz 1.5 Hz 3 Hz 4 Hz 5

A.

C.

B.

D.

Tf-To

Tf - Tm

To

Tm-To1.5 3 4 5

ms

ms

(Hz)(Hz)

1.5 3 4 5 (Hz)

1.5 3 4 5

(Hz)1.5 3 4 5

ms

ms

*†

‡ ¶

*

‡ ¶

*

*

A. To B. Tf-To

C. Tm-To D. Tf-Tm

To Tf -To

Tf -TmTm-To

Chapter IV

202

Figure IV-8: mRNA and protein expression profiles in WT and TG hearts

Expression profiles of sarcoplasmic reticulum calcium pump SERCA2 protein levels (A.), ryanodine receptor RyR2 mRNA (B.), sodium-calcium exchanger NCX1.1 mRNA (C.) and sodium-hydrogen exchanger NHE-1 mRNA (D.). In parenthesis are the number of animals per group. Statistical analysis: * p<0.05, TG vs. WT (one-way ANOVA, i.e. unpaired t-test); † p<0.05, age-genotype interaction effect (multiple-way ANOVA); ‡ p<0.05, TG vs. WT (multiple-way ANOVA); § p<0.05, 35-40 vs. 15-20 week old cardiomyocytes (multiple-way ANOVA).

TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

SERCA2 RyR2

NCX1.1 NHE-1

*

(n=10)(n=10) TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

SERCA2 RyR2

NCX1.1 NHE-1

TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

TG1306/1R (n=10)WT (n=10)

15-20 week 15-20 week

15-20 week15-20 week

35-40 week

35-40 week35-40 week

A. B.

C. D.

Expr

essi

on le

v el

(Arb

. Uni

ts)

*

NCX1.1

SERCA2 RyR2

NHE-1

§ ‡

0

40

80

120

160

010

2030

40

50

60

0

10

20

30

40

50

0

2

4

6

8

10

Expr

essi

on le

v el

(Arb

. Uni

ts)

Expr

essi

on le

vel

(Ar b

. Uni

ts)

Expr

essi

on le

vel

(Arb

. Uni

ts)

‡ ‡

SERCA2 RyR2

NCX1.1 NHE-1

*

(n=10)(n=10)

Chapter IV

203

3

6

9

12

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

7

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

7

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5

Lo/

sL

o/s

%L

o

(Hz)

(Hz)

(Hz)

1.5 3 4 5

1.5 3 4 5

*

*†

‡*

A. Normalized %S

B. Normalized MRS

C. Normalized MRL

35-40 week, LL

15-20 week, LL

35-40 week, LLC

15-20 week, LLC

35-40 week, LL

15-20 week, LL

35-40 week, LL

15-20 week, LL

35-40 week, LLC

15-20 week, LLC

35-40 week, LLC

15-20 week, LLC

Figure IV-9: Pacing responses of isotonically shortening adult myocytes from LL and LLC mice

Inotropic and lusitropic parameters. See legend of Table IV-2 for description of abbreviations. * p<0.05, LLC vs. LL cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

Chapter IV

204

35-40 week, LL

15-20 week, LL

35-40 week, LLC

15-20 week, LLC

35-40 week, LL

15-20 week, LL

35-40 week, LL

15-20 week, LL

35-40 week, LLC

15-20 week, LLC

35-40 week, LLC

15-20 week, LLC

Figure IV-10: Pacing responses of isotonically shortening adult myocytes from LL and LLC

Cycle time parameters. See legend of Table IV-2 for description of abbreviations. * p<0.05, LLC vs. LL cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

8

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 5

18

26

34

42

Hz 1.5 Hz 3 Hz 4 Hz 520

40

60

80

100

Hz 1.5 Hz 3 Hz 4 Hz 5

50

80

110

140

Hz 1.5 Hz 3 Hz 4 Hz 5

ms

ms

1.5 3 4 5 (Hz) (Hz)1.5 3 4 5

ms

ms

*

‡ ¶

*

‡ ¶

8

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5 (Hz)

*†

‡ ¶

50

80

110

140

Hz 1.5 Hz 3 Hz 4 Hz 5

*

(Hz)1.5 3 4 5

A. To B. Tf-To

C. Tm-To D. Tf-Tm

To Tf -To

Tf -TmTm-To

8

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 5

18

26

34

42

Hz 1.5 Hz 3 Hz 4 Hz 520

40

60

80

100

Hz 1.5 Hz 3 Hz 4 Hz 5

50

80

110

140

Hz 1.5 Hz 3 Hz 4 Hz 5

ms

ms

1.5 3 4 5 (Hz) (Hz)1.5 3 4 5

ms

ms

*

‡ ¶

*

‡ ¶

8

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5 (Hz)

*†

‡ ¶

50

80

110

140

Hz 1.5 Hz 3 Hz 4 Hz 5

*

(Hz)1.5 3 4 5

A. To B. Tf-To

C. Tm-To D. Tf-Tm

To Tf -To

Tf -TmTm-To

Chapter IV

205

Figure IV-11: mRNA and protein expression profiles in LL and LLC hearts

Expression profiles of sarcoplasmic reticulum calcium pump SERCA2 protein levels (A.), ryanodine receptor RyR2 mRNA (B.), sodium-calcium exchanger NCX1.1 mRNA (C.) and sodium-hydrogen exchanger NHE-1 mRNA (D.). In parenthesis are the number of animals per group. Statistical analysis: * p<0.05, LLC vs. LL (one-way ANOVA); † p<0.05, age-genotype interaction effect (multiple-way ANOVA); ‡ p<0.05 LLC vs. LL (multiple-way ANOVA): § p<0.05, 35-40 vs. 15-20 week myocytes (multiple-way ANOVA).

LLC (n=10)LL (n=10)

15-20 week 15-20 week 35-40 week

A. SERCA2 B. RyR2

C. NCX1.1 D. NHE-1

E xpr

essi

o n le

vel

(Arb

. Uni

ts)

*

Expr

essi

o n le

vel

(Arb

. Un i

ts)

Expr

e ssi

on le

v el

(Arb

. Uni

ts)

Expr

e ssi

on le

v el

(Arb

. Uni

ts)

20

60

100

140

0

10

20

30

40

50

0

10

20

30

40

50

15-20 week 35-40 week0

2

4

68

1012

0

2

4

68

1012

15-20 week 35-40 week

‡ ‡

0

3

69

12

15

18

0

3

69

12

15

18

‡ ‡

‡ ‡§

*

LLC (n=10)LL (n=10)

15-20 week 15-20 week 35-40 week

A. SERCA2 B. RyR2

C. NCX1.1 D. NHE-1

E xpr

essi

o n le

vel

(Arb

. Uni

ts)

*

Expr

essi

o n le

vel

(Arb

. Un i

ts)

Expr

e ssi

on le

v el

(Arb

. Uni

ts)

Expr

e ssi

on le

v el

(Arb

. Uni

ts)

20

60

100

140

0

10

20

30

40

50

0

10

20

30

40

50

15-20 week 35-40 week0

2

4

68

1012

0

2

4

68

1012

15-20 week 35-40 week

‡ ‡

0

3

69

12

15

18

0

3

69

12

15

18

‡ ‡

‡ ‡§

*

Chapter IV

206

%S

100

90

50 100 ms200150

Figure IV-12: After-contraction in LLC myocytes

Cell shortening profile and ‘delayed after contraction’ recorded from a 35-40 week old LLC cardiomyocyte at a pacing frequency of 1.5Hz.

CHAPTER V

Microarray analysis of global changes in gene

expression in Ang II-induced and insulin resistant

cardiac hypertrophy

Chapter V

208

1. INTRODUCTION

1.1. Application of microarray analysis

The microarray technique is rapidly developing as a general molecular biology

analytical technique. The technique is currently being exploited as a new tool to

perform genomic analysis and gene expression analysis.

Identification of new genes by examining nucleic acid sequences derived from open

reading frames has proved to be an efficient way of annotating the human genome and

facilitating the use of genomic information for experimental purposes (Shoemaker et

al., 2001). Furthermore, microarrays have been applied as a cost-effective technology

to screen for all possible mutations and sequence variations in genomic DNA (Hacia,

1999; Larsen et al., 2001). Samples can be sequenced using microarray hybridization

(Drobyshev et al., 1997), thus providing convenient means for identifying new genetic

variants.

Gene expression analysis examines the composition of cellular RNA populations. The

identity of transcripts that make up these populations and their expression levels are

informative of gene activity and cell state. As the precursors of translated proteins,

changes in mRNA levels are related to changes in the proteome and, ultimately, to the

physiology of the cell. A typical microarray gene expression analysis experiment

compares the relative expression levels of specific transcripts in two samples as

detailed in Chapter II (Section 8). One of these samples is a control and the other is

derived from cells or tissues whose response or status is being investigated (e.g. in

disease conditions, or after drug treatment and genetic modification) (Debouck and

Goodfellow, 1999, Duggan et al., 1999; Burgess, 2001). Each sample is labeled with

different fluorescent dyes (e.g. Cy3 and Cy5) and equal amounts of the labeled samples

are combined and hybridized with the microarray. The fluorescent signals

corresponding to the two dyes are measured independently from each spot on the

Chapter V

209

microarray (i.e. ‘probe’ nucleic acid sequences complementary to the labeled target

sample) after hybridization. After normalization, the intensity of the two hybridization

signals for each spot can be compared. Equal signal intensity from both samples

suggests equal expression in both samples (see Chapter II for further details).

1.2. cDNA versus oligonucleotides microarray assays

Before the availability of complete or near-complete eukaryotic genome sequences,

genes expressed in cells, tissues and organs were identified through sequence analysis

of cDNA banks. cDNA clones from cDNA banks of Arabidopsis and human peripheral

blood lymphocytes were used in the construction of the first cDNA expression arrays

(Schena et al., 1995). Briefly, a library of cDNA clones contained in 96- or 384-well

plates is cultured in bacteria. cDNA inserts from plasmids are amplified and products

are spotted onto activated glass slides. Two targets from mRNA are differentially

labeled, hybridized to slides and scanned. This method has been the foundation of most

gene expression research where the starting material is plasmids from cDNA banks (see

below and Duggan et al., 1999; Lou et al., 2001).

Oligonucleotide arrays were developed by Affymetrix (http://www.affymetrix.com)

using photolithographic methods similar to the production of computer chips that were

subsequently adapted to gene expression studies (Lockhart et al., 1996; Lipshutz et al.,

1999). In brief, pairs of oligonucleotides (~25 bases in length) with exact sequence (i.e.

perfect match) and one mismatch are synthesized in situ by photolithographic methods.

The mismatch oligonucleotide has one-base mismatch in the center position and it is

used as a control to detect background noise and cross-hybridization from unrelated

targets. Eleven-to-twenty different oligonucleotides are made for each gene or

transcript sequence to ‘tile’ or cover a portion of the 3’ end of the target mRNA. The

present commercial format is a 1.28 x 1.28 cm microarray ‘chip’ containing up to

500,000 different oligonucleotide sequences. One single target is hybridized to each

chip and differential analysis between targets is made by comparing different chip sets.

Chapter V

210

Technology has progressed to such an extent that publication of new genome sequences

has become commonplace. Therefore, the flexibility in designing and producing new

arrays to capitalize on this wealth of sequence data is an important issue. The

production of new arrays via the synthesis of long, 60-mer oligonucleotides by an ink-

jet printing process addresses this point (Blanchard and Hood., 1996; Hughes et al.,

2001). Algorithms are used to design long oligonucleotides of the same size, but with

little sequence homology. Protocols for printing, hybridizing and analyses are similar to

cDNA arrays. This versatile system can routinely produce arrays with 25,000 or more

elements.

1.3. Application of microarray analysis to the study of heart

development and cardiovascular diseases

The cDNA microarray technique has been used to investigate cellular and tissue

mechanisms of cardiac physiology and pathophysiology. In recent years, the number of

microarray studies reporting differentially upregulated genes in the cardiovascular

system under various pathophysiological conditions has literally exploded, producing

an exponential amount of data and information. The following Section summarizes a

selection of studies where the microarray technique was applied to pioneer the study of

gene expression in the cardiac tissue under different pathophysiological conditions.

1.3.1. Heart and cardiomyocyte development

Sehl et al. (2000) used cDNA microarray hybridization to determine differential gene

expression in embryonic and neonatal (vs. adult) rat ventricular cardiac tissue. The

greatest contrast was seen when comparing neonatal with adult myocardium and, as

expected, this reflected higher expression of signal transduction and growth regulatory

proteins in the developing heart. Among the upregulated genes in the embryonic state

(a ratio >1.5 between embryo and adult tissue) were the apoptotic Bcl-x, genes coding

Chapter V

211

for stress-related heat-shock proteins (Hsp70, Hsp90, Hsc70), transporters (β-globin,

ferritin, annexin VI), transcription factors (c-Myc, n-Myc, Ubf-1), translation factors

(Ef-1α), ribosomal units (11S, 20S, mitochondrial 12S and 16S), G proteins (Giα-1, -2,

-3 and Gs), receptors (Ppar-δ, IGF-1 receptor), structural proteins (vimentin, β-Actin,

MLC isoforms, collagen III, fibronectin, osteopontin), hormones (ANP, BNP, TNF-α)

and metabolism or energy-related proteins (GAPDH, aldehyde reductase, pyruvate

kinase). Other genes, such as SERCA2, troponin I and lipoprotein lipase were found to

be downregulated in the embryo relative to the adult (a ratio <0.6 between embryo and

adult tissue). All together, in that study 12 genes previously described as being

differentially expressed during myocardial development were identified together with

10 uncharacterized expressed sequence tags (ESTs) and 36 genes not previously

associated with cardiac development.

Peng et al. (2002) used cDNA microarray assays to study the gene expression profile of

a sub-line of P19 cells (P19CL6) differentiating into spontaneously beating

cardiomyocytes after 2 to 14 days of DMSO stimulation. Having demonstrated that

treatment with DMSO for a minimum of 4 days was required for differentiation and

initiation of contractions, the authors were particularly interested in genes for which

expression was altered during or after this 4 day time period. Thirteen genes for which

expression increased more than 2-fold before day 4 were identified (including c-Myc,

the downstream effector of Wnt signaling LEF1 and the metabolism-related genes

carbonic anhydrase 2 and phosphoenolpyruvate carboxykinase), in addition to a set of

16 unknown ESTs. Among the genes of which expression was increased more than 2-

fold after day 4, Peng et al. found 69 known genes and 105 unknown ESTs. Known

genes coded for transcription factors (including the pro-hypertrophic MEF2C), ECM

and signaling proteins (elastin, fibrillin 2, lamin and laminin isoforms, collagen V,

tenascin C, TIMP3, BMP-1), growth factors and hormones (TGF-β2, placental growth

factor, IGFBP), cytoskeleton (MLC isoforms) and intracellular signalling-related

molecules (annexin V, annexin A3 and A6, calcineurin catalytic subunit, Akt2,

MAPKK2).

Chapter V

212

1.3.2. Myocardial infarction and ischemia

Rodent models of myocardial infarction (MI) result in ischemic, inflammatory and

fibrotic repair responses within necrotic myocardium in addition to compensatory

hypertrophy of the remnant viable ventricle. Sehl et al. (2000) used cDNA microarrays

to compare gene expression profiles at different times after MI (1, 3 and 7 days).

Among the differentially expressed genes, they identified 14 genes not previously

associated with MI which exhibited significant change in their expression pattern (>1-

fold overexpression vs. sham-operated animals), including: an ATP synthase (at day 3

after MI), β-globin (at day 3 after MI), cathepsin B (from day 1 after MI), ferritin light

chain (at day 3 after MI), nonmuscular MLC (from 1 week after MI), phospholemman

(from day 3 after MI) and several ribosomal units (11S, 20S, mitochondrial 12S and

16S).

Stanton et al. (2000) examined the infarcted LV free wall and septum at 5 different

time points after MI in the rat, evaluating the expression of 7000 clones isolated from a

normalized rat left ventricular cDNA library. Over 700 genes, classified in 7 functional

clusters by the authors, were shown to have reproducible patterns of differential

expression (http://circres.ahajournals.org/cgi/content/full/86/9/939/DC2/1/). Among the

genes, Stanton et al. found several new genes not previously associated with the

process of repair and remodeling. Within the functional groups 'protein expression' and

'gene expression', most of the elevated expression was for ribosomal proteins,

elongation factors (eIF-5A, EF1α), genes encoding enzymes involved in protein

modification and degradation, transcription factors and total RNA synthesis (RNA

polymerase II, transcription factor S-II, basic transcription factors, MEF2C, GATA-

GT1); in the 'cell structure/motility' group, there was enhanced expression of genes

encoding for cytoskeletal and ECM proteins (collagens I, III, V, VI and XV, TIMP3,

gelsolin, fibrillins, fibronectins) and suppression of genes encoding for contractile

proteins. Many genes in the 'metabolism' category encoded proteins involved in energy

metabolism and, within this group, lipid metabolism genes were primarily suppressed.

In contrast, very few genes in the 'cell division' (a DNA polymerase and some histones)

Chapter V

213

or 'cell and organism defense' (hsp27 and hsp70) were found to be differentially

regulated. Among the genes whose the expression was upregulated in the 'cell

signaling/communication' category, Stanton et al. found some growth factors and

hormones (ANF, BNP, FGF12), calcyclin, some Rab and Rho molecule isoforms,

Grb2, and an L-type calcium channel isoform, while the downregulated genes included

the RyR receptor and SERCA2. Most of these genes showed different patterns of

expression between the LV and the septum, with some genes uniquely expressed in the

LV. Interestingly, Stanton et al. (2000) did not find genes with a unique and specific

expression in the septum when compared to the LV free wall.

1.3.3. Cardiac hypertrophy and heart failure

Friddle et al. (2000) applied microarray expression profiling to identify genes altered

during induction and regression phases of Ang II- and isoproterenol-induced cardiac

hypertrophy in the mouse. The set of genes exhibiting expression changes identified by

their analysis was limited to the induction (32 genes) or regression (8 genes) phase and

a group of 15 genes exhibiting biphasic expression changes. Their study identified 30

genes which were not previously associated with cardiac hypertrophy. Among the

genes differentially expressed during the induction phase of cardiac hypertrophy are: an

ATP synthase chain, the vasopressin receptor, natriuretic peptides, TGF alpha, annexin

I, PKC binding protein (beta and an alpha actin isoform) were upregulated, whereas

some ribosomal subunits (28S and mitochondrial 16S), the estrogen receptor and the

endothelin converting enzyme 2 were downregulated. The important finding of the

study was the identification of a set of genes specifically altered during the regression

of hypertrophy (cytochrome C oxidase polypeptide I and oligosaccharyltransferase,

both upregulated) and another set of genes whose expression was differentially

regulated between induction and regression (NADH dehydrogenase 6 and IGF-II,

upregulated during regression but downregulated during induction of cardiac

hypertrophy).

Chapter V

214

More recently, Yun et al. (2003) used transgenic mice overexpressing the α1b-

adrenergic receptor to identify genes whose expression was altered in a heart failure-

independent model of cardiac hypertrophy. Integrin α4, protein tyrosine phosphatase,

RAS GAP, phosphoinositide 3-kinase, uteroglobin, importin α, calbindin-28K, MEF2,

procollagen type XVIII and insulin II were all upregulated, whereas formin, FGF-7,

carboxypeptidase A3 and LIM were downregulated.

Ueno et al. (2003) used DNA microarray assays to study the development of blood

pressure-dependent heart failure in Dahl salt-sensitive rats fed a high (or low) sodium

diet. The development of cardiac hypertrophy and heart failure was associated with an

upregulation of genes such as ANP, actin and myosin isoforms, aldolase A, 12-

lipoxygenase and DBP transcription factor.

To restrict the set of genes to those specifically involved in the development of the

hypertrophic phenotype, Mirotsou et al. (2003) compared two different models of

cardiac hypertrophy to define genes concordantly regulated during the development of

cardiac remodeling in the LV or septum after MI or transverse aortic constriction

(TAC). Both MI and TAC resulted in cardiac and cardiomyocyte hypertrophy. Among

the genes that were upregulated in the LV and septum after MI but exhibiting little or

no change after TAC were the cell-cycle elements, cyclin D3 and tropomodulin 3, the

EGFR pathway substrate, growth arrest genes (specifically 1 and 2). In contrast genes

upregulated in the LV and septum after constriction but not after infarction included

genes encoding immune response proteins and genes involved in cytoskeleton

organization and biogenesis. Mirotsou et al. (2003) distinguished two additional groups

of genes: 1) those upregulated in the LV and septum after TAC and LV after MI, with

little or no change post MI in the septum (a group highly enriched with ECM genes and

genes involved in cell differentiation, proliferation and adhesion), and 2) genes

downregulated in LV and septum of both models (enriched for genes encoding

enzymes).

Chapter V

215

Larkin et al. (2004) evaluated alterations in cardiac gene expression in response to

pressure overload after acute (24h) and chronic (14 days) infusion of Ang II in mice. In

both groups the ribosome pathway was highly expressed, with upregulation of

ribosomal subunits, as well as initiation and elongation factors (eIF1A, iIF2B, eEF1-

α1). In addition to an increased ribosome and translation activity, Larkin et al. (2004)

identified a significant upregulation of genes linked to the RhoA-signaling pathway, the

ECM, cytoskeleton, cell cycle, steroid biosynthesis and the integrin-mediated signaling

pathway. Most of the genes were already documented to be increased in expression in

response to hypertension and Ang II treatment, but novel genes, such as follistatin-like

3 (Fstl3), epithelial membrane protein 1 (Emp1) and the transferrin receptor 1 (Tfr1)

were identified. Common downregulated pathways included: metabolic mitochondrial

fatty acid oxidation, citric acid cycle, carbohydrate metabolism, valine-leucine-

isoleucine degradation pathway and the oxidative phosphorylation. Other single

candidates included the non-muscle myosin heavy chain (MHC), SERCA2, estrogen

receptors, the peroxisomal protein (PeP) and the metabotropic glutamate receptor 1

(GRM1).

Recently, Rysä et al. (2005) used oligonucleotide microarrays to describe gene

expression profiles in the SHR during the development of LV hypertrophy (at 12

months) and the transition to diastolic heart failure (at 20 months). They identified 92

genes and 35 ESTs that were differentially regulated in the LV of 20 month SHR

compared to 12 month SHR. The relatively small number of altered genes agrees with

previous observations by Larkin et al. (2004), showing that more genes alter in

response to acute than chronic overload. The majority of upregulated genes encoded for

cell structure and signaling proteins (e.g. thrombospondin 4), whereas most of the

downregulated genes encoded for proteins involved in fatty acid and energy

metabolism (e.g. enoyl-CoA isomerase and acyl-CoA dehydrogenase).

Chapter V

216

1.3.4. Hereditary hypertrophic cardiomyopathy

The so-called ‘Toronto group’ used an extensive cardiac EST library to create a large

‘Cardiochip’ cDNA array to probe samples from patients with familial dilated and/or

hypertrophic cardiomyopathies associated with end-stage heart failure (Barrans et al.,

2002; Hwang et al., 2002). The genetic basis of inherited cardiomyopathy appeared to

revolve around genes involved in structural alterations in the sarcomere and

cytoskeleton, although some genes were also identified which have been found to be

upregulated or downregulated in other forms of cardiac hypertrophy and non-familial

heart failure (e.g. ANF, actins alpha and beta, elongation factor 2, hsp90, collagen type

I, ribosomal proteins S12, S6, L10, UBF2 transcription factor, SERCA2, elastin, fatty

acid binding protein, antigen CD36).

1.3.5. Uncomplicated and hypertension-associated human

obesity

Philip-Couderc et al. (2004) investigated heart transcriptome remodeling focusing on

changes specifically related to metabolic syndrome and obesity. In their study, they

used microarray analysis to compare right atrial samples from non-obese and obese

hypertensives and normotensive patients to discriminate between changes specifically

linked to obesity and arterial hypertension. The main finding of the study was that the

expression profile related to obesity dramatically differed to that observed in obesity-

related hypertension but also to that observed in arterial hypertension. One very

interesting example was the opposite change in genes encoding proteins involved in the

Wnt signaling pathway in obese and hypertensive patients. The Wnt pathway plays a

major role in cardiac myogenesis, hypertrophy and heart failure (Chapter I). Contrary

to observations in hypertension-related cardiac hypertrophy, obesity is characterized by

an overexpression of genes antagonizing and suppressing the Wnt pathway and

favoring β-catenin ubiquitination and degradation in proteasomes. Thus the expression

Chapter V

217

of several transcription factors that play a pivotal role in cardiac hypertrophy in other

contexts is suppressed in obesity (e.g. c-Myc, GATA4 and MEF2B).

1.4. In summary

With the development of the microarray technique and the advances in bioinformatics,

the candidate gene approach has been in part replaced with a broader, genomic scale

approach. However, analysis at this level can yield hundreds of genes that are

differentially regulated between control and experimental tissues, without indication of

their pathophysiological importance. Attempts at expression profiling in experimental

or human cardiac hypertrophy and failure have yielded ‘snapshots’ of differentially

regulated genes and sets of genes, with some overlapping (e.g. metabolic pathways

generally downregulated; structural proteins synthesis and cell cycle generally

upregulated). It is unclear whether these differences are important in mediating

pathophysiology or merely represent secondary and compensatory phenomena. These

findings emphasize the importance of an initial pathophysiological characterization of

the different experimental models of cardiac hypertrophy and the evaluation of model-

specific and condition-specific processes leading to heart remodeling prior to

undertaking gene profiling.

Chapter V

218

1.5. Aims

The goal of this study was to identify whether cardiac and cardiomyocyte remodeling

induced by chronic Ang II overproduction or impaired glucose transport in vivo could

have a primary and long-term impact on differential expression of the cardiac

transcriptome. It was hypothesized that the previously described chronic tissue and

myocyte remodeling in the TG1306/1R and GLUT4-KO mice would be associated with

a shift in expression of distinct sets of genes involved in various aspects of cell

physiology, morphogenesis and metabolism. It was also postulated that some

differentially regulated genes would be common to both models of remodeling.

Chapter V

219

2. METHODS

2.1. RNA and cDNA preparation for hybridization

Total RNA was extracted from 60 week old mouse ventricles (left and right ventricles

in addition to the septum) using the procedures detailed in Chapter II. The CyScribe

cDNA Post-Labelling Kit (Amersham Pharmacia, RPN5660) was utilized for the

preparation of Cy3- and Cy5-labelled cDNA for microarray hybridization according to

the manufacturer’s recommended protocol. Procedures are detailed in Chapter II.

2.2. cDNA microarray assays and experimental groups

2.2.1. Choice of array design

The microarrays used in the present study incorporate characteristics which are

particularly relevant in the investigation of differential gene expression induced by

cardiac remodeling in genetically manipulated mice. The NIA mouse embryonic clone

set printed on the microarray slides obtained from the Australian Genome Research

Facility

• is composed of embryonic clones, and thus allows the study of the possible re-

expression of embryonic genes, which has previously been observed during the

development of cardiac hypertrophy (and frequently interpreted to be

genetically a ‘fetal recapitulation’ state).

• consists of a substantial number of clones coding for unknown genes, making

these microarrays a dynamic tool to evaluate the expression of known and

unknown genes at the same time. As the mouse genome project progresses,

Chapter V

220

more ESTs will be linked to novel genes, unraveling new aspects of gene

expression and function in cardiac remodeling.

2.2.2. TG versus WT hearts

In the first set of experiments, comparison was made between 60 week male TG and

WT ventricles to screen for possible differences in gene expression induced by chronic

local overexpression of the Agt gene (gene ‘knock-in’). Eight mice (from different

parental crosses) were selected for each group. TG hearts were also selected according

to morphological criteria to avoid the presence of the dilated phenotype in the sample

population (see Chapter III). The mean CWI of the hearts were 4.7±0.2 mg/g for the

WT group (n=8) and 5.6±0.3 mg/g for the TG group (n=8). RNA from 2 ventricles

from the TG group were randomly pooled and directly compared with 2 pooled RNA

samples from the WT group. A dye-swap was performed for each comparison to

minimize the effects of any gene-specific dye-bias. This comparison was performed 4

times with different samples for a total of 8 cDNA slides hybridized (i.e. 4 replicates

with dye swap).

2.2.3. LLC versus LL hearts

In this experiment, comparison was made between 60 week male LLC and LL

ventricles to screen for possible differences in gene expression induced by the deletion

of the GLUT4 in the heart (gene ‘knock-out’). Eight mice (from different parental

crosses) were selected for each group. The mean CWI of the hearts used were 5.7±0.5

for the LL group (n=8) and 10.0±0.7 for the LLC group (n=8). RNA from 2 ventricles

of the LLC group were pooled and directly compared with 2 RNA samples pooled from

the LL group (the internal control group). Again, a dye-swap was performed for each

comparison to minimize the effects of any gene-specific dye-bias. This comparison was

performed 4 times with different samples for a total of 8 cDNA slides hybridized (i.e. 4

replicates with dye swap).

Chapter V

221

3. RESULTS

3.1. Differential gene expression and gene clustering in TG mice

3.1.1. Differential gene expression in TG mice

At 60 weeks chronic cardiac overexpression of Agt is associated with altered

expression of a number of genes. Compared with age-matched WT littermate controls,

TG mice exhibited differential expression of 181 clones in cardiac tissue, of which 91

were upregulated and 90 were downregulated. These data show that only ~1% of the

clone set printed on the microarray slide presented significant patterns of differential

expression after statistical normalization and ranking (M>1 or M<1 and B>0, see

Chapter II for further explanations). The 91 upregulated clones were successfully

linked to 72 annotated gene sequences and 9 unknown expressed ESTs. Similarly, of

the 90 downregulated clones, 76 known gene sequences and 3 ESTs were identified

(Tables V-1 and V-2).

3.1.2. Gene Ontology classification of candidates

Candidate genes (72 upregulated and 76 downregulated) were clustered according to

the Gene Ontology Consortium classification groups (Chapter II). Figures V-1, V-2 and

V-3 depict stacked bar charts illustrating differential distribution of GO terms for the

upregulated and downregulated genes according to the 3 common classifications: cell

component, molecular function and biological process. These data show that

considering all genes either up- or downregulated, ~37% of the selected genes coded

for cytoplasmic proteins (GO:0005737), ~21% for membrane proteins (GO:0016021)

and ~18% for nuclear proteins (GO:0005634) (Figure V-1). With respect to the

molecular function of the proteins translated from candidate genes, ~39% had binding

Chapter V

222

(or ligand) activity (GO:0005488), ~22% catalyzed biochemical reactions

(GO:0003824), ~10% functioned as transporter molecules (GO:0005215) and ~8%

transduced extracellular signals into the cells (GO:0004871) (Figure V-2). About 37%

of the putative proteins are involved in cellular metabolic regulation (GO:0008152),

~26% are involved in processes pertinent to the integrated physiology of the cell, such

as cell death, growth, defense or cell motility (GO:0050875), ~12% mediate interaction

of a cell with its surrounding, such as cell-to-cell communication or cell adhesion

(GO:0007154), and ~9% are known to modulate frequency, rate, or extent of a

physiological processes at the cellular or organ level (GO:0050791) (Figure V-3).

These data show that there is an even distribution of upregulated and downregulated

genes for each major GO category in TG ventricles (relative to WT), suggesting that

the pattern of differential upregulation and downregulation of genes in hypertrophic TG

hearts is relatively balanced. A statistical analysis across all GO categories at all levels

of classification (Chapter II) confirmed this general trend.

3.2. Differential gene expression and gene clustering in LLC mice

3.2.1. Differential gene expression in LLC hearts

Chronic suppression of myocardial GLUT4-mediated glucose uptake is associated with

altered expression of numerous genes at the 60 week time point. Compared with age-

matched LL littermate controls, LLC mice exhibited a differential expression of 96

clones in cardiac tissue, of which 50 were upregulated and 46 were downregulated.

These data showed that only ~0.5% of the clone set printed on the microarray slide

presented significant patterns of differential expression after statistical normalization

and ranking. The 50 upregulated clones were successfully linked to 39 annotated gene

sequences and 10 unknown ESTs. Similarly, of the 46 downregulated clones, 37 known

gene sequences and 5 ESTs were identified (Tables V-3 and V-4).

Chapter V

223

3.2.2. Gene Ontology classification in LLC hearts

Candidate genes (39 upregulated and 37 downregulated) were clustered according to

the Gene Ontology classification groups (Figures V-4, V-5 and V-6). The clustering of

candidate genes for GLUT4-KO hearts was relatively similar to the one observed for

TG1306/1R. Indeed, of all genes exhibiting differential regulation (either upregulated

or doenregulated), a majority of the selected genes encoded for cytoplasmic (~40%;

GO:0005737), membrane (~20%; GO:0016021) and/or nuclear proteins (~10%;

GO:0005634) (Figure V-4). At the level of the molecular function, the majority of the

candidate genes encoded proteins with binding (~37%; GO:0005488), catalytic (~30%;

GO:0003824), transporter (~20%; GO:0005215) and/or transducer (~5%; GO:0004871)

activity (Figure V-5). At a physiological level, the proteins encoded by the candidate

genes were involved in cell metabolism (~43%; GO:0008152), cell physiology (~29%;

GO:0050875) and cell communication (~10%; GO:0007154) (Figure V-6).

Interestingly, a statistical analysis performed across all GO categories at all levels of

classification (see Chapter II) showed that there was an uneven distribution of

upregulated and downregulated genes for the categories ‘mitochondrion’

(GO:0005739; a cellular component, with 12 downregulated genes vs. 3 upregulated,

p<0.05), ‘oxido-reductase activity’ (GO:0016491; a molecular function, with 5 genes

downregulated vs. 0 upregulated, p<0.05) and ‘ATP binding’ (GO:0006412; a

molecular function, with 8 genes downregulated vs. 2 upregulated, p<0.05). These data

suggest that the insulin resistant cardiomyopathy driven by the deletion of the insulin-

stimulated glucose transporter GLUT4 is associated with a significant downregulation

of genes encoding proteins involved in the mitochondrial energetic pathways.

Chapter V

224

3.3. Comparative analysis of TG1306/1R and GLUT4-KO

3.3.1. Common and distinct gene sets altered in TG and LLC

Gene expression profiling revealed both distinct and common sets of genes altered

during hormonally (TG) and metabolically (LLC) induced cardiac hypertrophies.

Among the 224 genes differentially regulated in LLC and TG ventricles, only 7

presented significant patterns of differential expression in both hypertrophic models.

The carboxylesterase 3 (Ces3), the mitochondrial cytochrome C oxidase 1 (mt-CO1)

and the mitochondrial 16S rRNA (mt-Rnr2) were found to be significantly

downregulated in both mouse models. The homologue of the human chromosome

condensation protein G (hCAPG, or 5730507H05Rik in the mouse), the CD2 antigen

binding protein 2 (Cd2bp2) and the ICOS ligand (Icosl) genes were upregulated in LLC

and downregulated in TG hearts. Finally, the fatty acid binding protein 3 (Fabp3) was

downregulated in LLC hearts and upregulated in TG.

3.3.2. Comparative GO analysis between TG and LLC

A comparative statistical analysis of upregulated and downregulated genes for the two

mouse models showed that TG hearts, when contrasted with LLC, exhibited a

significant upregulation of genes encoding for DNA-binding nuclear proteins

(GO:0005634; GO:0003677) involved in regulation of gene transcription

(GO:0006355) and protein biosynthesis (GO:0006412). TG hearts also showed

significant downregulation of genes encoding for proteins interacting with the

extracellular medium (GO:0003615) or involved in some type of receptor activity

(GO:0004872) when compared to LLC. LLC mice showed a significant

downregulation of genes encoding for proteins involved in mitochondrial proton

transport and ATP production when compared to TG hearts (GO:0005739;

GO:0015986; GO:0015992; GO:0016469; GO:0015078; GO:0046933; GO:0046961).

These results suggest that the major driving ‘signaling force’ underlying cardiac

Chapter V

225

hypertrophy in the TG hearts is protein synthesis, while remodeling in LLC hearts is

significantly influenced by downregulation of mitochondrial proteins and

mitochondrial pathophysiology.

Chapter V

226

4. DISCUSSION

4.1. Differential gene expression in TG1306/1R mice

In the present study, cDNA microarray expression profiling was used to identify genes

differentially expressed between TG and WT ventricles. Chronic overexpression of rat

Agt in the heart of 60 week old TG mice caused many genes to be altered in expression

(Table V-1) - a number exhibiting even higher M values than the Agt itself. Although it

is obvious that cardiac remodeling in TG hearts is primarily ‘triggered’ by chronic

overexpression of Agt, this study identifies novel candidate genes playing a major role

in the development of the cardiac phenotype in TG mice in response to the altered

myocardial humoral environment

4.1.1. Evidence for general transcriptional upregulation

in TG hearts

Although the cluster analysis suggested an even distribution of upregulated and

downregulated genes for each major GO category, a closer look at the list of the

differentially upregulated genes shows that in TG hearts several candidates code for

factors responsible for the de novo synthesis, folding and sorting of proteins. This is a

molecular trait that distinguishes the cardiac hypertrophy of TG mice from the LLC

cardiomyopathy. For instance, TG mice overexpress genes coding for various proteins

involved in DNA binding and transcription (Table V-1). Control of transcription by

RNA polymerase II involves the basal transcription machinery, a collection of proteins

which include TFIIB, TFIIE, TFIIF, TBP, TFIIA and TAFs. These, with RNA

polymerase II, assemble into complexes which are modulated by transactivator proteins

that bind to cis-regulatory elements located adjacent to the transcription start site. TG

mouse hearts overexpress the gene coding for the TATA box-binding protein-

Chapter V

227

associated factor 11 (TAF11), also known as TAFIID9, or TAFII28, suggesting that

gene transcription is enhanced in hearts from TG mice.

4.1.2. Evidence for enhanced protein synthesis in TG hearts

The protein translation factor SUI1 homolog (Sui1-rs1) codes for a protein related to

the translation initiation factor SUI1, a homologue of the human eukaryotic initiation

factor 1, eIF1. The latter is a low weight factor critical for stringent AUG (start codon)

selection in eukaryotic translation. It is recruited to the 40S ribosomal complex in the

multifactor complex (MFC) with eIF2 and eIF3 and eIF5 (Pestova et al., 2001; Singh et

al., 2004). Both these genes are upregulated in the TG (Table V-1). Sheikh et al. (1999)

demonstrated that the gene encoding eIF1 is induced during cellular stress and may also

represent an important adaptive response to ER stress. Interestingly, other specific

initiation factors are downregulated in TG hearts: the product of the gene Wbscr1 (also

known as Eif4h - Eukaryotic translation initiation factor 4h), stimulates protein

synthesis and it is highly expressed in heart, liver and testis. EIF4H stimulates the

RNA-dependent ATPase activities of eIF4A, eIF4B (also downregulated in TG hearts)

and eIF4F during translation initiation. It exerts its activity through protein-protein

interactions and possibly stabilizes conformational changes in eIF4A facilitating the

binding of the mRNA to the 40S ribosomal subunit (Richter-Cook et al., 1998).

Deletion of the Eif4h gene partly characterizes the multi-system developmental

disorder known as Williams-Beuren syndrome. This syndrome is characterized by a

constellation of features including mental retardation and cardiovascular diseases, such

as supravalvular stenosis, progressive left main coronary artery obstruction and

arrhythmias (Bruno et al., 2003). In general, the Williams-Beuren syndrome is caused

by the deletion of contiguous genes in the human locus 7q11.23, on chromosome 7,

which contains genes for elastin (ELN), LIM kinase-1 (LMK), RFC2, CYLN2,

CACNL2A, GPR56 and YWHAG, the latter also being downregulated in TG hearts. It

could be speculated that the specific downregulation of Wbscr1/Eif4h and Eif4b in TG

hearts is directly or indirectly associated with cardiac remodeling, and further

investigation is warranted.

Chapter V

228

Various elongation factors are upregulated in TG hearts. Examples are the eukaryotic

elongation factor-1 alpha 1 (Eef1a1) and delta (Eef1d). The alpha subunit of the

elongation factor-1 (eEF1α1) is a key factor in protein synthesis, where it promotes the

GTP-dependent transfer of aminoacylated tRNAs to the A-site of the 80S ribosomal

subunit (Bischoff et al., 2000). To do so, eEF1α1 needs first to interact with the

complex formed by the beta, gamma and delta (eEF1δ) subunits of the elongation

factor-1 to exchange bound GDP for GTP. Both Eef1a1 and Eef1d genes, encoding

eEF1α1 and eEF1δ proteins are upregulated in TG mouse hearts. In eukaryotic cells,

and in particular in human fibroblasts, the beta-gamma-delta complex colocalizes with

the endoplasmic reticulum (ER) and it is upregulated during ER stress responses, while

the alpha subunit shows a more diffuse distribution throughout the cytoplasm and it is

also associated with the nucleus (Sanders et al., 1996).

4.1.3. Overexpression of ribosomal subunits in TG hearts

Various components of the large 60S ribosomal subunit, such as Rpl10 (60S ribosomal

protein L10), Rpl5 (60S ribosomal protein L5), Rpl12 (60S ribosomal protein L12) and

Rpl4 (60S ribosomal protein L4) are upregulated in TG hearts. The mammalian

ribosome is composed of 4 RNA species and approximately 80 different proteins. In

eukaryotic cells, some ribosomes are free in the cytosol, whereas others are bound to

the ER. Membrane-bound ribosomes synthesize 3 major classes of proteins, that is

lysosomal proteins, secretory proteins and transmembrane proteins. Rpl10 gene

encoding for the 60S ribosomal protein L10 is also known as the QM gene. L10 binds

and negatively regulates c-Jun transcription activity by preventing homodimerization

between c-Jun units and therefore inhibiting transactivation of AP-1 transcription

factor-regulated promoters (Oh et al., 2002). L10 is differentially expressed in the

cardiac neural crest cells during embryonic heart development. In general, the pattern

of L10 expression suggests an inverse relationship with proliferative capacity and a

positive relationship with cell cycle arrest, differentiation and replicative senescence,

making it a tumor suppressor gene or a contributing factor in differentiation and aging

(Dimri et al., 1996; Mills et al., 1999). Thus, it could be speculated that an upregulation

of L10 in TG hearts could be associated with accelerated cellular senescence,

Chapter V

229

diminished cell proliferation and promotion of hypertrophy over hyperplasia in the

myocardium.

The cytosolic protein-synthesizing machinery is adapted to the protein-translocating

machinery in the ER membrane through ribonucleoproteins called signal recognition

particles (SRPs) (Larsen et al., 1998). A SRP complex consists of a 7S RNA and 6

different protein subunits. SRP14 and SRP9, the latter being upregulated in TG hearts,

constitute the Alu domain of 7S, whereas the other 4 proteins belong to the S domain.

While the S domain of SRP binds the N-terminal signal sequence on the nascent

polypeptide, the Alu domain temporarily interferes with the ribosomal elongation cycle

until the membrane translocation pore and the translocation machinery in the ER is

correctly engaged (Weichenrieder et al., 2001). SRP has at least 3 distinct functions

that can be associated with the protein subunits: signal recognition, translational arrest

and ER membrane targeting by interaction with the docking protein. Thus, an

overexpression of the Srp9 gene in hypertrophic hearts could be an indication of

increased trafficking of nascent proteins towards the plasma membrane via the ER, a

phenomenon that could contribute to ER stress.

4.1.4. Overexpression of chaperones in TG hearts

Heat-shock proteins (HSPs) are molecular chaperones which are essential for cell

survival. They play an important role in protein-protein interactions, including protein

folding, stability and turnover. HSPs have been named according to their molecular

weights. The family of approximately 70 kDa heat-shock proteins, forming the

complex HSC70, has been shown to be upregulated and synthesized under pathological

and stress conditions in the heart, including inflammation, ischemia and familial

hypertrophic cardiomyopathy (Hammerer-Lercher et al., 2001; Tanonaka et al., 2001).

TG mice present an upregulation of at least two members of the HSP70 family in their

hearts: Hspa8, coding for the HSPA8/HSP73 cognate protein, and Hspa5, coding for

HSPA5, also known as glucose-regulated protein 78 kDa, or GRP78.

Chapter V

230

HSPA8/HSP73 plays an important role in cells by transiently associating with nascent

polypeptides to facilitate correct folding and by functioning as an ATPase in the

disassembly of clathrin-coated vesicles during transport of membrane components

through the cell (Tavaria et al., 1995). It also interacts with tubulin and may protect

selected elements of the microtubule network to limit myofibril disruption during

cardiac remodeling (Decker et al., 2002).

HSPA5/GRP78 is a calcium-binding protein which is involved in the folding and

assembly of proteins in the ER (Nigam et al., 1994; Dierks et al., 1996). It may also

monitor protein transport through the cell. Barnes et al. (2000) demonstrated that

HSPA5/GRP78 is elevated in embryonic mouse hearts, it decreases significantly by the

fetal period and it can be re-induced following severe hypoglycemic stress, hence the

name ‘glucose-regulated protein’ - GRP. They concluded that HSPA5/GRP78 may play

a significant role in the normal differentiation and development of cardiac tissue and

that compensatory HSPA5/GRP78 induction may be involved in hypoglycemia-

associated myocyte dysmorphogenesis. Furthermore, HSPA5/GRP78 possesses an

ATP-binding domain which is also involved in the binding of the protein with

procaspase-7. Specific overexpression of HSPA5/GRP78 in cell lines (CHO, T cell or

carcinoma-derived) results in reduced apoptosis induced by caspase-7 and higher

colony survival when challenged with topoisomerase inhibitors, such as doxorubicin

and camptothecin (Reddy et al., 2003).

Thus, overexpression of this chaperone could provide protection against apoptosis in

TG hearts. Also, upregulation of Hspa5 could function as a mechanism to prevent or

limit cellular oxidative injury in TG. Indeed, Hung et al. (2003) demonstrated in renal

epithelial cells that transfection of cells with antisense RNA targeted against Hspa5

prevented the induction of HSPA5/GRP78, disabled the compensatory ER stress

response, sensitized the cells to oxidative injury (H2O2 toxicity), and prevented the

development of tolerance to H2O2 that normally occurs with preconditioning. Thus,

upregulation of genes involved in quality control of protein folding and calcium

buffering in the ER could be an adaptive response to chronic oxidative injury and

hypoxia

Chapter V

231

Various types of chaperones function in the ER to ensure proper synthesis of most

proteins. Some chaperones and escort proteins are highly specialized and are involved

in the folding of specific proteins. This is the case for the mesoderm development

candidate gene 2 (Mesdc2), also known for its Drosophila’s counterpart boca, which is

upregulated in TG hearts. Hsieh et al. (2003) demonstrated that Mesdc2, a gene

identified in the mesoderm development (Mesd) deletion interval on mouse

chromosome 7, is essential for specification of embryonic polarity and mesoderm

induction. They determined that the patterning and cell differentiation defects observed

in Mesd deletion homozygotes resulted solely from loss of the Mesdc2 gene. The latter

functions in the ER as a specific chaperone for the LDL receptor family members,

specifically Lrp5 and Lrp6, which in conjunction with frizzled are co receptors for the

canonical wingless (Wnt) signal transduction (Schweizer and Varmus, 2003). In the

absence (or mutation) of MESDC2 protein, LRP5 and LRP6 fail to reach the cell

surface and instead remain sequestered as insoluble aggregates in the ER. Thus, it could

be speculated that an upregulation of Mesdc2 in TG hearts could enhance membrane

sorting of LRP5 and LRP6 proteins, leading to activation of the Wnt signaling pathway

and also controlling cholesterol and glucose disposal in the hypertrophic heart, possibly

as a compensatory mechanism to preserve ATP/ADP ratio and energy production.

Interestingly, recent collaborative work at the University of Lausanne, Switzerland,

extends these findings and confirms upregulation of the Wnt signaling pathway in 50-

60 week old TG ventricles using cDNA microarray methodology (Domenighetti AA.,

data not shown).

4.1.5. Overexpression of calreticulin in TG hearts

Monoglucosylated oligosaccharides are recognized in the ER by two calcium-binding

lectins, calnexin and its soluble homologue calreticulin. Calreticulin is a

multifunctional protein that acts as a major calcium-binding/storage protein in the

lumen of the ER (Nigam et al., 1994; Llewellyn and Roderick, 1998). Calreticulin

alone retains misfolded or incompletely assembled proteins in the ER, where they are

degraded without transport to the Golgi complex (Zhang et al., 1997). In addition,

calreticulin and HSPA5/GRP78 can associate in an ATP-dependent manner in the ER

Chapter V

232

to retain misfolded protein fragments and to achieve cellular recovery from acute stress

(Zhang et al., 1997; Jethmalani and Henle, 1998). Thus, both proteins can combine to

serve as a quality control backup in retaining misfolded protein fragments in the ER.

Upregulation of both genes in the myocardium of TG mice would suggest that an

enhanced quality control in the pathway that leads to protein maturation is needed.

4.1.6. SR and ER stress: are they related phenomena?

As described in Chapter IV, cardiomyocyte remodeling and hypertrophy induced by

Ang II or metabolic disturbance is associated with impaired contractility and possible

dysfunctional calcium cycling by the SR. The results presented in this Chapter confirm

the pivotal role played by the endoplasmic reticulum (ER) in cell stress during Ang II-

induced cardiac hypertrophy. The ER stress response in TG hearts is certainly

associated with increased protein synthesis and trafficking through the reticulum. Also,

ER stress could function as a mechanism to prevent or limit oxidative injury in TG

hearts. From the present data it is not possible to determine which type of cell is subject

to ER stress in the heart.

Calreticulin is a functional homologue of calsequestrin in the ER of adult non-myocyte

cells, thus it could be first hypothesized that overexpression of calreticulin in TG hearts

is related to enhanced protein synthesis and ER stress in non-myocytes. Imanaka-

Yoshida et al. (1996) demonstrated that calreticulin is abundant in 2 week old fetal rat

cardiomyocytes and it is progressively replaced by calsequestrin after birth. This

suggests that maturation of cardiomyocytes involves the qualitative and/or quantitative

shift from non-muscle ER structures toward highly specialized muscle-specific SR

membranes. Calreticulin is also expressed in dedifferentiating adult cardiomyocytes in

long-term culture. These results indicate that calreticulin expression is downregulated

in cardiomyocytes during differentiation but can be expressed in vitro after

dedifferentiation. Further information comes from the study of calreticulin knock-out

mice (Mesaeli et al., 1999). The gene deletion is embryonic lethal and it is associated

with a marked decrease in the thickness of the ventricular wall (Mesaeli et al., 1999).

Study of the knock-out model indicated that calreticulin is abundantly expressed in

cardiomyocytes during early stages of development. While calreticulin-deficient

Chapter V

233

cardiomyocytes develop a functional SR and contract spontaneously, calreticulin

knock-out fibroblasts have ER impaired calcium homeostasis.

This indicates that SR and ER may exist as functionally distinct compartments in

cardiomyocytes. In addition, studies on the formation of SR and triads during skeletal

and cardiac muscle development suggest that the SR is not just a continuation of the

existing ER membranes but can be physically distinct (Franzini-Armstrong and

Jorgensen, 1994; Flucher and Franzini-Armstrong, 1996). At the functional level, it has

been demonstrated that ER calcium handling linked with the IP-dependent pathway and

nuclear translocation of NF-AT are impaired in calreticulin deficient cells. This

suggests that calreticulin is a part of the calcineurin/NF-AT/GATA-4 pathway

described for cardiac hypertrophy (see Chapter I) and that this unique pathway is also

activated during cardiac development by the ER. Taken together, these various findings

indicate that the ER membrane plays a critical role not only in protein synthesis,

modification and secretion, but also in control of calcium homeostasis in the

developing heart. Thus, it is tempting to speculate that ER membrane proteins such as

calreticulin play a role in the development of cardiomyocyte hypertrophy in TG hearts.

Stress-dependent stimulation of cardiomyocytes could result in a qualitative and

quantitative modification of ER and SR membrane structures and activation of both the

NF-AT and calreticulin pathways.

4.1.7. Differential regulation of Wnt signaling in TG hearts

As detailed in Chapter I, members of the Wnt gene family of secreted glycoproteins are

involved in different developmental processes such as cell differentiation, migration,

adhesion and organ development. The canonical Wnt/β-catenin pathway leads to the

stabilization of cytoplasmic β-catenin which subsequently enters the nucleus where, in

combination with various transcription factors, it regulates gene expression. A second

Wnt signaling pathway, triggered by Wnt5a (and possibly other Wnt isoforms),

stimulates the intracellular increase of calcium via PKC- and/or calmodulin-dependent

activation. These Wnt signaling pathways have been previously linked to the

development of the cardiovascular system (Chapter I). The present data suggest that

both the activity and the regulation of Wnt signaling pathways are enhanced in TG

Chapter V

234

hearts. Indeed, TG ventricles overexpress Mesdc, which is involved in the sorting of

LRP5 and LRP6, which have well known co-receptors of the canonical Wnt pathway.

In parallel, TG hearts exhibited downregulated Lrp1 encoding a suppressor of the same

pathway. Among other genes, the canonical Wnt-repressor adenomatous polyposis coli

gene (or Apc) is downregulated, while a subunit of the protein phosphatase 2A

(Ppp2r1a), a positive regulator of the Wnt signaling pathway, is upregulated. Finally,

TG ventricles show a positive regulation of the Wnt regulator Dickkopf3 (Dkk3) and

the metalloproteinase Mmp7 (known also as ‘matrilysin’ or ‘uterine’), the latter being a

target gene of the canonical Wnt signaling pathway (Crawford et al., 1999).

Interestingly, phospholipase C, delta subunit (Plcd), previously shown to be involved in

the calcium rise induced by Wnt activation, is downregulated in TG hearts.

In addition, calreticulin (overexpressed in TG hearts as discussed above) has been

shown to increase cell adhesiveness via activation of the Wnt pathway leading to

dephosphorylation of β-catenin and overexpression of N-cadherin (Fadel et al., 2001).

As noted above, additional experiments using cDNA microarray technology have been

undertaken to characterize alterations in Wnt pathway signaling in more detail. These

more extensive Wnt pathway-directed investigations have revealed TG upregulation of

Wnt1, Wnt3, Wnt3a, Wnt4, Wnt7a, Wnt10b and Wnt11 isoforms, together with

overexpression of disheveled 2 (Dvl2), frizzled 1 (Fzd1) and various phosphatase

subunits, which could dephosphorylate β-catenin and induce gene expression. Target

genes of the Wnt signaling such as Mmp7, cyclin D1 (Ccnd1) and bone morphogenic

protein 4 (Bmp4) were also upregulated on the complimentary microarray

(Domenighetti A., data not shown). Taken together, these data suggest that the Wnt

signaling pathways are upregulated in ventricles from TG mice. The reasons are

unknown, but possible re-activation of specific pro-growth fetal programs of gene

expression is a valid hypothesis. Haq et al. (2003) demonstrated that β-catenin is both

sufficient to induce growth in cardiomyocytes in culture and in vivo and is necessary

for hypertrophic stimulus-induced growth.

Chapter V

235

Interestingly, Bond et al. (2003) showed that endothelin-1 stimulates the expression of

Wnt11 and Wnt7a in differentiating cardiac cells destined for functional specialization

in the electrical conduction system in the developing heart. Marvin et al. (2001) found

that suppression of Wnt signaling by Dikkopf1 and/or Crescent promotes heart

formation in the anterior lateral mesoderm in chick embryos, whereas active Wnt

signaling in the posterior lateral mesoderm promotes blood cell development. They also

showed that Wnt signals can repress heart formation from anterior mesoderm in vitro

and in vivo, while forced expression of either Wnt3a or Wnt8c in the same region can

promote development of primitive erythrocytes from the precardiac region. Intriguingly

TG hearts overexpress both adult and embryonic forms of hemoglobin genes (Hbb-a1,

Hbb-b2, Hbb-y) indicating that a form of cell differentiation leading to erythropoiesis,

rather that myogenesis could be activated in these hearts.

4.1.8. Cytoskeletal remodeling in TG hearts

In TG hearts there is upregulation of genes involved in cytoskeletal rearrangements,

including the ARF1 GTPase activating protein 1 (Arfgap1) and the protein kinase 9

(Ptk9), also known as Twinfilin-1.

Furman et al. (2002) demonstrated that the actin-bound protein ARFGAP1 co-localizes

with focal adhesions and it can function as a regulator of the cytoskeleton. ARFGAP1

upregulation interferes with formation of focal adhesions and with PDGF-induced

membrane ruffling in 3T3 cells. Also, ectopic ARFGAP1 enhances 3T3 cell migration

toward PDGF or IGF-1 and increases cell spreading in general. Thus, it could be

speculated that an upregulation of the gene coding for ARFGAP1 in TG hearts could

enhance or stimulate non-myocyte motility and ‘myocyte slippage’ within the

myocardium, contributing to dysmorphogenesis.

Ptk9/Twinfilin is a ubiquitous actin monomer-binding protein that is composed of two

ADF-homology domains. It forms a 1:1 complex with ADP-actin monomers, inhibits

nucleotide exchange on actin monomers and prevents assembly of the monomers into

filaments. Vartiainen et al. (2003) showed that mammals have two forms of twinfilin.

Chapter V

236

Twinfilin-1 is the major isoform in embryos and in most adult mouse non-muscle cell-

types, whereas twinfilin-2 is the predominant isoform of adult heart and skeletal

muscles. They are regulated by distinct cellular signaling pathways. Studies on budding

yeast suggest that twinfilin contributes to actin filament turnover by localizing actin

monomers, in their ‘inactive’ ADP form, to the sites of rapid filament assembly. This is

mediated through direct interactions between twinfilin and capping proteins (Falck et

al., 2004). Thus, an upregulation of twinfilin-1 in TG hearts could be associated with

re-expression of the fetal isoform in myocytes or an upregulation of this isoform in

non-myocyte cells. Either way, it could be speculated that re-expression (or

upregulation) of twinfilin-1 in TG hearts leads to cytoskeletal rearrangement and

remodeling, possibly in association with the re-expression of the fetal isoform alpha-

skeletal actin in the same mouse model (Clement et al., 2001).

4.1.9. Differential protein turnover and catabolism in TG hearts

Selective protein degradation plays an important role in cellular regulation, since it can

protect cells against environmental stress by eliminating aberrant proteins generated

under physiologic and pathologic conditions (Hochstrasser, 1996; Patton et al., 1998).

In eukaryotes, selective protein degradation proceeds primarily through the enzymes of

the ubiquitin conjugation system, which requires essential action of 3 enzymes: an

activating enzyme E1, a conjugating enzyme E2 and a ligase E3 (Hochstrasser, 1996).

TG hearts exhibit an upregulation of the gene Ube4 which codes for a novel

ubiquitination factor E4. The latter has been linked to stress tolerance and degradation

of stress-induced aberrant proteins. Substrates of an E4-dependent degradation pathway

are supposed to be proteins whose stabilization does not interfere with vital functions of

the cell under normal growth conditions. However, E4-dependent degradation becomes

crucial when cells are exposed to stress. Indeed, the yeast homologue Ufd2 is

functionally implicated in cell survival under stressful conditions. Histones and actin

are among the substrates whose turnover might be regulated by E4 (Koegl et al., 1999).

Meacham et al. (2001) demonstrated that for mutant forms of the cystic-fibrosis

transmembrane-conductance regulator (CFTF), a plasma-membrane chloride channel,

the chaperone CHIP converts the heat-shock complex 70 (Hsc70) from a protein-

folding machine into a degradation factor that functions in ER quality control through

Chapter V

237

an E4-dependent pathway. These data reinforce the idea that TG hearts suffer from

manifest ER and more generalized cell stress, probably caused by enhanced protein

biosynthesis. Upregulation of the Ube4 in TG hearts is associated with the positive

regulation of the Apc7, which is one of the subunits composing the anaphase promoting

complex, or cyclosome, involved in protein turnover and cell cycle progression.

Interestingly, the positive regulation of the E4-dependent degradation pathway in TG

hearts is associated with downregulation of the ubiquitin protein ligase E3b (Ube3b),

the ubiquitin-conjugating enzyme E2G2 (Ube2g2) and the subunits Psmd2 and Psmd13

of the proteasome complex. The latter is a key enzymatic complex responsible for

selective turnover of short-lived and misfolded proteins. These data suggest that

cellular remodeling in TG hearts is associated with differential downregulation of

proteolytic pathways in favor of a stress-mediated ubiquitination pathway.

4.1.10. Are sex hormones involved in cardiac remodeling?

One of the most upregulated genes in TG hearts encodes for the Estradiol 17 beta-

dehydrogenase 4 protein (HSD17B4) This enzyme is one of the isoforms of the

estradiol 17 beta-dehydrogenase enzyme family that is responsible for the

interconversion of estrone and estradiol, as well as the interconversion of

androstenedione and testosterone (Andersson and Moghrabi, 1997). In general,

HSD17B4 inactivates the sex steroids by catalyzing the production of the keto forms. It

exhibits catalytic preference for oxidation with cofactor preference for NAD+ and

substrate preference for C18 steroids (i.e. preference for estrogens over androgens). It is

the first steroid metabolizing enzyme found to be localized in peroxisomes (in contrast

to the other HSDs which are found in microsomes or in the cytoplasm).

HSD17B4 is also known as D-bifunctional protein (DBP). HSD17B4 catalyzes the

formation of 3-ketoacyl-CoA intermediates from both straight-chain and 2-methyl-

branched-chain fatty acids and therefore it is involved in peroxisomal beta-oxidation of

fatty acids (Moller et al., 2001). Indeed, mutations or deletions in HSD17B4 have been

shown to cause defects in liver and fibroblast peroxisomal beta-oxidation, with

accumulation of both very long-chain fatty acids and bile acid intermediates, as it is

observed in Zellweger syndrome or in adrenoleukodystrophy (de Launoit and Adamski,

Chapter V

238

1999; Moller etal., 2001). As oxygen radicals are both produced and scavenged in

peroxisomes, evidence for increased oxidative stress in HSD17B4 deficiency has also

been reported (Ferdinandusse et al., 2003).

HSD17B4 has yet to be fully characterized for capacity for stimulation by modulators

of both steroid and fatty acid metabolism, such as progesterone and peroxisome

proliferators - the latter mechanism being dependent on activation of peroxisome

proliferator-activated receptor alpha (PPARα) (see Chapter I). Accordingly, in PPARα

knock-out mice the expression of HSD17B4 is low and does not change after treatment

with peroxisome proliferators, such as WY 14643 (Aoyama et al., 1998). Structurally,

HSD17B4 is expressed in many tissues as an approximately 3.0-kb transcript, with

highest expression in liver, kidney, ovary, lungs, heart, prostate, and testis. The enzyme

is primarily translated as an 80 kDa protein. The post-translational modifications

include an N-terminal cleavage leading to a 32 kDa peptide, which is covalently bound

to actin (Leenders et al., 1994). The N-terminal part shows homologies to the family of

short chain alcohol dehydrogenases, especially to the two short chain alcohol

dehydrogenases domains of the multifunctional (hydratase-dehydrogenase) enzymes of

peroxisomal beta-oxidation of fatty acids. The C-terminal extension of the 80-kDa

protein shows an intriguing similarity to the sterol carrier protein 2 (SCP2) which is

assumed to participate in the intracellular transport of sterols and lipids (Leenders et al.,

1994). In SCP2 knock-out mice marked alterations in gene expression, peroxisome

proliferation, hypolipidemia, impaired body weight control and neuropathy were

observed (Seedorf et al., 1998). The gene disruption led to impaired catabolism of

methyl-branched fatty acyl-CoAs and inefficient import of phytanoyl-CoA into

peroxisomes.

Although no information is available concerning the role played by HSD17B4 in the

myocardium, estrogens have been previously shown to confer cardio-protection.

17β-estradiol has been shown to attenuate the development of pressure-overload

cardiac hypertrophy in mice (van Eickels et al., 2001), while the same steroid has been

shown to activate ERK1/2 and JNK (but only marginally p38) MAPK pathways

through an estrogen receptor-dependent mechanism in isolated adult rat

cardiomyocytes (Nuedling et al., 1999). Thus it could be hypothesized that during Ang

Chapter V

239

II-induced cardiac hypertrophy, an increased conversion of estradiol to the less-active

estrone by HSD17B4 overexpression (and induction), could lead to decrease in tissue

(and serum) estradiol levels with suppression of some of the cardio-protective effects of

estrogens mentioned above. Hence, the diminished cardio-protective role played by

estrogens could be associated with decreased availability of free fatty acids at the

cellular level and therefore decreased availability of energy for cardiomyocyte function.

Alternatively, it is known that peroxisomal beta-oxidation enzymes acyl-coA oxidase

and HSD17B4 can produce a retro-conversion pathway for the production of

docosahexaenoic acid (DHA), a long-chain polyunsaturated fatty acid of the omega-3

family (C22:6n-3) (Su et al., 2001). DHA is relatively abundant in fish oil, and has long

been recognized for its beneficial effects on the cardiovascular system. Clinically, DHA

and other long-chain polyunsaturated fatty acids are believed to have anti-hypertensive

properties by inhibiting ACE activity (and Ang II formation) and enhancing endothelial

NO production. Also, perinatal supplementation of DHA and other long-chain

polyunsaturated fatty acids decrease insulin resistance and prevent the development of

hypertension in adult life (Diep et al., 2002; Engler et al., 2003; Das, 2004). At the

cardiomyocyte level, DHA inhibits calcium sparks and RyR activity in rat and sheep

ventricular myocytes (Honen et al., 2003) and can directly or indirectly block several

calcium and potassium ionic currents and therefore modulate contraction (Fournier et

al., 1995; Jude et al., 2003). Thus, an overexpression of HSD17B4 in the heart could be

associated with enhanced DHA production, leading to NO- and bradykinin-induced

vasodilation and to the modulation of E-C coupling in cardiomyocytes. These effects

could be a compensatory mechanism to counterbalance the increased cellular oxidative

stress and the remodeling effects of high levels of Ang II produced in the TG hearts.

4.1.11. Metabolic disturbances in TG hearts

TG mice exhibit differential regulation of genes involved in various metabolic

processes of the heart.

Among the upregulated processes, there is also evidence of increased

• purine and pyrimidine metabolism, with upregulation of the ribonucleotide

reductase M2 (Rrm2),

Chapter V

240

• N-glycan degradation (Aga),

• nitrogen metabolism, with upregulation of the carbonic anhydrases 8 and 14 (Car8

and Car14).

• nuclear-coded mitochondrial oxidative phosphorylation, with upregulation of the

cytochrome c oxidase subunit VII Cox7a2l and cytochrome b of complex III (mt-

Cytb).

Downregulated processes apparently include:

• glycolysis and gluconeogenesis, with downregulation of the phosphofructokinase

(Pfk1),

• mitochondrial oxidative phosphorylation, with downregulation of the mitochondrial

cytochrome c oxidase 1 of complex IV (mt-Co1),

• flux through the pentose phosphate pathway (Pfk1),

• fructose/mannose/galactose metabolism (Pfk1),

• mitochondrial fatty acid metabolism, with downregulation of the acetyl-coA

dehydrogenase (Acads),

• chondroitin/heparan sulfate metabolism, with downregulation of the carbohydrate

sulfotransferase 11 (Chst11),

• the inositol phosphate metabolism, with downregulation of the phospholipase C,

delta, (Plcd1),

• valine/leucine/isoleucine degradation, with downregulation of the branched chain

keto-acid dehydrogenase E1 (Bckdha) and Acads,

• citrate cycle, with downregulation of the citrate synthase (Cs),

• glyoxylate and dicarboxylate metabolism (Cs),

• butanoate metabolism (Acads) and

• sulfur metabolism (Chst11).

In overview, TG hearts show differential expression of key enzymes involved in

multiple metabolic processes, such as glycolysis and fatty acid oxidation. This suggests

that cardiac and cardiomyocyte remodeling in TG hearts is linked with important

metabolic adaptations. As detailed in Chapter I, during cardiac remodeling and

hypertrophy the heart is believed to switch from free fatty acid to glucose metabolism,

in a recapitulation of the embryonic metabolic state. The data reported here partially

Chapter V

241

disagrees with this ‘generic’ concept, suggesting that the decrease in mitochondrial

fatty acid metabolism is also associated with a decrease in the glucose metabolism. The

downregulation of mt-Co1 in the electron transport chain could be balanced by the

upregulation of mt-Cyb and Cox7a2l. Interestingly, there is a possible compensatory

activation of the peroxisomal free fatty acid beta oxidation (with the upregulation of the

HSD17B4 enzyme).

4.1.12. Carbonic anhydrase and NHE-1 upregulation

Carbonic anhydrases form a large family of genes encoding zinc metalloenzymes are of

great physiological importance. As catalysts of the reversible hydration of carbon

dioxide, these enzymes participate in a variety of physiological processes, including

respiration, calcification, acid-base balance, bone resorption and the formation of

aqueous humor. TG hearts overexpress the genes encoding carbonic anhydrases 8 and

14 (CAR8 and CAR14). cDNA microarray studies (University of Lausanne) not

mentioned in detail here suggest there is upregulation of three other carbonic anhydrase

enzymes, namely CAR2, CAR4 and CAR6. CAR2 and other isoforms bearing CO2

hydration activity enhance the activity of the NHE-1 and the anion exchanger AE3 in a

synergistic way (Figure V-7) (Li et al., 2002; Alvarez et al., 2003; Loiselle et al., 2004).

Interestingly, Alvarez et al. (2004) proposed that catecholamine-induced

cardiomyocyte hypertrophy in vitro can be suppressed by treating the cells with ETZ,

an inhibitor of the carbonic anhydrases, suggesting an important role played by

synergistic activity of carbonic anhydrases, NHE-1 and anion exchangers in the process

of cardiomyocyte functional remodeling. As shown in Chapter IV (Figure IV-8), TG

hearts upregulate the NHE-1. The present data strengthen the idea that sodium-

hydrogen exchange is enhanced in hearts from Agt-overexpressing transgenic mice,

leading to cellular sodium load and cardiomyocyte hypertrophy.

Chapter V

242

4.2. Differential gene expression in LLC hearts

Microarray expression profiling was used to identify genes differentially expressed

between LLC and LL ventricles. In hearts of 60 week LLC, genetically manipulated

ablation of the glucose transporter, GLUT4, was associated with altered expression of

numerous genes, including the hypertrophic markers BNP (Nppb) and SERCA2

(Atp2a2), which are up- and downregulated respectively in LLC hearts (Tables V-3 and

V-4). This study provides novel insights into the genetic alterations underlying the

development of insulin resistant cardiomyopathy.

4.2.1. Metabolic and mitochondrial impairment in LLC hearts

The most evident shift in gene expression in LLC ventricles is a general

downregulation of ‘key players’ of the mitochondrial oxidative phosphorylation. LLC

hearts exhibit downregulation of proton-transporting ATP synthase subunits Atp5b,

Atp5c1 and Atp5j (complex V) as well as the NADH dehydrogenase subunits Ndufs2

and Ndufb9 (complex I) and the cytochrome c1 oxidase (mt-Co1) (complex IV). As a

result, ATP production would be expected to be impaired in the LLC hearts. These data

suggest that cardiac remodeling in insulin resistant cardiomyopathy is associated with

mitochondrial dysfunction, possibly linked with the disruption of the glucose

metabolism following GLUT4 deletion. Other evidence of mitochondrial impairment in

LLC hearts comes from the downregulation of acetyl-Coenzyme A acyltransferase 2

(Acaa2), which is involved in the fatty acid beta oxidation. Interestingly, the

cytoplasmic acyl-CoA synthetase Acsl5, which activates fatty acids and requires ATP,

is upregulated. Together, these data propose a disruption of the mitochondrial electron

chain transport, decreased ATP synthesis and downregulation of beta-oxidation, which

is consistent with the interpretation that fatty acid metabolism is downregulated during

cardiac hypertrophy.

LLC hearts also exhibit downregulation of soluble malate dehydrogenase 1 (Mdh1) and

isocitrate dehydrogenase 2 (Idh2), suggesting impaired citric acid cycle and decreased

Chapter V

243

pyruvate and glyoxylate/dicarboxylate metabolism, with decreased CO2 fixation.

Interestingly, the glyceraldehyde-3-phosphate dehydrogenase (Gapd) involved in

glycolysis and gluconeogenesis, is upregulated in LLC hearts, suggesting that the

glycolytic pathway may be enhanced in LLC hearts. Together these data could suggest

that there is a shift in the production of ATP from the aerobic pathway, involving

mitochondrial oxidation of NADH equivalents, to an anaerobic pathway, where glucose

and glycogen are metabolized to pyruvate and then reduced to lactate by the lactate

dehydrogenase. This could be a compensatory mechanism to salvage ATP production

in an environment characterized by decreased glucose intake and possibly low oxygen

levels. Lactate and proton accumulation in the heart could potentially induce changes in

pH linked with activation of the sodium/hydrogen exchanger (NHE) (Chapter IV),

which in turn could induce the decrease of GLUT4 translocation toward the membrane

and further reduce myocyte glucose uptake (Yang et al., 2002).

There is evidence that other metabolic pathways are differentially regulated in LLC

hearts. For instance, downregulation of the methionine adenosyltransferase I alpha

(Mat1a) could decrease methionine and the selenoamino acid metabolism, while

downregulation of Acaa2 (see above) could also prevent the degradation of valine,

leucine and isoleucine, as well as the degradation of benzoate. The dual-specificity

tyrosine-phosphorylation regulated kinase Dyrk1a, also downregulated in LLC hearts,

is linked to the starch/sucrose metabolism and the nicotinate/nicotinamide metabolism.

Dyrk1a, together with the kinase GSK-3, is known to modulate glycogen synthesis by

phosphorylating glycogen synthase (Skurat and Dietrich, 2004). Downregulation of

Drk1a would decrease glycogen synthase phosphorylation and enhance glycogen

storage. This is a surprising and unexpected finding as increased glycogen breakdown

and suppressed glycogen synthesis is generally observed in type 2 diabetic skeletal

muscles, in association with insulin resistance and diminished glucose disposal.

However, if similar expression shifts occur in the LLC skeletal muscle, this would

explain the lack of hyperglycemia in the GLUT4-KO mice (Kaczmarczyk et al., 2003).

Chapter V

244

4.2.2. Metal and iron homeostasis in LLC hearts

Hearts of LLC mice exhibit upregulation of proteins playing a central role in iron

homeostasis. The solute Carrier 11 alpha 2 (Slc11a2), is best known as natural

resistance-associated macrophage protein 2 (NRAMP2). It is a proton-dependent cation

transporter, which plays an important role in iron homeostasis. NRAMP2 is present in

the plasma membrane and in acidified endomembrane compartments (Tabuchi et al.,

2000). Their subcellular colocalization and parallel trafficking suggest that NRAMP2

and transferrin receptors are functionally coupled to effect pH-dependent iron uptake

across the endosomal membrane (Touret et al., 2003).

Ferritin light chain 1 (Ftl1) is also upregulated in LLC hearts. Ferritin is the major

intracellular iron storage protein in all organisms. This gene has been shown to be

upregulated in rat heart after myocardial infarction (Zhu et al., 2000). Ferritin

accumulation in the heart causes alterations in systolic and diastolic function and can

provide an arrhythmogenic substrate inducing atrial and ventricular tachyarrhythmia

(Parkes et al., 1993). Metal ions are essential cofactors for a variety of biologic

processes, including oxidative phosphorylation, gene regulation, and free-radical

homeostasis. Iron-overload is associated with hereditary disorders such as

hemochromatosis, which causes cirrhosis of the liver, hepatocarcinoma, diabetes,

hypermelanotic pigmentation of the skin, cardiomyopathy, myocardial inflammation

and heart failure (Burke et al., 2001; Imperatore et al., 2003). Interestingly, Oudit et al.

(2003) showed that iron-overload in mice was associated with increased myocardial

fibrosis and elevated oxidative stress, in association with increased mortality, systolic

and diastolic dysfunction, bradycardia and hypotension. They also showed that L-type

voltage-operated calcium channels were key transporters of iron into cardiomyocytes

under iron-overloaded conditions. Thus, it could be speculated that cardiomyopathy in

LLC mice is associated with (and possibly induced by) non-hereditary form of iron-

overload.

Chapter V

245

4.2.3. Genes associated with hereditary and non-hereditary

cardiomyopathy in LLC hearts

Differential regulation of genes that have been previously associated with various

forms of hereditary cardiomyopathy (see Chapter I for more details) are observed in

LLC hearts. A gene which is commonly implicated in hereditary cardiomyopathy is

tropomyosin 1 alpha chain (Tpm1). This gene is downregulated in LLC hearts. It

usually binds to actin filaments in muscle and non-muscle cells and plays a central role,

in association with the troponin complex, in the calcium dependent regulation of

vertebrate striated muscle contraction (Murakami et al., 2005). In non-muscle cells it is

involved in stabilizing cytoskeleton actin filaments. Mutations and downregulation of

this gene causes hypertrophic cardiomyopathy CMH3 in humans.

Another gene involved in heart defects in Down’s syndrome, is the SH3 domain-

binding glutamic acid-rich protein (Sh3bgr). This gene maps to chromosome 21 within

the Down’s syndrome congenital heart disease minimal region in humans. The

expression of Sh3bgr is restricted to the heart between embryonic E7.75-E10.5 and it

plays a possible role in heart morphogenesis (Egeo et al., 2000). This gene belongs to a

new family of highly conserved small proteins related to the thioredoxin super family

(Mazzocco et al., 2002). Upregulation of this gene in Down’s syndrome (due to the

presence of an extra copy of chromosome 21) causes important cardiac defects which

ultimately reduce the life expectancy of Down’s individuals. Recently, Sandri et al.

(2004) reported that transgenic mice overexpressing the Sh3bgr gene do not present

any morphogenetic and developmentally regulated heart defect. Nevertheless, non-

familial upregulation of Sh3bgr and downregulation of Tpm1 could play an important

part in the development of the cardiomyopathic phenotype in LLC hearts.

Chapter V

246

4.2.4. Adrenomedullin receptor overexpression in LLC hearts

Transcription of the gene coding for the receptor activity-modifying protein 2 precursor

is upregulated in LLC hearts. This gene transports the calcitonin-receptor-like receptor

to the plasma membrane where it acts as an adrenomedullin receptor (Bomberger et al.,

2005). The calcitonin-receptor-like receptor binds calcitonin gene-related peptide one

of the most potent known endogenous vasodilators known. Calcitonin gene-related

peptide and its related peptide adrenomedullin are potent smooth muscle relaxants in a

variety of tissues and upregulation of adrenomedullin and the receptor activity-

modifying protein 2 precursor in the myocardium and aorta may be significant in the

pathogenesis of ischemic cardiomyopathy. Upregulation of the cardiac adrenomedullin

system, including the ligand, receptor, and activity, may modulate pathophysiology

during the transition from LV hypertrophy to heart failure in hypertensive rats

(Romppanen et al., 1997). Furthermore, adrenomedullin stimulates heart rate, cardiac

output, plasma levels of cAMP, prolactin, norepinephrine and renin whilst inhibiting

any concomitant response in plasma aldosterone (Troughton et al., 2000). Acute

hyperinsulinemia induces a significant increase in the plasma levels of adrenomedullin

in patients with type 2 diabetes mellitus, thus suggesting that increased plasma insulin

levels may regulate circulating levels of adrenomedullin in patients with type 2 diabetes

mellitus (Katsuki et al., 2002). Furthermore, the adrenomedullin response is

upregulated by hypoxia and inflammatory stimuli and may play an anti-inflammatory

role via a PI3K pathway (Okumura et al., 2004). All these data could suggest that

overexpression of receptor activity-modifying protein 2 precursor in the heart of LLC

mice could enhance adrenomedullin activity at the vascular and myocardial levels.

4.2.5. Cell cycle regulation and fibroblast proliferation

in LLC hearts

As suggested in Chapter III, collagen production in LLC hearts could be associated

with fibroblast proliferation. The gene encoding the S100 calcium binding protein A6

(S100A6), also known as calcyclin, is upregulated in the LLC ventricles. Calcyclin, a

Chapter V

247

prolactin receptor-associated protein that binds calcium and zinc ions is highly

expressed in fibroblasts and epithelial cells of various organs, including the heart

(Kuznicki et al., 1989; Kuznicki et al., 1992). Breen and Tang (2003) reported that

calcyclin expression modulates calcium-dependent signaling events that regulate

progression through the cell cycle and induce proliferation of fibroblasts. Calcyclin also

seems to interact with other proteins in a calcium-dependent manner, such as GAPD,

annexin II and a protein CacyBP/SIP (Zeng et al., 1993; Filipek et al., 1995; Nowotny

et al., 2003). Interestingly, CacyBP/SIP is a component of a novel ubiquitinylation

pathway regulating β-catenin degradation (Filipek et al., 2002) and inhibiting the

canonical Wnt signaling pathway. LLC hearts upregulate another mitotic checkpoint

gene, the budding uninhibited by benzimidazoles 1 homolog beta gene (Bub1b). Bub1b

codes for a protein that can target other proteins for degradation by proteasomes during

mitosis (Davenport et al., 1999). In synchronized cells, expression of Bub1 gene peaks

in G2/M phase, suggesting that it controls transition from the G phase to mitosis.

Finally, the most upregulated gene in LLC hearts encodes for the n-Myc downstream

regulated gene 4 (Ndr4). This gene encodes for a protein which is identical to the

mouse cardiac-specific SMAP8 (Nishimoto et al., 2003). Interestingly, Nishimoto et al.

(2003) showed that PDGF-induced proliferation was significantly enhanced in SMAP8-

transfected smooth muscle-derived cell lines compared to control. Consistent with this,

Ciani et al. (2004) demonstrated that downregulation of n-Myc negatively regulates

proliferation and promotes neuronal differentiation in neuroblastoma cell lines. These

data strengthen the hypothesis that collagen production in LLC hearts is associated with

fibroblast proliferation and suppression of differentiation processes. Proliferation could

be associated with the activation of novel ubiquitinylation pathways leading to mitosis

and to the inhibition of the Wnt signaling pathways that regulate differentiation.

Chapter V

248

4.3. Common set of genes altered in TG and LLC hearts

Gene expression profiling revealed a large number of distinct genes altered during

hormonally (TG) and metabolically (LLC) induced cardiac hypertrophies. Among the

224 genes differentially regulated in LLC and TG ventricles, only 7 presented

significant patterns of differential expression in both models. In both TG and LLC, 3

genes were downregulated, 4 genes were reciprocally regulated and none were

upregulated in common. It is important to note that statistical normalization of

microarray experiments and gene ranking leads to the positive selection of the most

robustly regulated genes in each model. Thus, genes that were found to be differentially

regulated only in one model could be differentially regulated in the other model too, but

less markedly.

4.3.1. Common downregulated genes

The mitochondrial cytochrome C oxidase 1 (mt-CO1) and the mitochondrial 16S rRNA

(mt-Rnr2) were found to be significantly downregulated in both mouse models. mt-

CO1 and mt-Rnr2 are genes expressed by the mitochondrial genome. Downregulation

of these genes could suggest that mitochondrial gene expression is impaired in both

LLC and TG hearts. Independently of the etiology and the severity of remodeling, this

could be regarded as a major common factor for the development of heart failure in

both mouse models. Interestingly, LLC hearts downregulate the gene coding for the

optic atrophy 1 homolog (human) (OPA1). This protein may be involved in

mitochondrial biogenesis. It is highly expressed in retina, but also expressed in brain,

testis, heart and skeletal muscle. Downregulation of this gene by siRNA leads to

fragmentation of the mitochondrial network, dissipation of the mitochondrial

membrane potential, cytochrome C release and caspase-dependent apoptosis (Olichon

et al., 2003). OPA1 may be involved in cytochrome C sequestration. This is in

agreement with the concept that developing mitochondrial damage could cause the

downregulation of the aerobic metabolic pathways, ultimately leading to energy

starvation and the development of heart failure.

Chapter V

249

The carboxyl esterase 3 (Ces3) was also found to be downregulated in both LLC and

TG hearts. Carboxyl esterases metabolize ester, thioester, carbamate and amide

compounds to more soluble acid, alcohol and amine products. The consequence of a

downregulation of Ces3 in the hearts of TG and LLC mice is unknown. Ces3 exhibits

relatively lower enzymatic activity than the other two isoforms Ces1 and Ces2

(Sanghani et al., 2004).

4.3.2. Reciprocally regulated genes

The homologue of the human chromosome condensation protein G (hCAPG, or

5730507H05Rik in the mouse), the CD2 antigen binding protein 2 (Cd2bp2) and the

ICOS ligand (Icosl) were upregulated in LLC but were downregulated in TG hearts.

hCAPG plays a central role in mitotic chromosome condensation and chromatine

stability (Geiman et al., 2004). Analysis of hCAPG mRNA expression showed highest

expression in the testis among normal tissues and variable expression in tumor cells,

reflecting the proliferative activity in these cells. This mitosis-related expression

suggests hCAPG as a possible proliferation marker and it could be associated with the

proliferative activity of fibroblasts in the LLC hearts. This would strengthen the idea

that a major difference between LLC and TG could be the balance between cell

proliferation and cell differentiation of non-myocyte cells. Some hypertrophic signaling

pathways present in the LLC hearts could activate non-myocyte proliferation in the

myocardium leading to collagen production and fibrosis, while other hypertrophic

signals specific to the TG mouse could preferentially activate non-myocyte cell

differentiation and prevent collagen production.

The ICOS ligand (Icosl) and the CD2 antigen binding protein 2 (Cd2bp2) were

upregulated in LLC and downregulated in TG hearts. ICOS ligand precursor (ICOSL),

regulates CD4 as well as CD8 T-cell responses via interaction with its receptor ICOS

on activated T cells. ICOSL expression is markedly increased in muscle fibres in

inflammatory myopathies and in myocarditis (Wiendl et al., 2003). Induction of ICOSL

is dependent on TNF-alpha and is regulated via NF-kappa B. Blockade of T cell

Chapter V

250

activation through ICOS suppresses expression of cytokines including INF-gamma,

IL4, IL6, IL-0, IL1beta, and TNF-alpha and inhibits T cell proliferation in vitro.

The CD2 antigen binding protein 2 (CD2BP2) is expressed in a wide variety of tissues

and organs. It interacts with the cytoplasmic tale of CD2, a surface antigen of the

human T-lymphocyte lineage that is expressed on all peripheral blood T cells. CD2BP2

expression selectively enhances IL2 production on cross-linking with CD2 (Nishizawa

et al., 1998). Thus, both ICOSL and CD2BP2 are markers of tissue penetration of

immune cell, inflammation and cytokine production. This evidence suggests that LLC

hearts suffer from fibrotic cardiomyopathy associated with cytokine-related

inflammation, while TG hearts suffer from inflammation-independent cardiomyocyte

and cardiac hypertrophy.

Finally, the fatty acid binding protein 3 (Fabp3) was found to be downregulated in LLC

hearts and upregulated in TG. Fatty acid metabolism in mammalian cells depends on

flux of fatty acids, between the plasma membrane and mitochondria or peroxisomes for

beta-oxidation, and between other cellular organelles for lipid synthesis. FABP proteins

play a role as transport vehicles of fatty acid compounds throughout the cytoplasm.

FABP3, also known as heart-type fatty acid binding protein (H-FABP) is abundant in

the heart and has low concentrations in the blood and in tissues outside the heart

(Alhadi and Fox, 2004). Overexpression of this gene in TG hearts could relate to

increased (or increased need for) fatty acid shuttling into the cytoplasm, possibly in

association with an enhancement of peroxisomal beta-oxidation. Interestingly, FABP3

has been reported as a sensitive and specific marker for the early diagnosis of acute

myocardial infarction, myocardial ischemia and development of congestive heart

failure (Goto et al., 2003; Chan et al., 2004; Tambara et al., 2004). FABP3 can be

secreted into the interstitial space by increased permeability of the cardiomyocyte

membrane associated with ischemic conditions, thus high serum of FABP3 can indicate

recent occurrence of an infarctive event within the previous 24 hours, while high

pericardial fluid levels can reflect chronic hypoxic conditions. To date there is no

evidence of a positive or negative correlation between mRNA expression levels and

serum or fluid levels for FABP3.

Chapter V

251

5. IN SUMMARY

The present study demonstrates differential expression of various set of genes in forms

of Ang II-induced cardiac hypertrophy and insulin resistant cardiomyopathy. Ang II-

stimulated cardiac hypertrophy is predominantly associated with differential expression

of pro-hypertrophic genes involved in protein biosynthesis and cell differentiation. In

contrast, cardiac remodeling induced by impaired myocardial glucose uptake is

associated with strong downregulation of genes involved in mitochondrial energy

production and upregulation of genes involved in cell proliferation and tissue

inflammation. Both forms of hypertrophy show evidence of impairment of

mitochondrial function, although mitochondrial dysfunction and damage seems to be

more important in LLC hearts. Collectively these data suggest that cardiac remodeling

in LLC and TG hearts is associated with the expression of separate sets of genes which

ultimately determine the morphological and mechanical differences observed between

the two hypertrophic models. The concept of a generic metabolic fetal recapitulation in

these two forms of normotensive hypertrophy is not supported.

These data are further considered in conjunction with other functional and morphologic

findings in the next Chapter.

Chapter V

252

M value Clone Symbol Name Heart*2.27 H3117D07 Hsd17b4 Hydroxysteroid (17-beta) dehydrogenase 4 ***2.19 H3114C04 EST no1.94 H3120F02 Tia1 Cytotoxic granule-associated RNA binding protein 1 **1.92 H3118A03 Hbb-y Hemoglobin Y, beta-like embryonic chain **1.87 H3118A04 Hbb-y Hemoglobin Y, beta-like embryonic chain **1.82 H3129G10 EST no1.80 H3118A05 Hbb-y Hemoglobin Y, beta-like embryonic chain **1.77 H3045A12 Hba-a1 Hemoglobin alpha, adult chain 1 ***1.74 H3097C06 Tgfb1i4 Transforming growth factor beta 1 induced transcript 4 ***1.72 H3054E05 AI385631 Expressed sequence AI385631 no1.72 H3144C04 1300010M03Rik RIKEN cDNA 1300010M03 gene ***1.61 H3136E01 E130307J07Rik RIKEN cDNA E130307J07 gene ***1.58 H3122A02 Hoxb3 Homeo box B3 **1.56 H3120G01 EST no1.53 H3065D08 Morc Microrchidia no1.52 H3117D02 Hbb-b2 Hemoglobin, beta adult minor chain ***1.51 H3128H08 LOC114601 Tangerin - provisional ***1.49 H3001B06 9330177P20Rik RIKEN cDNA 9330177P20 gene **1.48 H3032A08 Hspa5 Heat shock 70kD protein 5 (glucose-regulated protein) ***1.48 H3115C01 Car8 Carbonic anhydrase 8 **1.47 H3133A09 Sfrs4 Splicing factor, arginine/serine-rich 4 (SRp75) ***1.46 H3129G11 Dscr2 Down syndrome critical region homolog 2 (human) **1.45 H3139E01 Hspa8 Heat shock protein 8 ***1.43 H3144C11 Hspa5 Heat shock 70kD protein 5 (glucose-regulated protein) ***1.42 H3119D02 2810480G15Rik Riken cDNA 2810480G15 gene ***1.42 H3149E05 Agt Angiotensinogen ***1.39 H3118A06 Hbb-y Hemoglobin Y, beta-like embryonic chain **1.39 H3148G03 Dkk3 Dickkopf homolog 3 (Xenopus laevis) ***1.37 H3005A12 Taf11 TAF11 RNA polymerase II, TATA box binding protein (TBP)-associated factor **1.35 H3128A04 Arfgap1 ADP-ribosylation factor GTPase activating protein 1 **1.34 H3099G04 Slc28a3 Solute carrier family 28 (Na-coupled nucleoside transporter), member 3 **1.33 H3130A11 Fabp3 Fatty acid binding protein 3, muscle and heart ****1.29 H3126G11 Col15a1 Procollagen, type XV, alpha 1 ***1.29 H3072G05 Elavl2 ELAV (embryonic lethal, abnormal vision, Drosophila)-like 2 **1.28 H3127H08 EST no1.27 H3069D12 Ube4b Ubiquitination factor E4B, UFD2 homolog (S. cerevisiae) **1.26 H3036C07 Rpl10 Ribosomal protein 10 ***1.26 H3120C09 Rpl10 Ribosomal protein 10 ***1.26 H3125D02 EST no1.24 H3068D06 1110007A06Rik RIKEN cDNA 1110007A06 gene no1.21 H3022B03 Srp9 Signal recognition particle 9 ***1.20 H3018H04 Ptk9 Protein tyrosine kinase 9 **1.20 H3024A11 Rrm2 Ribonucleotide reductase M2 **1.20 H3017H10 Lasp1 LIM and SH3 protein 1 ***1.20 H3073F12 EST no1.19 H3017A12 Syngr1 Synaptogyrin 1 **1.19 H3001H10 Tmsb10 Thymosin, beta 10 ****1.19 H3007H07 EST no1.19 H3074B11 EST no1.19 H3100H03 3110001A13Rik RIKEN cDNA 3110001A13 gene *1.19 H3010F12 Smyd4 SET and MYND domain containing 4 ***1.18 H3021B10 Mta2 Metastasis-associated gene family, member 2 no1.18 H3013D11 Mt2 Metallothionein 2 ***1.18 H3028D05 Rpl5 Ribosomal protein L5 ***1.18 H3027G10 Zfp94 Zinc finger protein 94 **1.18 H3025A06 2400006P09Rik RIKEN cDNA 2400006P09 gene **1.18 H3014D04 Gtpbp1 GTP binding protein 1 **1.17 H3073C07 Phip Pleckstrin homology domain interacting protein **1.17 H3023H05 mt-Cytb cytochrome b, mitochondrial ?1.17 H3024F03 Cox7a2l Cytochrome c oxidase subunit VIIa polypeptide 2-like ***1.17 H3023C04 Itm1 Intergral membrane protein 1 **1.17 H3032F12 EST no1.17 H3021G11 Calr Calreticulin **1.16 H3022F12 Csh1 Chorionic somatomammotropin hormone 1 no1.16 H3028B06 D4Ertd478e DNA segment, Chr 4, ERATO Doi 478, expressed no1.15 H3140F07 Eef1a1 Eukaryotic translation elongation factor 1 alpha 1 ****1.15 H3031B10 Usf2 Upstream transcription factor 2 ***1.15 H3019G12 2510048O06Rik RIKEN cDNA 2510048O06 gene ***1.15 H3055F03 Gtl7 gene trap locus 7 **1.15 H3025B09 Pdcd11 Programmed cell death 11 *1.15 H3011H11 Rpl4 Ribosomal protein L4 ***1.14 H3032B09 Mesdc2 Mesoderm development candiate 2 **1.14 H3093H12 Mmp7 Matrix metalloproteinase 7 *1.14 H3062H08 1810030M08Rik RIKEN cDNA 1810030M08 gene ***1.14 H3010F11 Liph Lipase, member H **1.13 H3028C10 Farp1 FERM, RhoGEF (Arhgef) and pleckstrin domain protein 1 **1.13 H3079A02 Rpl12 Ribosomal protein L12 ***1.13 H3053H06 Slc25a13 Solute carrier family 25 (mitoch. carrier; adenine nucleotide translocator), member 13 **1.13 H3073C03 1500041J02Rik RIKEN cDNA 1500041J02 gene no1.13 H3021C04 Sui1-rs1 Suppressor of initiator codon mutations, related sequence 1 (S. cerevisiae) ****1.12 H3025F03 Car14 Carbonic anhydrase 14 **1.12 H3005H03 Rab18 RAB18, member RAS oncogene family **1.12 H3018F10 Gprc5c G protein-coupled receptor, family C, group 5, member C **1.12 H3025D10 Aga Aspartylglucosaminidase **1.11 H3021B04 Sypl Synaptophysin-like protein ***1.10 H3022D11 Ppp4r1 Protein phosphatase 4, regulatory subunit 1 **1.09 H3021F12 Pcna Proliferating cell nuclear antigen **1.09 H3022D12 Anapc7 Anaphase promoting complex subunit 7 **1.09 H3028E04 Ppp2r1a Protein phosphatase 2 (formerly 2A), regulatory subunit A (PR 65), alpha isoform ***1.09 H3005H10 Eef1d Eukaryotic translation elongation factor 1 delta (guanine nucleotide exchange protein) ***1.09 H3055C04 G3bp Ras-GTPase-activating protein SH3-domain binding protein -pending ***

Table V-1: Upregulated clones in TG ventricle

Clones are ranked according to their M value, in descending order. *The last column reports evidence for cardiac expression of the candidate gene. Evidence is based either on Weizmann Institute of Science DNA array experiments, performed with the Affymetrix HG-U95 set A-E, or from SAGE tags listed by the Cancer Genome Anatomy Project (*<10 copies/tag; **10-100 copies/tag; ***100-1000 copies/tag; ****>1000 copies/tag).

Chapter V

253

M value Clone Symbol Name Heart*-2.85 H3083C11 Ywhag 3-monooxgenase/tryptophan 5-monooxgenase activation protein, gamma polypeptide ***-2.45 H3080D12 Wbscr1 Williams-Beuren syndrome chromosome region 1 homolog (human) ***-2.44 H3036C12 Hdlbp High density lipoprotein (HDL) binding protein ***-2.36 H3135D05 Col9a3 Procollagen, type IX, alpha 3 **-2.15 H3020C06 Akt1 Thymoma viral proto-oncogene 1 ***-1.93 H3141D03 Ube3b Ubiquitin protein ligase E3B ***-1.91 H3136E12 Ucp2 Uncoupling protein 2, mitochondrial ***-1.84 H3012B04 Cs Citrate synthase ****-1.82 H3122A09 Wbscr18 Williams-Beuren syndrome chromosome region 18 homolog (human) **-1.82 H3064F01 2610507B11Rik RIKEN cDNA 2610507B11 gene ***-1.75 H3032G06 Cd97 CD97 antigen **-1.70 H3124B12 4732495E13Rik RIKEN cDNA 4732495E13 gene ***-1.69 H3091H09 Arnt Aryl hydrocarbon receptor nuclear translocator **-1.66 H3095C06 4933439F18Rik RIKEN cDNA 4933439F18 gene **-1.64 H3023C05 Nrbp Nuclear receptor binding protein ***-1.61 H3144H01 Immt Inner membrane protein, mitochondrial ***-1.57 H3015C12 Elf3 E74-like factor 3 **-1.56 H3151D03 Kif1b Kinesin family member 1B ***-1.55 H3004A05 Tbrg4 Transforming growth factor beta regulated gene 4 ***-1.53 H3092D12 C530043G21Rik RIKEN cDNA C530043G21 gene ***-1.52 H3022B01 mt-Co1 Cytochrome c oxidase I, mitochondrial ?-1.51 H3032G02 Ankrd17 Ankyrin repeat domain 17 **-1.50 H3113A09 Lamc1 Laminin, gamma 1 ***-1.50 H3158H01 Ppp5c Protein phosphatase 5, catalytic subunit **-1.49 H3118F06 Pctk1 PCTAIRE-motif protein kinase 1 ***-1.49 H3010A04 Ptbp1 Polypyrimidine tract binding protein 1 **-1.49 H3136B09 Bckdha Branched chain ketoacid dehydrogenase E1, alpha polypeptide ***-1.48 H3024B02 Sec61a1 Sec61 alpha 1 subunit (S. cerevisiae) ***-1.47 H3138A09 Eif4b Eukaryotic translation initiation factor 4B ****-1.46 H3044D09 Scd2 Stearoyl-Coenzyme A desaturase 2 *-1.45 H3026F05 Acads Acyl-Coenzyme A dehydrogenase, short chain ***-1.45 H3122B10 Ap1b1 Adaptor protein complex AP-1, beta 1 subunit ***-1.42 H3059G03 Trfr2 Transferrin receptor 2 *-1.41 H3078F08 Spata2 Spermatogenesis associated 2 *-1.40 H3076A12 Tbx20 T-box 20 no-1.39 H3066B09 Ankhd1 Ankyrin repeat and KH domain containing 1 **-1.36 H3089D02 Srpr Signal recognition particle receptor ('docking protein') ***-1.36 H3019B10 Mtap4 Microtubule-associated protein 4 ***-1.36 H3025D11 Pfkl Phosphofructokinase, liver, B-type ***-1.34 H3146C06 C1qr1 Complement component 1, q subcomponent, receptor 1 ***-1.33 H3146C12 Lrp1 Low density lipoprotein receptor-related protein 1 ***-1.33 H3025H05 Usp5 Ubiquitin specific protease 5 (isopeptidase T) ***-1.30 H3115A05 Plcd Phospholipase C, delta *-1.30 H3020H06 Mark2 MAP/microtubule affinity-regulating kinase 2 no-1.30 H3053E02 Poll Polymerase (DNA directed), lambda **-1.30 H3032C05 Sgta Small glutamine-rich tetratricopeptide repeat (TPR)-containing, alpha ***-1.29 H3089C02 D2Ertd391e DNA segment, Chr 2, ERATO Doi 391, expressed **-1.29 H3145H01 Nxf2 Nuclear RNA export factor 2 **-1.29 H3119B01 EST no-1.28 H3098G12 1700088E04Rik RIKEN cDNA 1700088E04 gene **-1.27 H3052H04 mt-Rnr2 16S rRNA, mitochondrial ?-1.26 H3080G10 Apc Adenomatosis polyposis coli **-1.26 H3032E06 Tulp4 Tubby like protein 4 *-1.25 H3150D05 Mtmr3 Myotubularin related protein 3 **-1.23 H3063B12 Fmo4 Flavin containing monooxygenase 4 **-1.22 H3080C09 D17Wsu92e DNA segment, Chr 17, Wayne State University 92, expressed no-1.22 H3054A08 C3 Complement component 3 no-1.22 H3122A04 4930438M06Rik RIKEN cDNA 4930438M06 gene ***-1.22 H3123A09 Emilin1 Elastin microfibril interfacer 1 **-1.22 H3052C01 D15Ertd417e DNA segment, Chr 15, ERATO Doi 417, expressed ***-1.22 H3018G12 Chst11 Carbohydrate sulfotransferase 11 **-1.22 H3117H11 Rxrb Retinoid X receptor beta **-1.19 H3104C11 AU045220 Expressed sequence AU045220 no-1.19 H3112E03 Mrps12 Mitochondrial ribosomal protein S12 ***-1.18 H3092G06 A630007B06Rik RIKEN cDNA A630007B06 gene **-1.18 H3025C08 Psmd13 Proteasome (prosome, macropain) 26S subunit, non-ATPase, 13 **-1.18 H3023F04 5430432P15Rik RIKEN cDNA 5430432P15 gene **-1.17 H3159C09 AW610627 Expressed sequence AW610627 no-1.17 H3017A02 Oaz2 Ornithine decarboxylase antizyme 2 ***-1.16 H3097F01 Lman2 Lectin, mannose-binding 2 ***-1.16 H3159B07 2900057K09Rik RIKEN cDNA 2900057K09 gene no-1.16 H3065F10 5730507H05Rik RIKEN cDNA 5730507H05 gene *-1.16 H3061C09 2810404F18Rik RIKEN cDNA 2810404F18 gene **-1.16 H3007G04 Psmd2 Proteasome (prosome, macropain) 26S subunit, non-ATPase, 2 ***-1.16 H3082B01 5730403B10Rik RIKEN cDNA 5730403B10 gene ***-1.15 H3103C07 Icosl Icos ligand **-1.14 H3135D01 Tde1 Tumor differentially expressed 1 ***-1.14 H3009F11 Rrbp1 Ribosome binding protein 1 **-1.13 H3088D09 Ece1 Endothelin converting enzyme 1 ***-1.13 H3111D10 Ces3 Carboxylesterase 3 no-1.13 H3106H04 Cd2bp2 CD2 antigen (cytoplasmic tail) binding protein 2 **-1.13 H3149B06 Fndc5 Fibronectin type III domain containing 5 ***-1.13 H3115G07 Cdkn3 cyclin-dependent kinase inhibitor 3 *-1.12 H3158C05 EST no-1.11 H3154G06 Ube2g2 Ubiquitin-conjugating enzyme E2G 2 **-1.11 H3063A02 Chd1l Chromodomain helicase DNA binding protein 1-like ***-1.10 H3086A11 AU015584 Expressed sequence AU015584 no-1.10 H3152B03 Grid2 Glutamate receptor, ionotropic, delta 2 no-1.08 H3159B06 EST no-1.06 H3025B01 Atp1a1 ATPase, Na+/K+ transporting, alpha 1 polypeptide ***

Table V-2: Downregulated clones in TG ventricle

Clones are ranked according to their M value, in descending order. *The last column reports evidence for cardiac expression of the candidate gene. Evidence is based either on Weizmann Institute of Science DNA array experiments, performed with the Affymetrix HG-U95 set A-E, or from SAGE tags listed by the Cancer Genome Anatomy Project (*<10 copies/tag; **10-100 copies/tag; ***100-1000 copies/tag; ****>1000 copies/tag).

Chapter V

254

M value Clone Symbol Name Heart*

2.93 H3078D03 Ndr4 N-myc downstream regulated gene 4 ***2.72 H3014C07 Nppb Natriuretic peptide precursor type B ***2.50 H3147B06 Gapd Glyceraldehyde-3-phosphate dehydrogenase ****2.38 H3097B03 Slc11a2 Solute carrier family 11 (proton-coupled divalent metal ion transporters), member 2 **2.36 H3103D04 Bub1b Budding uninhibited by benzimidazoles 1 homolog, beta (S. cerevisiae) **2.28 H3087D01 EST no2.21 H3159D08 2610014H22Rik RIKEN cDNA 2610014H22 gene no2.20 H3103B09 Cdh22 Cadherin 22 *2.17 H3102E11 EST no2.15 H3065F10 5730507H05Rik RIKEN cDNA 5730507H05 gene *2.02 H3045E04 Ybx2 Y box protein 2 **1.99 H3087A06 Falz Fetal Alzheimer antigen **1.96 H3087H09 S100a6 S100 calcium binding protein A6 (calcyclin) ***1.91 H3031E11 Gapd Glyceraldehyde-3-phosphate dehydrogenase ****1.88 H3050D06 Sh3bgr SH3-binding domain glutamic acid-rich protein **1.83 H3083H03 EST no1.82 H3098A01 EST no1.81 H3022B04 EST no1.80 H3092G11 Lmbr1 Limb region 1 **1.77 H3133E08 9630010G10Rik RIKEN cDNA 9630010G10 gene no1.76 H3102A09 Synj2bp Synaptojanin 2 binding protein **1.74 H3066G03 Gna14 Guanine nucleotide binding protein, alpha 14 **1.73 H3090C11 Rsb30 RAB30, member RAS oncogene family **1.68 H3106H04 Cd2bp2 CD2 antigen (cytoplasmic tail) binding protein 2 **1.66 H3093C06 6030413G23Rik RIKEN cDNA 6030413G23 gene no1.65 H3101C10 Gapd Glyceraldehyde-3-phosphate dehydrogenase ****1.64 H3010H10 Ramp2 Receptor (calcitonin) activity modifying protein 2 ***1.64 H3149G11 Mrvldc1 MARVEL (membrane-associating) domain containing 1 ***1.61 H3145A03 Prkar1a Protein kinase, cAMP dependent regulatory, type I, alpha ***1.60 H3083B05 Fusip1 FUS interacting protein (serine-arginine rich) 1 **1.60 H3093F01 1190002L16Rik RIKEN cDNA 1190002L16 gene ***1.59 H3089D10 AU016599 Expressed sequence AU016599 no1.58 H3097F06 Acsl5 Acyl-CoA synthetase long-chain family member 5 **1.57 H3062G07 EST no1.46 H3101A09 2810453I06Rik RIKEN cDNA 2810453I06 gene ***1.45 H3083D07 C87580 Expressed sequence C87580 no1.37 H3100B12 EST no1.36 H3097A02 Lap3 Leucine aminopeptidase 3 ***1.36 H3061H07 Greb1 Gene regulated by estrogen in breast cancer protein ***1.34 H3022B06 Triobp TRIO and F-actin binding protein ***1.33 H3103C07 Icosl Icos ligand **1.32 H3104H06 Nbr1 Neighbor of Brca1 gene 1 **1.31 H3025G07 Cdc2l2 Cell division cycle 2 homolog (S. pombe)-like 2 ***1.20 H3009D08 EST no1.19 H3102C03 C87499 Expressed sequence C87499 no1.15 H3061D06 Fath Fat tumor suppressor homolog (Drosophila) ***1.15 H3138E04 EST no1.14 H3020B08 Ftl1 Ferritin light chain 1 ****1.12 H3107F10 Ssr1 Signal sequence receptor, alpha **1.10 H3121E09 4930513H15Rik RIKEN cDNA 4930513H15 gene no

Table V-3: Upregulated clones in LLC ventricle

Clones are ranked according to their M value, in descending order. *The last column reports evidence for cardiac expression of the candidate gene. Evidence is based either on Weizmann Institute of Science DNA array experiments, performed with the Affymetrix HG-U95 set A-E, or from SAGE tags listed by the Cancer Genome Anatomy Project (*<10 copies/tag; **10-100 copies/tag; ***100-1000 copies/tag; ****>1000 copies/tag).

Chapter V

255

M value Clone Symbol Name Heart*

-3.11 H3131A07 Ech1 Enoyl coenzyme A hydratase 1, peroxisomal ****-2.86 H3109F07 Atp2a2 ATPase, Ca++ transporting, cardiac muscle, slow twitch 2 ****-2.75 H3120G06 Tpm1 Tropomyosin 1, alpha ****-2.18 H3112B07 Acaa2 Acetyl-Coenzyme A acyltransferase 2 (mitochondrial 3-oxoacyl-Coenzyme A thiolase) ***-2.14 H3137E02 Ndufs2 NADH dehydrogenase (ubiquinone) Fe-S protein 2 ***-2.04 H3140D12 Etfa Electron transferring flavoprotein, alpha polypeptide ***-2.00 H3024G04 mt-Co1 Cytochrome c oxidase I, mitochondrial ?-1.94 H3033C07 Fhl2 Four and a half LIM domains 2 ****-1.94 H3115C07 Opa1 Optic atrophy 1 homolog (human) **-1.93 H3120E09 BC020002 cDNA sequence BC020002 **-1.90 H3118E09 Tpm1 Tropomyosin 1, alpha ****-1.86 H3122F01 Atp5b ATP synthase, H+ transporting mitochondrial F1 complex, beta subunit ***-1.81 H3122E12 Atp5b ATP synthase, H+ transporting mitochondrial F1 complex, beta subunit ***-1.78 H3099H03 EST no-1.77 H3054E07 2610019A05Rik RIKEN cDNA 2610019A05 gene ***-1.75 H3138F10 Nsmaf Neutral sphingomyelinase (N-SMase) activation associated factor **-1.74 H3025C05 Mdh1 Malate dehydrogenase 1, NAD (soluble) ****-1.73 H3145E01 Slc25a3 Solute carrier family 25 (mitochondrial carrier; phosphate carrier), member 3 ****-1.72 H3144B03 Ndufb9 NADH dehydrogenase (ubiquinone) 1 beta subcomplex, 9 ****-1.72 H3011G11 EST no-1.69 H3066F05 mt-Rnr2 16S rRNA, mitochondrial ?-1.69 H3080H02 AF322649 cDNA sequence AF322649 **-1.67 H3003E04 Gcnt2 Glucosaminyl (N-acetyl) transferase 2, I-branching enzyme **-1.65 H3058C12 Stk11 Serine/threonine kinase 11 **-1.64 H3111D10 Ces3 Carboxylesterase 3 no-1.63 H3005E04 Atp5b ATP synthase, H+ transporting mitochondrial F1 complex, beta subunit ****-1.62 H3104D07 Fabp3 Fatty acid binding protein 3, muscle and heart ****-1.58 H3057E08 C87286 expressed sequence C87286 no-1.54 H3084G09 Tcl1b5 T-cell leukemia/lymphoma 1B, 5 *-1.51 H3059E05 Mdh1 Malate dehydrogenase 1, NAD (soluble) ****-1.46 H3115C02 Atp5j ATP synthase, H+ transporting, mitochondrial F0 complex, subunit F ***-1.45 H3066C07 Scye1 Small inducible cytokine subfamily E, member 1 **-1.43 H3110F07 EST no-1.37 H3061A08 Dyrk1a Dual-specificity tyrosine-(Y)-phosphorylation regulated kinase 1a ***-1.33 H3090C05 Gpiap1 GPI-anchored membrane protein 1 **-1.33 H3111F04 EST no-1.30 H3054E09 C88045 Expressed sequence C88045 no-1.27 H3115H10 Atp5c1 ATP synthase, H+ transporting, mitochondrial F1 complex, gamma polypeptide 1 ****-1.22 H3062C03 D8Ertd514e D8Ertd514e no-1.22 H3128D01 Mat1a Methionine adenosyltransferase I, alpha **-1.21 H3131C01 1700022L09Rik RIKEN cDNA 1700022L09 gene **-1.15 H3002D07 Rps6ka1 Ribosomal protein S6 kinase polypeptide 1 **-1.14 H3059G06 Mtf1 Metal response element binding transcription factor 1 *-1.13 H3062E02 Idh2 Isocitrate dehydrogenase 2 (NADP+), mitochondrial **-1.10 H3121D12 EST (similar to general transcription factor IIIC, polypeptide 3, 102kDa) no-1.10 H3110G03 Pink1 PTEN induced putative kinase 1 ***

Table V-4: Downregulated clones in LLC ventricle

Clones are ranked according to their M value, in descending order. *The last column reports evidence for cardiac expression of the candidate gene. Evidence is based either on Weizmann Institute of Science DNA array experiments, performed with the Affymetrix HG-U95 set A-E, or from SAGE tags listed by the Cancer Genome Anatomy Project (*<10 copies/tag; **10-100 copies/tag; ***100-1000 copies/tag; ****>1000 copies/tag).

Chapter V

256

0 10 20 30 40 50 60

cytoplasm (GO:0005737)

integral to membrane(GO:0016021)

nucleus (GO:0005634)

plasma membrane(GO:0005886)

ribonucleoprotein complex(GO:0030529)

inner membrane(GO:0019866)

mitochondrial membrane(GO:0005740)

membrane fraction(GO:0005624)

endomembrane system(GO:0012505)

chromosome (GO:0005694)

collagen (GO:0005581)

cyclin-dep. prot. kinaseholoenzyme (GO:0000307)

synapse (GO:0045202)

basement membrane(GO:0005604)

respiratory chain complexIV (GO:0045277)

Figure V-1: Gene Ontology classification of genes in TG hearts according to the cellular location

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘cellular component’ classification, at level 4.

Chapter V

257

0 10 20 30 40 50 60 70 80 90

binding (GO:0005488)

catalytic activity(GO:0003824)

transporter activity(GO:0005215)

signal transducer activity(GO:0004871)

structural molecule activity(GO:0005198)

transcription regulatoractivity (GO:0030528)

molecular_functionunknown (GO:0005554)

chaperone activity(GO:0003754)

translation regulatoractivity (GO:0045182)

enzyme regulator activity(GO:0030234)

motor activity(GO:0003774)

antioxidant activity(GO:0016209)

Figure V-2: Gene Ontology classification in TG hearts according to the molecular function

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘molecular function’ classification at level 2.

Chapter V

258

0 10 20 30 40 50 60 70 80

metabolism (GO:0008152)

cellular physiologicalprocess (GO:0050875)

cell communication(GO:0007154)

regulation of physiologicalprocess (GO:0050791)

response to stimulus(GO:0050896)

organismal physiologicalprocess (GO:0050874)

death (GO:0016265)

regulation of cellularprocess (GO:0050794)

morphogenesis(GO:0009653)

homeostasis (GO:0042592)

pattern specification(GO:0007389)

reproduction (GO:0000003)

pathogenesis (GO:0009405)

embryonic development(GO:0009790)

extracell. structure org. &bio. (GO:0043062)

Figure V-3: Gene Ontology classification in TG hearts according to the biological process

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘biological process’ classification at level 3.

Chapter V

259

0 5 10 15 20 25 30

cytoplasm (GO:0005737)

integral to membrane(GO:0016021)

nucleus (GO:0005634)

mitochondrial membrane(GO:0005740)

inner membrane(GO:0019866)

plasma membrane(GO:0005886)

membrane fraction(GO:0005624)

chromosome (GO:0005694)

endosome membrane(GO:0010008)

endomembrane system(GO:0012505)

outer membrane(GO:0019867)

extrinsic to membrane(GO:0019898)

ribonucleoprotein complex(GO:0030529)

proton-transp. ATP syntha.complex (GO:0045259)

respiratory chain complexIV (GO:0045277)

Figure V-4: Gene Ontology classification of genes in LLC mice according to the cellular location

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘cellular component’ classification, at level 4.

Chapter V

260

0 5 10 15 20 25 30 35 40

binding (GO:0005488)

catalytic activity(GO:0003824)

transporter activity(GO:0005215)

signal transducer activity(GO:0004871)

molecular_functionunknown (GO:0005554)

structural molecule activity(GO:0005198)

enzyme regulator activity(GO:0030234)

translation regulatoractivity (GO:0045182)

motor activity(GO:0003774)

transcription regulatoractivity (GO:0030528)

Figure V-5: Gene Ontology classification in LLC hearts according to the molecular function

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘molecular function’ classification at level 2.

Chapter V

261

0 5 10 15 20 25 30 35

metabolism (GO:0008152)

cellular physiologicalprocess (GO:0050875)

cell communication(GO:0007154)

regulation of cellularprocess (GO:0050794)

organismal physiologicalprocess (GO:0050874)

response to stimulus(GO:0050896)

morphogenesis(GO:0009653)

regulation of physiologicalprocess (GO:0050791)

mesoderm development(GO:0007498)

embryonic development(GO:0009790)

homeostasis (GO:0042592)

Figure V-6: Gene Ontology classification in LLC hearts according to the biological process

Differential distribution of GO terms for the upregulated (green) and downregulated (red) genes according to the ‘biological process’ classification at level 3.

Chapter V

262

Carbonic anhydrase

NH

E1

AE

3

CO2 + H2O

H+ HCO3-

Cl-Na+

Carbonic anhydrase

NH

E1

AE

3

CO2 + H2O

H+ HCO3-

Cl-Na+

Figure V-7: Depiction of the physical and biochemical interactions between CAR - NHE-1 - AE3

TG hearts differentially upregulate various isoforms of the carbonic anhydrases (CAR). Together with the upregulation of the NHE-1, it could be speculated that increased acid load could be compensated by an exchange with sodium chloride, leading to sodium overload and cardiomyocyte hypertrophy. CAR= carbonic anhydrase; NHE-1= sodium/hydrogen exchanger 1; AE3= anion exchanger 3.

CHAPTER VI

Mechanisms of cardiac remodeling in

Ang II-induced and insulin resistant cardiac

hypertrophy: a comparative overview

Chapter VI

264

1. MORPHOLOGIC CHANGES IN CARDIAC

HYPERTROPHY

As it is demonstrated for both TG1306/1R and GLUT4-KO mouse models, that cardiac

hypertrophy is associated with an increase in ventricular mass and chamber volume

(Chapter III, Tables III-1 and III-2). Common features observed in the development of

cardiac hypertrophy in both models are the stimulation of the intra-cardiac renin-

angiotensin system, with overexpression of Agt mRNA, and decreased GLUT4 protein

levels (Chapter III, Figures III-3 & III-4 and III-8 & III-9). A comparative analysis of

WT, TG, LL and LLC hearts shows that the severity of morphologic remodeling, that is

myocardial collagen deposition, fibroblast proliferation and cardiomyocyte

hypertrophy, is determined by the extent and cumulative effect of neuro-humoral

activation and metabolic imbalance.

When compared to WT mice (that is non-transgenic littermate mice from the

TG1306/1R strain), LL mice present evidence of modest cardiac and cardiomyocyte

hypertrophy (Figure VI-1). As described by Kaczmarczyk et al. (2003), LL mice are

exposed to systemic metabolic disturbances (insulin-signaling desensitization, impaired

glucose disposal) and downregulation of cardiac GLUT4 levels to 10-15 % of WT

(Chapter III, Figure III-10). These systemic and tissue-specific modifications could

predispose LL hearts to a degree of cardiac and cardiomyocyte remodeling, the

magnitude of which is similar to that observed in Agt-overexpressing transgenic mice:

that is, 15-20% increase in CWI and cardiomyocyte dimensions without collagen

deposition, when compared to non-transgenic littermates (i.e. WT) (Chapter III,

comparative analysis of Tables III-1 and III-2). Strain-specific differences due to

variability in genetic background should also be considered, since WT and TG mice are

estimated to be ~99% C57BL6-derived, while LL and LLC mice carry a more

heterogeneous genetic background composed of C57BL6 (~60%), 129Sv (~35%) and

CBA (~5%).

Chapter VI

265

While cardiac hypertrophy in TG mice is induced by the stimulation of the intracardiac

renin-angiotensin system, a downregulation of cardiac GLUT4 levels is also observed,

the extent of which correlates with the degree of remodeling (Chapter III, Figure III-4).

These data suggest that a decrease in cardiac GLUT4 protein level is associated with

the development of hypertrophy, independently of the upregulation of the cardiac RAS

or insulin resistance.

The data presented in Chapter III also demonstrate that without additional provocation,

systemic insulin resistance (as observed in LL mice) and cardiac overexpression of Agt

(induced in TG mice) are insufficient to stimulate fibroblast proliferation and collagen

deposition in the myocardium. The insulin-resistant LLC mice exhibit a very different

degree of remodeling, which associates cardiomyocyte hypertrophy with fibrosis

(Table III-2 and Figure III-7), activation of the intra-cardiac RAS (Figure III-8) and

evidence of reduction in glucose uptake in the heart (Kaczmarczyk et al., 2003). Thus it

could be proposed that deranged intracardiac glucose metabolism (due to the severity

of GLUT4 deficiency), in combination with other events, such as hypoxia, increased

production of reactive oxygen species, inflammation and tissue damage, are necessary

to cause dramatic restructuring of the extracellular matrix of the myocardium and to

stimulate non-myocyte proliferation and collagen accumulation.

Chapter VI

266

2. MECHANICAL CHANGES IN CARDIAC

HYPERTROPHY

The morphological analysis presented in Chapter III shows that the severity of tissue

remodeling is determined by the extent and the cumulative effect of neuro-humoral

activation and metabolic imbalance. Chapter IV reports that cardiomyocyte

hypertrophy is associated with impaired contractile function of isolated myocytes and

differential expression of key transporters involved in calcium and pH homeostasis.

In Figures VI-2 and VI-3 a comparative evaluation of the cardiomyocyte contractile

function in WT and LL is presented. The two ‘control’ groups exhibit equivalent

duration of excitation-contraction coupling latency (To) and duration of the shortening

and lengthening periods (Tm-To; Tf -Tm), summing to produce a comparable whole

contraction period (Tf -To) (Figure VI-2). As NHE-1 and SERCA2 expression levels in

WT and LL hearts are similar (Figure VI-4), the stability observed in twitch duration

may suggest that pH regulation of myofilament sensitivity and calcium removal

mechanisms are preserved in these two mouse lines. This hypothesis is strengthened by

additional information showing that the age-dependent shift in the same contraction

parameters is comparable for WT and LL cardiomyocytes and it is consistent with

NHE-1 downregulation in both types of myocytes (Figure VI-4). Decrease in proton

export, possibly secondary to decreased rates of glycolysis and reduced production of

acid charges (H+), would reduce myofilament calcium sensitivity and shortening

duration, causing age-dependent dysregulation of intracellular pH homeostasis in both

WT and LL hearts. An explanation for the age-dependent alteration for the lengthening

period (Tf -Tm) in WT and LL myocytes is not immediately apparent, but it could be

secondary to depressed NHE activity, leading to sodium and calcium depletion,

ultimately reducing the duration of the whole contraction period.

In contrast to what it is seen in WT cardiomyocytes, ageing has no effect on LL

cardiomyocyte inotropy (%S and MRS) and lusitropy (MRL) (Figure VI-3). A similar

Chapter VI

267

pattern is also observed for LLC cardiomyocytes (Chapter IV, Table IV-4 and Figure

IV-9), suggesting that ageing does not impair cardiomyocyte inotropy, and surprisingly

improves relaxation dynamics in LL and LLC mice. As explained in Chapter IV, an

increase in rate of lengthening in LL and LLC myocytes could represent an important

compensatory mechanism to shorten the contractile cycle and preserve peak shortening

(%S). Further experiments are needed to determine whether these disparities in

contractile and relaxation performances are determined by the genetic background or

whether they are related to systemic insulin-signaling desensitization and impaired

glucose disposal in LL and LLC mice.

In Figures VI-5 and VI-6 a comparative evaluation of the contractile function in TG

and LLC cardiomyocyte is shown. Not surprisingly, cardiac hypertrophy in TG and

LLC mice is associated with different age- and genotype-dependent alterations in

cardiomyocyte contractile function. Generally lower rates of shortening and

lengthening (MRS and MRL) and longer contractile cycle timing periods are observed

for the older TG cardiomyocytes when compared to age-matched LLC. This could be

caused by a slower calcium re-uptake into the SR by the SR calcium ATPase SERCA2,

which normally accounts for the removal of ~90% of the intracellular calcium

concentrations in the rodent. Indeed, a characteristic of the TG heart is a more

pronounced reduction in myocardial SERCA2 protein expression already at 15-20

weeks of age when compared to LLC (28.3±1.6 vs. 65.4±0.4 in arbitrary units, p<0.05).

Downregulation of the SERCA2 calcium pump would suppress the re-uptake of

calcium into the SR and delay myocyte relaxation. The concomitant upregulation of the

NCX in TG hearts (at 15-20 weeks: 37.4±4.6 vs. 5.4±0.4 in arbitrary units; at 35-40

weeks: 28.0±2.5 vs. 4.9±0.4 in arbitrary units, when compared to LLC; p<0.05, 2-way

ANOVA), could enhance export of calcium ions to the extracellular medium and shift

the calcium cycling from an intracellular pathway toward an extracellular one. This

would reduce the levels of releasable SR calcium in TG mice and chronically erode

systolic functional reserve with ageing.

Differential expression of contractile protein isoforms and metabolic genes could also

account for the differences observed in contractile performance between TG and LLC

mice and further investigation of these possibilities is merited.

Chapter VI

268

3. DIFFERENTIAL GENE EXPRESSION OF METABOLIC

SUBSTRATES IN CARDIAC HYPERTROPHY

The findings reported in Chapter V demonstrate cDNA microarray assays to be a useful

tool in describing the differential expression of the whole transcriptome in cardiac

tissue of TG1306/1R and GLUT4 knock-out mice. The large amount of information

reported in Chapter V makes the interpretation of the data a very complex task to

perform. Nevertheless the results suggest that metabolic regulatory responses are

crucial in the development of the cardiac phenotypes in TG and LLC mice. These

metabolic responses include differential activation of signal transduction events and the

regulation of genes encoding rate-limiting enzymes and proteins. Indeed, the results

presented in Chapter V show that many of the metabolic regulatory events that dictate

fuel selection and capacity for ATP production in the failing heart occur at the level of

gene expression.

Metabolic regulation is inevitably linked with cardiac and cardiomyocyte function. This

metabolism-function relationship is relevant to diseases that lead to cardiac

hypertrophy and heart failure. The progression to heart failure of any cause is

associated with a gradual but progressive decline in the activity of mitochondrial

respiratory pathways leading to diminished capacity for ATP production (reviewed in

Russell et al., 2005). The functional results presented in Chapter IV, combined with the

gene expression data reported in Chapter V suggest that energy deficiency can be a

cause and an effect of cardiomyocyte functional deterioration in the contrasting

hypertrophic etiologies exhibited by TG and LLC mice.

As mentioned in the introductory Chapter I, the oxidation of free fatty acids and

glucose in mitochondria accounts for the vast majority of ATP production in the

healthy adult heart. Studies using animal models of ventricular hypertrophy induced by

pressure overload have consistently demonstrated a myocardial shift away from fatty

acid oxidation toward glycolytic metabolism. Changes in gene expression in pressure-

Chapter VI

269

load hypertrophied and failing hearts, showing downregulation of mitochondrial

enzymes involved in fatty acid oxidation are consistent with the observed metabolic

alterations. Data presented in Chapter V for the TG mouse model partially conflict with

this ‘generic’ hypertrophy concept, since the decrease in mitochondrial fatty acid

metabolism is accompanied by a decrease in enzymes involved in glucose metabolism

(e.g. phosphofructokinase), combined with an apparent compensatory activation of

peroxisomal free fatty acid beta oxidation (Chapter V). This ‘double downregulation’

of glucose and fatty acid metabolism could account for the slower rates of shortening

and lengthening observed in TG cardiomyocytes (Figure VI-6) and the transition to

heart failure.

On the contrary, The data derived from the GLUT4-KO model supports the concept of

a myocardial shift from fatty acid oxidation toward glucose oxidation during the

development of this cardiomyopathy (Chapter V). Furthermore, the data from LLC

would suggest a shift from aerobic production of ATP, involving mitochondrial

oxidation of NADH equivalents, to an anaerobic pathway, where glucose and glycogen

are metabolized to pyruvate and then reduced to lactate by the lactate dehydrogenase.

Interestingly, exogenous pyruvate supplementation “by-passes” dysfunctional GLUT4

dependent glucose transport, apparently facilitates ATP production, and selectively

enhances cardiac performance ex vivo in LLC mice (Huggins et al., 2004), suggesting

that the machinery for glucose-dependent ATP production is not disrupted but just

‘silenced’, due to diminished glucose uptake in LLC hearts. This major difference in

fuel selection and energetics could account for the significant differences observed in

contractile parameters between TG and LLC cardiomyocytes when the contractile

behavior is examined under similar ex vivo conditions (Figure VI-5 and VI-6).

It is important to note that previous studies have reported a primary increase in the

cardiac fatty acid metabolism during cardiac remodeling induced by insulin resistance

and hyperglycemia (Chapter I, Section 8.3.4). This fuel switch is believed to be linked

to the combined effects of myocyte insulin- and pH-dependent decrease in GLUT4

membrane translocation, leading to diminished glucose uptake, with systemic

dyslipidemia and increased cardiac PPARα-dependent fatty acid oxidation rates. One

proposed mechanism for cardiac and cardiomyocyte dysfunction in these diabetic

Chapter VI

270

conditions is excess intracellular lipid accumulation resulting in myocyte dysfunction

or death, termed ‘lipotoxicity’. In LLC hearts the opposite occurs: upregulation of

enzymes involved in glycolysis and gluconeogenesis is associated with downregulation

of enzymes involved in the beta-oxidation and disruption of the mitochondrial electron

chain. It could be speculated that in the insulin resistant LLC heart early compensatory

mechanisms could lead to increased fatty acid oxidation and diminished glucose

utilization, leading to myocyte death by lipotoxicity and scar formation (Chapter III,

Figure III-7). Despite increased rates of fatty acid oxidation, chronic downstream

defects in mitochondrial oxidative phosphorylation could progressively develop with

age, leading to mitochondrial oxidative damage, myocyte death and eventual

impairment of aerobic respiration. Moreover, an ‘uncoupling’ between rates of fatty

acid oxidation and oxidative phosphorylation could lead to toxic intermediate

accumulation and perpetuate a continued cycle of mitochondrial damage which would

eventually undermine any mitochondrion-related metabolic process. Further

investigation of these possibilities is required.

Thus, chronic alterations in myocardial and cardiomyocyte fuel selection and energetics

are linked to the development of cardiomyocyte dysfunction and progression of heart

failure. The direction of the ‘fuel shift’ varies according to the etiology and possibly

according to the time course and the combination of substrate and functional insults.

Chapter VI

271

4. IN CONCLUSION

This Thesis demonstrates that overexpression of the cardiac renin-angiotensin system

(in TG mice) or decreased cardiac glucose uptake and insulin-resistance (in LLC mice)

leads to tissue and cellular remodeling, producing cardiac hypertrophy and failure. In

these models, important differences in the cardiac adaptive response to chronic neuro-

humoral or metabolic imbalance emerge:

1. Compared to TG hearts, myocardial remodeling in the LLC involves more

pronounced cardiomyocyte hypertrophy, with extracellular matrix accumulation

and development of fibrosis and fibrotic scars (Chapter III).

2. Although morphological remodeling is exacerbated in LLC hearts,

cardiomyocytes from TG hearts show longer contraction cycle timing periods

when compared to age-matched LLC myocytes (Chapter IV).

3. In addition, inotropic and lusitropic parameters are reduced in the older TG

cardiomyocytes when compared to age-matched LLC (Chapter IV).

4. These differences in myocyte contractile performance between TG and LLC

mice are associated with differential gene (and protein) expression of key

players governing pH, calcium and metabolic homeostasis in the heart

(Chapters III to V).

There is a critical threshold of cardiac GLUT4 protein expression below which

exacerbated cardiomyocyte hypertrophy with fibroblast proliferation and collagen

production is triggered. Metabolically speaking, there is a threshold under which

GLUT4 deficiency impairs cardiac insulin-stimulated glucose uptake and compromises

cardiomyocyte function. Under these conditions the intra-cardiac renin-angiotensin

system is also triggered, with elevated Agt mRNA overexpression in LLC hearts. It

Chapter VI

272

demonstrates the important role played by the glucose transporter GLUT4 in regulating

cardiomyocyte metabolic balance, gene expression and contractile function. As low as

5-15% of GLUT4 protein expression is sufficient to preserve cardiomyocytes from

exacerbated hypertrophy and dysfunction.

Ang II, a peptide known for its pro-hypertrophic and pro-fibrotic properties, is an

important trigger of cardiomyocyte hypertrophy, as demonstrated in TG mice, but it is

not sufficient to induce fibroblast proliferation and collagen accumulation in vivo

conditions. Other coincident processes, such as inhibition of extracellular matrix

degradation, pronounced glucose metabolic imbalance, development of inflammation

and production of cytokines and other peptides are probably required to jointly

stimulate Ang II-mediated collagen production and fibrosis in the heart.

Thus, overexpression of the cardiac renin-angiotensin system and decreased cardiac

glucose intake lead to cardiac and cardiomyocyte remodeling, producing cardiac

hypertrophy and failure in mice. The severity of cardiomyocyte and tissue remodeling

is determined by the extent of cardiomyocyte hypertrophy and myocardial collagen

deposition. The present study support the concept that cardiac-specific activation of the

renin-angiotensin system and decrease in cardiac glucose uptake are sufficient

‘triggers’ to activate pathologic cardiac remodeling, the severity of which is dependent

on the cumulative activation of the intra-cardiac renin-angiotensin system and the

extent of glucose metabolic perturbation.

The maladaptive responses to cardiac-specific angiotensinogen overexpression and to

decreased cardiac glucose uptake and insulin-resistance are different, suggesting that

not all hypertrophic responses are the same. The concept that cardiac hypertrophy

represents a generic growth ‘recapitulation’ is not supported by these findings.

Regardless of the phenotype observed, this study demonstrates that overexpression of

cardiac angiotensinogen mRNA producing Ang II, and low expression of GLUT4

protein are independent but cumulative triggers of cardiac hypertrophy, leading to

mechanical dysfunction and failure

Chapter VI

273

0

2

4

6

8

10

WT TG LL LLC

0

2

4

6

8

WT TG LL LLC

0

60

120

180

240

WT TG LL LLC

A. Cardiac weight index (mg/g)

B. LV collagen (%tsa)

C. Myocyte length (μm)

0

2

4

6

8

10

WT TG LL LLC

0

2

4

6

8

WT TG LL LLC

0

60

120

180

240

WT TG LL LLC

A. Cardiac weight index (mg/g)

B. LV collagen (%tsa)

C. Myocyte length (μm)

Figure VI-1: Comparative analysis of cardiac remodeling in WT, TG, LL and LLC mice

Comparison of the cardiac weight index (A.), the myocardial collagen content (B.) and the cardiomyocyte length (C.) between the different mouse strains. For each group, data from animals aged 15-20 and 35-40 weeks were pooled. The severity of tissue remodeling, that is cardiac and cardiomyocyte hypertrophy and myocardial collagen deposition, is maximal in the LLC mouse where neuro-humoral activation and metabolic imbalance produce a cumulative effect.

Chapter VI

274

Figure VI-2: Pacing responses of isotonically shortening myocytes from WT and LL (1.5-5.0 Hz)

Cycle time parameters. (To= the excitation-contraction coupling latency; Tf -To= duration of the whole contractile cycle; Tm-To= duration of shortening; Tf -Tm,= duration of lengthening). † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

8

9

10

11

12

13

14

Hz 1.5 Hz 3 Hz 4 Hz 550

70

90

110

130

Hz 1.5 Hz 3 Hz 4 Hz 5

15

20

25

30

35

40

45

Hz 1.5 Hz 3 Hz 4 Hz 530

40

50

60

70

80

90

Hz 1.5 Hz 3 Hz 4 Hz 5

1.5 3 4 5

ms

ms

(Hz)(Hz)

1.5 3 4 5 (Hz)

1.5 3 4 5

(Hz)1.5 3 4 5

ms

ms

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

A. To B. Tf-To

C. Tm-To D. Tf-Tm

To Tf -To

Tf -TmTm-To

8

9

10

11

12

13

14

Hz 1.5 Hz 3 Hz 4 Hz 550

70

90

110

130

Hz 1.5 Hz 3 Hz 4 Hz 5

15

20

25

30

35

40

45

Hz 1.5 Hz 3 Hz 4 Hz 530

40

50

60

70

80

90

Hz 1.5 Hz 3 Hz 4 Hz 5

1.5 3 4 5

ms

ms

(Hz)(Hz)

1.5 3 4 5 (Hz)

1.5 3 4 5

(Hz)1.5 3 4 5

ms

ms

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

A. To B. Tf-To

C. Tm-To D. Tf-Tm

To Tf -To

Tf -TmTm-To

Chapter VI

275

4

6

8

10

12

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

7

8

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

7

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5

Lo/

sL

o/s

%L

o

(Hz)

(Hz)

(Hz)

1.5 3 4 5

1.5 3 4 5

*‡

‡*

*‡ ¶

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

35-40, WT35-40, LL15-20, WT15-20, LL

A. Normalized %S

B. Normalized MRS

C. Normalized MRL

Figure VI-3: Pacing responses of isotonically shortening myocytes from WT and LL (1.5-5.0 Hz)

Inotropic and lusitropic parameters (%S= maximum cell shortening; MRS= maximal rate of cell shortening; MRL= maximal rate of cell lengthening). * p<0.05, LL vs. WT cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

Chapter VI

276

0

40

80

120

160

CTRL LL

01

23

45

67

15-20 35-40Cell

A.Ex

pres

sion

leve

l(A

rb. U

nits

)

15-20 week 35-40 week

NHE-1

LLWT

B. SERCA2

Expr

essi

on le

vel

(Arb

. Uni

ts)

0

40

80

120

160

CTRL LL

01

23

45

67

15-20 35-40Cell

A.Ex

pres

sion

leve

l(A

rb. U

nits

)

15-20 week 35-40 week

NHE-1

LLWT

B. SERCA2

Expr

essi

on le

vel

(Arb

. Uni

ts)

Figure VI-4: mRNA and protein expression profiles in WT and LL hearts

Expression profiles of sodium-calcium exchanger NHE-1 mRNA (A.) and sarcoplasmic reticulum calcium pump SERCA2 protein levels at 15-20 weeks of age (B.). † p<0.05, 35-40 vs. 15-20 week old hearts (2-way ANOVA). Comparative analysis across the two sets of data is validated by equal conditions of PCR amplification, electrophoresis and Western blotting for GLUT4-KO and TG1306/1R mice.

Chapter VI

277

Figure VI-5: Pacing responses of isotonically shortening myocytes from TG and LLC (1.5-5.0 Hz)

Cycle time parameters. (To= the excitation-contraction coupling latency; Tf -To= duration of the whole contractile cycle; Tm-To= duration of shortening; Tf -Tm,= duration of lengthening). * p<0.05, TG vs. LLC cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

20

30

40

50

Hz 1.5 Hz 3 Hz 4 Hz 5 30

50

70

90

110

Hz 1.5 Hz 3 Hz 4 Hz 5

50

75

100

125

150

Hz 1.5 Hz 3 Hz 4 Hz 58

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5

ms

ms

(Hz)(Hz)

1.5 3 4 5 (Hz)

1.5 3 4 5

(Hz)1.5 3 4 5

ms

ms

*

*

*

35-40, TG35-40, LLC15-20, TG15-20, LLC

A. To B. Tf-To

C. Tm-To D. Tf-Tm

35-40, TG35-40, LLC15-20, TG15-20, LLC

To Tf -To

Tf -TmTm-To

20

30

40

50

Hz 1.5 Hz 3 Hz 4 Hz 5 30

50

70

90

110

Hz 1.5 Hz 3 Hz 4 Hz 5

50

75

100

125

150

Hz 1.5 Hz 3 Hz 4 Hz 58

12

16

20

24

28

Hz 1.5 Hz 3 Hz 4 Hz 51.5 3 4 5

ms

ms

(Hz)(Hz)

1.5 3 4 5 (Hz)

1.5 3 4 5

(Hz)1.5 3 4 5

ms

ms

*

*

*

35-40, TG35-40, LLC15-20, TG15-20, LLC

A. To B. Tf-To

C. Tm-To D. Tf-Tm

35-40, TG35-40, LLC15-20, TG15-20, LLC

To Tf -To

Tf -TmTm-To

Chapter VI

278

2

3

4

5

6

7

Hz 1.5 Hz 3 Hz 4 Hz 5

2

3

4

5

6

Hz 1.5 Hz 3 Hz 4 Hz 5

3

5

7

9

11

Hz 1.5 Hz 3 Hz 4 Hz 5

1.5 3 4 5

Lo/

sL

o/s

%L

o

(Hz)

(Hz)

(Hz)

1.5 3 4 5

1.5 3 4 5

*‡

‡¶

*‡ ¶

35-40, TG35-40, LLC15-20, TG15-20, LLC

A. Normalized %S

B. Normalized MRS

C. Normalized MRL

35-40, TG35-40, LLC15-20, TG15-20, LLC

35-40, TG35-40, LLC15-20, TG15-20, LLC

Figure VI-6: Pacing responses of isotonically shortening myocytes from TG and LLC (1.5-5.0 Hz)

Inotropic and lusitropic parameters (%S= maximum cell shortening; MRS= maximal rate of cell shortening; MRL= maximal rate of cell lengthening). * p<0.05, TG vs. LLC cardiomyocytes; † p<0.05, 35-40 vs. 15-20 week old cardiomyocytes; ‡ p<0.05, pacing frequency effect; ¶ p<0.05, age-genotype interaction effect (all values derived from analysis by multiple-way ANOVA for repeated measures).

Bibliography

Bibliography

280

AbdAlla S, Lother H and Quitterer U (2000). AT1-receptor heterodimers show enhanced G-protein activation and altered receptor sequestration. Nature, 407: 94-98. AbdAlla S, Lother H, Abdel-tawab AM and Quitterer U (2001). The angiotensin II AT2 receptor is an AT1 receptor antagonist. J Biol Chem, 276: 39721-39726. Abel ED (2004). Glucose transport in the heart. Front Biosci, 9: 201-215. Abel ED, Kaulbach HC, Tian R, Hopkins JC, Duffy J, Doetschman T, Minnemann T, Boers ME, Hadro E, Oberste-Berghaus C, Quist W, Lowell BB, Ingwall JS, Kahn BB (1999). Cardiac hypertrophy with preserved contractile function after selective deletion of GLUT4 from the heart. J Clin Invest, 104: 1703-1714. Abel ED, Peroni O, Kim JK, Kim YB, Boss O, Hadro E, Minnemann T, Shulman GI and Kahn BB (2001). Adipose-selective targeting of the GLUT4 gene impairs insulin action in muscle and liver. Nature, 409: 729-733. ACC/AHA (2001). Guidelines for the evaluation and management of chronic heart failure in the adult: executive summary. Circulation, 104: 2996-3007. Adams Jr KF (2001). New epidemiologic perspectives concerning mild-to-moderate heart failure. Am J Med, 110 (Suppl 7A): 6S-13S. Adams RJ, Cohen DW, Gupte S, Johnson JD, Wallick ET, Wang T and Schwartz A (1979). In vitro effects of palmitylcarnitine on cardiac plasma membrane Na+, K+-ATPase and sarcoplasmic reticulum Ca2+-ATPase and Ca2+ transport. J Biol Chem, 254: 12404-12410. Ai T, Horie M, Obayashi K and Sasayama S (1998). Accentuated antagonism by angiotensin II on guinea-pig cardiac L-type Ca-currents enhanced by beta-adrenergic stimulation. Pflugers Arch, 436: 168-174. AIRE (Acute Infarction Ramipril Efficacy) Study Investigators (1993). Effect of ramipril on mortality and morbidity of survivors of acute myocardial infarction with clinical evidence of heart failure. Lancet, 342: 821-828. Aiello EA and Cingolani HE (2001). Angiotensin II stimulates cardiac L-type Ca(2+) current by a Ca(2+)- and protein kinase C-dependent mechanism. Am J Physiol, 280: H1528-H1536. Aiello EA, Villa-Abrille MC and Cingolani HE (2002). Autocrine stimulation of cardiac Na(+)-Ca(2+) exchanger currents by endogenous endothelin released by angiotensin II. Circ Res, 90: 374-376. Aikawa R, Nawano M, Gu Y, Katagiri H, Asano T, Zhu W, Nagai R and Komuro I (2000). Insulin prevents cardiomyocytes from oxidative stress-induced apoptosis through activation of PI3 kinase/Akt. Circulation, 102: 2873-2879.

Bibliography

281

Allard MF, Schonekess BO, Henning SL, English DR and Lopaschuck GD (1994). Contribution of oxidative metabolism and glycolysis to ATP production in hypertrophied hearts. Am J Physiol, 267: H742-H750. Allen AM, Zhuo J and Mendelsohn FAO (2000). Localization and function of angiotensin AT1 receptors. Am J Hypertens, 13: 31S-38S. Allen DG and Xiao XH (2003). Role of the cardiac Na+/H+ exchanger during ischemia and reperfusion. Cardiovasc Res, 57: 934-941. Allen ND, Norris ML and Surani MA (1990). Epigenetic control of transgene expression and imprinting by genotype-specific modifiers. Cell, 61: 853-861. Alpert NR and Mulieri LA (1982). Increased myothermal economy of isometric force generation in compensated cardiac hypertrophy induced by pulmonary artery constriction in rabbit: a characterization of heat liberation in normal and hypertrophied right ventricular papillary muscle. Circ Res, 50: 491-500. Alvarez BV, Loiselle FB, Supuran CT, Schwartz GJ and Casey JR (2003). Direct extracellular interaction between carbonic anhydrase IV and the human NBC1 sodium/bicarbonate co-transporter. Biochemistry, 42: 12321-12329. American Diabetes Association (1997). Clinical practice recommendations. Diabetes Care, 20 (Suppl 1): S1-S70. Anan F, Takahashi N, Ooie T, Hara M, Yoshimatsu H and Saikawa T (2004). Candesartan, an angiotensin II receptor blocker, improves left ventricular hypertrophy and insulin resistance. Metabolism, 53: 777-781. Andersson S and Moghrabi N (1997). Physiology and molecular genetics of 17 beta-hydroxysteroid dehydrogenases. Steroids, 62: 143-147. Andreka P, Zang J, Dougherty C, Slepak TI, Webster KA and Bishopric NH (2001). Cytoprotection by Jun kinase during nitric oxide-induced cardiac myocyte apoptosis. Circ Res, 88: 305-312. Anker SD, Ponikowski P, Varney S, Chua TP, Clark AL, Webb-Peploe KM, Harrington D, Kox WJ, Poole-Wilson PA and Coats AJ (1997). Wasting as independent risk factor for mortality in chronic heart failure. Lancet, 349: 1050-1053. Antoons G, Mubagwa K, Nevelsteen I and Sipido KR (2002). Mechanisms underlying the frequency dependence of contraction and [Ca2+]i transients in mouse ventricular myocytes. J Physiol, 543: 889-898. Anversa P, Leri A, Beltrami CA, Guerra S and Kajstura J (1998). Myocyte death and growth in the failing heart. Lab. Invest., 78: 767-786. Anversa P, Leri A, Kajstura J and Nadal-Ginard B (2002). Myocyte growth and cardiac repair. J Mol Cell Cardiol, 34: 91-105.

Bibliography

282

Aomine M and Yamato T (2000). Electrophysiological properties of ventricular muscle obtained from spontaneously diabetic mice. Exp Anim, 49: 23-33. Aoyama T, Peters JM, Iritani N, Nakajima T, Furihata K, Hashimoto T and Gonzalez FJ (1998). Altered constitutive expression of fatty acid-metabolizing enzymes in mice lacking the peroxisome proliferator-activated receptor alpha (PPARalpha). J Biol Chem, 273: 5678-5684. Arai M, Yoguchi A, Iso T, Takahashi T, Imai S, Murata K and Suzuki T (1995). Endothelin-1 and its binding sites are upregulated in pressure overload cardiac hypertrophy. Am J Physiol, 268: H2084-H2091. Arai M, Suzuki T and Nagai R (1996). Sarcoplasmic reticulum genes are upregulated in mild cardiac hypertrophy but downregulated in severe cardiac hypertrophy induced by pressure overload. J Mol Cell Cardiol, 28: 1583-1590. Argaman A and Livne A (1988). Angiotensin II and Na+/H+ exchange in human blood platelets. J Hum Hypertens, 2: 161-166. Armoundas AA, Wu R, Juang G, Marban E and Tomaselli GF (2001). Electrical and structural remodeling of the failing ventricle. Pharmacol Ther, 92: 213-230. Arnold JM, Yusuf S, Young J, Mathew J, Johnstone D, Avezum A, Lonn E, Pogue J, Bosch J and HOPE Investigators (2003). Prevention of Heart Failure in Patients in the Heart Outcomes Prevention Evaluation (HOPE) Study. Circulation, 107: 1284-1290. Asakawa M, Takano H, Nagai T, Uozumi H, Hasegawa H, Kubota N, Saito T, Masuda Y, Kadowaki T and Komuro I (2002). Peroxisome proliferator-activated receptor gamma plays a critical role in inhibition of cardiac hypertrophy in vitro and in vivo. Circulation, 105: 1240-1246. Ashavaid TF, Colvin RA, Messineo FC, MacAlister T and Katz AM (1985). Effects of fatty acids on Na/Ca exchange in cardiac sarcolemmal membranes. J Mol Cell Cardiol, 17: 851-861. Ashkenazi A and Dixit VM (1998). Death receptors: signaling and modulation. Science, 281: 1305-1308. Assayag P, Charlemagne D, de Leiris J, Boucher F, Valere PE, Lorte S, Swynghedauw B and Besse S (1997). Senescent heart compared with pressure overload-induced hypertrophy. Hypertension, 29: 15-21. Avendano GF, Agarwal RK, Bashey RI, Lyons MM, Soni BJ, Jyothirmayi GN and Regan TJ (1999). Effects of glucose intolerance on myocardial function and collagen-linked glycation. Diabetes, 48: 1443-1447. Avogaro A, Nosadini R, Doria A, Fioretto P, Velussi M, Vigorito C, Sacca L, Toffolo G, Cobelli C, Trevisan R, et al. (1990). Myocardial metabolism in insulin-deficient diabetic humans without coronary artery disease. Am J Physiol, 258: E606-E618.

Bibliography

283

Azhar M, Schultz Jel J, Grupp I, Dorn GW 2nd, Meneton P, Molin DG, Gittenberger-de Groot AC and Doetschman T (2003). Transforming growth factor beta in cardiovascular development and function. Cytokine Growth Factor Rev, 14: 391-407. Baartscheer A, Schumacher CA, van Borren MM, Belterman CN, Coronel R, Opthof T and Fiolet JW (2005). Chronic inhibition of Na+/H+-exchanger attenuates cardiac hypertrophy and prevents cellular remodeling in heart failure. Cardiovasc Res, 65: 83-92. Badaue-Passos D Jr, Ventura RR, Silva LF, Olivares EL, Ramalho MJ, Antunes Rodrigues J and Reis LC (2001). Effect of losartan on sodium appetite of hypothyroid rats subjected to water and sodium depletion and water, sodium and food deprivation. Exp Physiol, 86: 621-628. Badenhorst D, Veliotes D, Maseko M, Tsotetsi OJ, Brooksbank R, Naidoo A, Woodiwiss AJ and Norton GR (2003). Beta-adrenergic activation initiates chamber dilatation in concentric hypertrophy. Hypertension, 41: 499-504. Baker KM and Singer HA (1988). Identification and characterization of guinea pig angiotensin II ventricular and atrial receptors: coupling to inositol phosphate production. Circ Res, 62: 896-904. Ballard C and Schaffer S (1996). Stimulation of the Na+/Ca2+ exchanger by phenylephrine, angiotensin II and endothelin 1. J Mol Cell Cardiol, 28: 11-17. Banyasz T, Kalapos I, Kelemen SZ and Kovacs T (1996). Changes in cardiac contractility in IDDM and NIDDM diabetic rats. Gen Physiol Biophys, 15: 357-369. Barany M (1967). ATPase activity of myosin correlated with speed of muscle shortening. J Gen Physiol, 50 (Suppl): 197-218. Bardales RH, Hailey LS, Xie SS, Schaefer RF and Hsu SM (1996). In situ apoptosis assay for detection of early acute myocardial infarction. Am J Pathol, 149: 821-829. Barger PM, Brandt JM, Leone TC, Weinheimer CJ and Kelly DP (2000). Deactivation of peroxisome proliferator-activated receptor-α during cardiac hypertrophic growth. J Clin Invest, 105: 1723-1730. Barki-Harrington L, Luttrell LM and Rockman HA (2003). Dual inhibition of beta-adrenergic and angiotensin II receptors by a single antagonist: a functional role for receptor-receptor interaction in vivo. Circulation, 108: 1611-1618. Barnes JA and Smoak IW (2000). Glucose-regulated protein 78 (GRP78) is elevated in embryonic mouse heart and induced following hypoglycemic stress. Anat Embryol (Berl), 202: 67-74. Barrans JD, Allen PD, Stamatiou D, Dzau VJ and Liew CC (2002). Global gene expression profiling of end-stage dilated cardiomyopathy using a human cardiovascular-based cDNA microarray. Am J Pathol, 160: 2035-2043.

Bibliography

284

Barry WH, Matsui H, Bridge JH and Spitzer KW (1995). Excitation-contraction coupling in ventricular myocytes: effects of angiotensin II. Adv Exp Med Biol, 382: 31-39. Bassani RA, Bassani JW and Bers DM (1992). Mitochondrial and sarcolemmal Ca2+ transport reduce [Ca2+]i during caffeine contractures in rabbit cardiac myocytes. J Physiol, 453: 591-608. Bassani JWM, Bassani RA and Bers DM (1994). Relaxation in rabbit and rat cardiac cells: species-dependent differences in cellular mechanisms. J Physiol, 476: 279-293. Basso C, Calabrese F, Corrado D and Thiene G (2001). Postmortem diagnosis in sudden cardiac death victims: macroscopic, microscopic and molecular findings. Cardiovasc Res, 50: 290-300. Basso N and Terragno NA (2001). History about the discoveryof the renin-angiotensin system. Hypertension, 38: 1246-1249. Becker M, Umrani D, Lokhandwala MF and Hussain T (2003). Increased renal angiotensin II AT1 receptor function in obese Zucker rat. Clin exp Hypertens, 25: 35-47. Beddington RS and Robertson EJ (1999). Axis development and early asymmetry in mammals. Cell, 96: 195-209. Belke DD, Betuing S, Tuttle MJ, Graveleau C, Young ME, Pham M, Zhang D, Cooksey RC, McClain DA, Litwin SE, Taegtmeyer H, Severson D, Kahn CR and Abel ED (2002). Insulin signaling coordinately regulates cardiac size, metabolism, and contractile protein isoform expression. J Clin Invest, 109: 629-639. Belke DD, Larsen TS, Gibbs EM and Severson DL (2000). Altered metabolism causes cardiac dysfunction in perfused hearts from diabetic (db/db) mice. Am J Physiol, 279: E1104-E1113. Belke DD, Larsen TS, Gibbs EM and Severson DL (2001). Glucose metabolism in perfused mouse hearts overexpressing human GLUT-4 glucose transporter. Am J Physiol, 280: E420-E427. Belloni AS, Andreis PG, Macchi V, Gottardo G, Malendowicz LK and Nussdorfer GG (1998). Distribution and functional significance of angiotensin-II AT1- and AT2-receptor subtypes in the rat adrenal gland. Endocr Res, 24: 1-15. Beltrami CA, Finato N, Rocco M, Feruglio GA, Puricelli C, Cigola E, Quaini F, Sonnenblick EH, Olivetti G and Anversa P (1994). Structural basis of end-stage failure in ischemic cardiomyopathy in humans. Circulation, 89: 151-163. Berridge MJ, Lipp P and Bootman MD (2000). The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol, 1: 11-21.

Bibliography

285

Bers DM (2001). In: ‘Excitation-contraction coupling and cardiac contractile force’. Kluwe Academic Publishers, (2nd Edition). Berry MF, Woo YJ, Pirolli TJ, Bish LT, Moise MA, Burdick JW, Morine KJ, Jayasankar V, Gardner TJ and Sweeney HL (2004). Administration of a tumor necrosis factor inhibitor at the time of myocardial infarction attenuates subsequent ventricular remodeling. J Heart Lung Transplant, 23: 1061-1068. Beuckelmann DJ, Nabauer M and Erdmann E (1992). Intracellular calcium handling in isolated ventricular myocytes from patients with terminal heart failure. Circulation, 85: 1046-1055. Beuckelmann DJ, Nabauer M and Erdmann E (1993). Alterations of K+ currents in isolated human ventricular myocytes from patients with terminal heart failure. Circ Res, 73: 379-385. Bing OH, Conrad CH, Boluyt MO, Robinson KG and Brooks WW (2002). Studies of prevention, treatment and mechanisms of heart failure in the aging spontaneously hypertensive rat. Heart Fail Rev, 7: 71-88. Bischoff C, Kahns S, Lund A, Jorgensen HF, Praestegaard M, Clark BF and Leffers H (2000). The human elongation factor 1 A-2 gene (EEF1A2): complete sequence and characterization of gene structure and promoter activity. Genomics, 68: 63-70. Bishop SP and Altschuld RA (1970). Increased glycolytic metabolism in cardiac hypertrophy and congestive failure. Am J Physiol, 218: 153-159. Bishop JE, Greenbaum R, Gibson DG, Yacoub M and Laurent GJ (1990). Enhanced deposition of predominantly type I collagen in myocardial disease. J Mol Cell Cardiol, 22: 1157-1165. Blair-West JR, Coghlan JP, Denton DA, Funder JW, Scoggins BA and Wright RD (1971). The effect of the heptapeptide (2-8) and hexapeptide (3-8) fragments of angiotensin II on aldosterone secretion. J Clin Endocrinol Metab, 32: 575-578. Blanchard AP and Hood L (1996). Sequence to array: probing the genome's secrets. Nat Biotechnol, 14: 1649. Boateng SY, Naqvi RU, Koban MU, Yacoub MH, MacLeod KT and Boheler KR (2001). Low-dose ramipril treatment improves relaxation and calcium cycling after established cardiac hypertrophy. Am J Physiol, 280: H1029-H1038. Boateng SY, Seymour AM, Bhutta NS, Dunn MJ, Yacoub MH and Boheler KR (1998). Sub-antihypertensive doses of ramipril normalize sarcoplasmic reticulum calcium ATPase expression and function following cardiac hypertrophy in rats. J Mol Cell Cardiol, 30: 2683-2694.

Bibliography

286

Bohm M, Lippoldt A, Wienen W, Ganten D and Bader M (1996). Reduction of cardiac hypertrophy in TGR(mREN2)27 by angiotensin II receptor blockade. Mol Cell Biochem, 163-164: 217-221. Bohm M, Flesch M and Schnabel P (1997). Beta-adrenergic signal transduction in the failing and hypertrophied myocardium. J Mol Med, 75: 842–848. Boluyt MO, Robinson KG, Meredith AL, Sen S, Lakatta EG, Crow MT, Brooks WW, Conrad CH and Bing OH (2005). Heart failure after long-term supravalvular aortic constriction in rats. Am J Hypertens, 18: 202-212. Bomberger JM, Spielman WS, Hall CS, Weinman EJ and Parameswaran N (2005). Receptor activity-modifying protein (RAMP) isoform-specific regulation of adrenomedullin receptor trafficking by NHERF-1. J Biol Chem, 280: 23926-23935. Bond J, Sedmera D, Jourdan J, Zhang Y, Eisenberg CA, Eisenberg LM and Gourdie RG (2003). Wnt11 and Wnt7a are up-regulated in association with differentiation of cardiac conduction cells in vitro and in vivo. Dev Dyn, 227: 536-543. Bonne G, Carrier L, Richard P, Hainque B and Schwartz K (1998). Familial hypertrophic cardiomyopathy: from mutations to functional defects. Circ Res, 83: 580-593. Bonneux L, Barendregt JJ, Meeter K, Bonsel GJ and van der Maas PJ (1994). Estimating clinical morbidity due to ischaemic heart disease and congestive heart failure: the future rise of heart failure. Am J Public Health, 84: 20-28. Bouzegrhane F and Thibault G (2002). Is angiotensin II a proliferative factor of cardiac fibroblasts? Cardiovasc Res, 53: 304-312. Bowling N, Huang X, Sandusky GE, Fouts RL, Mintze K, Esterman M, Allen PD, Maddi R, McCall E and Vlahos CJ (2001). Protein kinase C-alpha and –epsilon modulate connexin-43 phosphorylation in human heart. J Mol Cell Cardiol, 33: 789-798. Braun-Mendez E and Page IH (1958). Suggested revision of nomenclature: angiotensin. Science, 127: 242. Braz JC, Bueno OF, Liang Q, Wilkins BJ, Dai YS, Parsons S, Braunwart J, Glascock BJ, Klevitsky R, Kimball TF, Hewett TE and Molkentin JD (2003). Targeted inhibition of p38 MAPK promotes hypertrophic cardiomyopathy through upregulation of calcineurin-NFAT signaling. J Clin Invest, 111: 1475-1486. Brede M, Roell W, Ritter O, Wiesmann F, Jahns R, Haase A, Fleischmann BK and Hein L (2003). Cardiac hypertrophy is associated with decreased eNOS expression in angiotensin AT2 receptor-deficient mice. Hypertension, 42: 1177-1182.

Bibliography

287

Breen EC and Tang K (2003). Calcyclin (S100A6) regulates pulmonary fibroblast proliferation, morphology, and cytoskeletal organization in vitro. J Cell Biochem, 88: 848-854. Brennan KJ and Hardeman EC (1993). Quantitative analysis of the human alpha-skeletal actin gene in transgenic mice. J Biol Chem, 268: 719-725. Brilla CG, Janicki JS and Weber KT (1991). Cardioreparative effects of lisinopril in rats with genetic hypertension and left ventricular hypertrophy. Circulation, 83: 1771-1779. Brilla CG, Rupp H, Funck R and Maisch B (1995). The renin-angiotensin-aldosterone system and myocardial collagen matrix remodelling in congestive heart failure. Eur Heart J, 16 (Suppl O):107-109. Brillantes AM, Allen P, Takahashi T, Izumo S and Marks AR (1992). Differences in cardiac calcium release channel (ryanodine receptor) expression in myocardium from patients with end-stage heart failure caused by ischemic versus dilated cardiomyopathy. Circ Res, 71: 18-26. Brink M, de Gasparo M, Rogg H, Whitebread S and Bullock G (1995). Localization of angiotensin II receptor subtypes in the rabbit heart. J Mol Cell Cardiol, 27: 459-470. Brink M, Erne P, de Gasparo M, Rogg H, Schmid A, Stulz P and Bullock G (1996). Localization of the angiotensin II receptor subtypes in the human atrium. J Mol Cell Cardiol, 28: 1789-1799. Brixius K, Hoischen S, Reuter H, Lasek K and Schwinger RH (2001). Force/shortening-frequency relationship in multicellular muscle strips and single cardiomyocytes of human failing and nonfailing hearts. J Card Fail, 7: 335-341. Brooks WW and Apstein CS (1996). Effect of treppe on isovolumic function in the isolated blood-perfused mouse heart. J Mol Cell Cardiol, 28: 1817-1822. Brooks WW, Bing OH, Boluyt MO, Malhotra A, Morgan JP, Satoh N, Colucci WS and Conrad CH (2000). Altered inotropic responsiveness and gene expression of hypertrophied myocardium with captopril. Hypertension, 35: 1203-1209. Brooksby P, Levi AJ and Jones JV (1993). The electrophysiological characteristics of hypertrophied ventricular myocytes from the spontaneously hypertensive rat. J Hypertens, 11: 611-622. Broome M, Hanley M, Haggmark S, Johansson G, Aneman A and Biber B (2001). Acute effects of angiotensin II on myocardial performance. Acta Anaesthesiol Scand, 45: 1147-1154. Brown CO 3rd, Chi X, Garcia-Gras E, Shirai M, Feng XH and Schwartz RJ (2004). The cardiac determination factor, Nkx2-5, is activated by mutual cofactors GATA-4 and Smad1/4 via a novel upstream enhancer. J Biol Chem, 279: 10659-10669.

Bibliography

288

Bruno E, Rossi N, Thuer O, Cordoba R and Alday LE (2003). Cardiovascular findings, and clinical course, in patients with Williams syndrome. Cardiol Young, 13: 532-536. Burgess JK (2001). Gene expression studies using microarrays. Clin Exp Pharmacol Physiol, 28: 321-328. Burgess ML, Carver WE, Terracio L, Wilson SP, Wilson MA and Borg TK (1994). Integrin-mediated collagen gel contraction by cardiac fibroblasts. Effects of angiotensin II. Circ Res, 74: 291-298. Burke W, Imperatore G and Reyes M (2001). Iron deficiency and iron overload: effects of diet and genes. Proc Nutr Soc, 60: 73-80. Burson JM, Aguilera G, Gross KW and Sigmund CD (1994). Differential expression of angiotensin receptor 1A and 1B in mouse. Am J Physiol, 267: E260-E267. Cadre BM, Qi M, Eble DM, Shannon TR, Bers DM and Samarel AM (1998). Cyclic stretch down-regulates calcium transporter gene expression in neonatal rat ventricular myocytes. J Mol Cell Cardiol, 30: 2247-2259. Cai L, Li W, Wang G, Guo L, Jiang Y and Kang YJ (2002). Hyperglycemia-induced apoptosis in mouse myocardium: mitochondrial cytochrome C-mediated caspase-3 activation pathway. Diabetes, 51: 1938-1948. Cambien F, Poirier O, Lecerf L, Evans A, Cambou JP, Arveiler D, Luc G, Bard JM, Bara L, Ricard S, et al. (1992). Deletion polymorphism in the gene for angiotensin-converting enzyme is a potent risk factor for myocardial infarction. Nature, 359: 641-644. Camilion de Hurtado MC, Alvarez BV, Perez NG, Ennis IL and Cingolani HE (1998). Angiotensin II activates Na+-independent Cl--HCO3

- exchange in ventricular myocardium. Circ Res, 82: 473-481. Campbell DJ (1987). Circulating and tissue angiotensin systems. J Clin Invest, 79: 1-6. Campbell DJ, Bouhnik J, Coezy E, Menard J and Corvol P (1985). Characterization of precursor and secreted forms of human angiotensinogen. J Clin Invest, 75: 1880-1893. Campbell SE and Katwa LC (1997). Angiotensin II stimulated expression of transforming growth factor-beta1 in cardiac fibroblasts and myofibroblasts. J Mol Cell Cardiol, 29: 1947-1958. Candido R, Forbes JM, Thomas MC, Thallas V, Dean RG, Burns WC, Tikellis C, Ritchie RH, Twigg SM, Cooper ME and Burrell LM (2003). A breaker of advanced glycation end products attenuates diabetes-induced myocardial structural changes. Circ Res, 92: 785-792.

Bibliography

289

Capasso JM, Li P, Zhang X, Meggs LG and Anversa P (1993). Alterations in Ang II responsiveness in left and right myocardium after infarction-induced heart failure in rats. Am J Physiol, 264: H2056-H2067. Capogrossi MC, Kort AA, Spurgeon HA and Lakatta EG (1986). Single adult rabbit and rat cardiac myocytes retain the Ca2+- and species-dependent systolic and diastolic contractile properties of intact muscle. J Gen Physiol, 88: 589-613. Capogrossi MC, Suarez-Isla BA and Lakatta EG (1986). The interaction of electrically stimulated twitches and spontaneous contractile waves in single cardiac myocytes. J Gen Physiol, 88: 615-633. Carluccio E, Tommasi S, Bentivoglio M, Buccolieri M, Filippucci L, Prosciutti L and Corea L (2000). Prognostic value of left ventricular hypertrophy and geometry in patients with a first, uncomplicated myocardial infarction. Int J Cardiol, 74: 177-183. Caspari PG, Newcomb M, Gibson K and Harris P (1977). Collagen in the normal and hypertrophied human ventricle. Cardiovasc Res, 11: 554-558. Castellano M and Bohm M (1997). The cardiac beta-adrenoceptor-mediated signaling pathway and its alterations in hypertensive heart disease. Hypertension, 29: 715-722. Castello A, Rodriguez-Manzaneque JC, Camps M, Perez-Castillo A, Testar X, Palacin M, Santos A and Zorzano A (1991). Perinatal hypothyroidism impairs the normal transition of GLUT4 and GLUT1 glucose transporters from fetal to neonatal levels in heart and brown adipose tissue. J Biol Chem, 269: 5905-5912. Celentano A, Vaccaro O, Tammaro P, Galderisi M, Crivaro M, Oliviero M, Imperatore G, Palmieri V, Iovino V and Riccardi G (1995). Early abnormalities of cardiac function in non-insulin-dependent diabetes mellitus and impaired glucose tolerance. Am J Cardiol, 76: 1173-1176. Cerbai E, Barbieri M, Li Q and Mugelli A (1994). Ionic basis of action potential prolongation of hypertrophied cardiac myocytes isolated from hypertensive rats of different ages. Cardiovasc Res, 28: 1180-1187. Cerbai E, Pino R, Porciatti F, Sani G, Toscano M, Maccherini M, Giunti G and Mugelli A (1997). Characterization of the hyperpolarization-activated current, If, in ventricular myocytes from human failing heart. Circulation, 95: 568–571. Cerbai E, Sartiani L, DePaoli P, Pino R, Maccherini M, Bizzarri F, DiCiolla F, Davoli G, Sani G and Mugelli A (2001). The properties of the pacemaker current I(F) in human ventricular myocytes are modulated by cardiac disease. J Mol Cell Cardiol, 33: 441–448. Ch'en FF, Vaughan-Jones RD, Clarke K and Noble D (1998). Modelling myocardial ischaemia and reperfusion. Prog Biophys Mol Biol, 69: 515-538.

Bibliography

290

Chahine M, Bkaily G, Nader M, Al-Khoury J, Jacques D, Beier N and Scholz W (2005). NHE-1-dependent intracellular sodium overload in hypertrophic hereditary cardiomyopathy: prevention by NHE-1 inhibitor. J Mol Cell Cardiol, 38: 571-582. Chan CP, Sanderson JE, Glatz JF, Cheng WS, Hempel A and Renneberg R (2004). A superior early myocardial infarction marker. Human heart-type fatty acid-binding protein. Z Kardiol, 93: 388-397. Chancey AL, Brower GL, Peterson JT and Janicki JS (2002). Effects of matrix metalloproteinase inhibition on ventricular remodeling due to volume overload. Circulation, 105: 1983-1988. Chapman D, Weber KT and Eghbali M (1990). Regulation of fibrillar collagen types I and III and basement membrane type IV collagen gene expression in pressure overloaded rat myocardium. Circ Res, 67: 787-794. Charron MJ, Katz EB and Olson AL (1999). GLUT4 gene regulation and manipulation. J Biol Chem, 274: 3253-3256. Chattou S, Diacono J and Feuvray D (1999). Decrease in sodium-calcium exchange and calcium currents in diabetic rat ventricular myocytes. Acta Physiol Scand, 166: 137-144. Chen F, Jarmakani JM and Van Dop C (1995). Developmental changes in mRNA encoding cardiac Na+/H+ exchanger (NHE-1) in rabbit. Biochem Biophys Res Commun, 212: 960-967. Chen X, Piacentino V 3rd, Furukawa S, Goldman B, Margulies KB and Houser SR (2002). L-type Ca2+ channel density and regulation are altered in failing human ventricular myocytes and recover after support with mechanical assist devices. Circ Res, 91: 517-524. Chen L, Chen CX, Gan XT, Beier N, Scholz W and Karmazyn M (2004) Inhibition and reversal of myocardial infarction-induced hypertrophy and heart failure by NHE-1 inhibition. Am J Physiol, 286: H381–H387. Chen Y, Rajashree R, Liu Q and Hofmann P (2003). Acute p38 MAPK activation decreases force development in ventricular myocytes. Am J Physiol, 285: H2578-H2586. Cheng CP, Suzuki M, Ohte N, Ohno M, Wang ZM and Little WC (1996). Altered ventricular and myocyte response to angiotensin II in pacing-induced heart failure. Circ Res, 78: 880-892. Cheng W, Li B, Kajstura J, Li P, Wolin MS, Sonnenblick EH, Hintze TH, Olivetti G and Anversa P (1995). Stretch-induced programmed myocyte cell death. J Clin Invest, 96: 2247-2259.

Bibliography

291

Chirgwin JM, Schaefer IM, Diaz JA and Lalley PA (1984). Mouse kidney renin gene is on chromosome one. Somat Cell Molec Genet, 10: 633-637. Choi KM, Zhong Y, Hoit BD, Grupp IL, Hahn H, Dilly KW, Guatimosim S, Lederer WJ and Matlib MA (2002). Defective intracellular Ca2+ signaling contributes to cardiomyopathy in type 1 diabetic rats. Am J Physiol, 283: H1398-H1408. Ciani E, Severi S, Contestabile A, Bartesaghi R and Contestabile A (2004). Nitric oxide negatively regulates proliferation and promotes neuronal differentiation through N-Myc downregulation. J Cell Sci, 117: 4727-4737. Cicoira M, Bolger AP, Doehner W, Rauchhaus M, Davos C, Sharma R, Al-Nasser FO, Coats AJ and Anker SD (2001). High tumour necrosis factor-alpha levels are associated with exercise intolerance and neurohormonal activation in chronic heart failure patients. Cytokine, 15: 80-86. Cingolani HE, Alvarez BV, Ennis IL and Camilion de Hurtado MC (1998). Stretch-induced alkalinization of feline papillary muscle. An autocrine-paracrine system. Circ Res, 83: 775-780. Clark AF, Sharp MG, Morley SD, Fleming S, Peters J and Mullins JJ (1997). Renin-1 is essential for normal renal juxtaglomerular cell granulation and macula densa morphology. J Biol Chem, 272: 18185-18190. Clauser E, Curnow KM, Davies E, Conchon S, Teutsch B, Vianello B, Monnot C and Corvol P (1996). Angiotensin II receptors: protein and gene structures, expression and potential pathological involvements. Eur J Endocrinol, 134: 403-411. Clement S, Pellieux C, Chaponnier C, Pedrazzini T and Gabbiani G (2001). Angiotensin II stimulates α-skeletal actin expression in cardiomyocytes in vitro and in vivo in the absence of hypertension. Differentiation, 69: 66-74. Clerico A and Emdin M (2004). Diagnostic accuracy and prognostic relevance of the measurement of cardiac natriuretic peptides: a review. Clin Chem, 50: 33-50. Codd MB, Sugrue DD, Gersh BJ and Melton 3rd LJ (1989). Epidemiology of idiopathic dilated and hypertrophic cardiomyopathy: a population-based study in Olmsted County, Minnesota, 1975–1984. Circulation, 80: 564–572. Coker ML, Jolly JR, Joffs C, Etoh T, Holder JR, Bond BR and Spinale FG (2001). Matrix metalloproteinase expression and activity in isolated myocytes after neurohormonal stimulation. Am J Physiol, 281: H543-H551. Collins AR, Schnee J, Wang W, Kim S, Fishbein MC, Bruemmer D, Law RE, Nicholas S, Ross RS and Hsueh WA (2004). Osteopontin modulates angiotensin II-induced fibrosis in the intact murine heart. J Am Coll Cardiol, 43: 1698-1705. Cooklin M, Wallis WR, Sheridan DJ and Fry CH (1997). Changes in cell-to-cell electrical coupling associated with left ventricular hypertrophy. Circ Res, 80: 765–771.

Bibliography

292

Cooper EL (1976). Evolution of blood cells. Ann Immunol (Paris), 127: 817-825. Coppen SR, Dupont E, Rothery S and Severs NJ (1998). Connexin45 expression is preferentially associated with the ventricular conduction system in mouse and rat heart. Circ Res, 82: 232–243. Cowie MR, Mosterd A, Wood DA, Deckers JW, Poole-Wilson PA, Sutton GC and Grobbee DE (1997). The epidemiology of heart failure. Eur Heart J, 18: 208-225. Crackower MA, Sarao R, Oudit GY, Yagil C, Kozieradzki I, Scanga SE, Oliveira-dos-Santos AJ, da Costa J, Zhang L, Pei Y, Scholey J, Ferrario CM, Manoukian AS, Chappell MC, Backx PH, Yagil Y and Penniger JM (2002). Angiotensin-converting enzyme 2 is an essential regulator of heart function. Nature, 417: 822-828. Crawford HC, Fingleton BM, Rudolph-Owen LA, Goss KJ, Rubinfeld B, Polakis P and Matrisian LM (1999). The metalloproteinase matrilysin is a target of beta-catenin transactivation in intestinal tumors. Oncogene, 18: 2883-2891. Criddle DN, Dewar GH, Wathey WB and Woodward (1990). The effects of novel vasodilator long chain acyl carnitine ester in the isolated perfused heart of the rat. Br J Pharmacol, 99: 477-480. Cubillos-Garzon LA, Casas JP, Morillo CA and Bautista LE (2004). Congestive heart failure in Latin America: the next epidemic. Am Heart J, 147: 412-417. D'Angelo DD, Sakata Y, Lorenz JN, Boivin GP, Walsh RA, Liggett SB, Dorn GW 2nd (1997). Transgenic Galphaq overexpression induces cardiac contractile failure in mice. Proc Natl Acad Sci U S A, 94: 8121-8126. Danik SB, Liu F, Zhang J, Suk HJ, Morley GE, Fishman GI and Gutstein DE (2004). Modulation of cardiac gap junction expression and arrhythmic susceptibility. Circ Res, 95: 1035-1041. Dargie HJ, McMurray JJV and McDonald TA (1996). Heart failure- implications of the true size of the problem. J Intern Med, 239: 309-315. Darrow BJ, Fast VG, Kleber AG, Beyer EC and Saffitz JE (1996). Functional and structural assessment of intercellular communication: increased conduction velocity and enhanced connexin expression in dibutyryl cAMP-treated cultured cardiac myocytes. Circ Res, 79: 174–183. Das UN (2004). Long-chain polyunsaturated fatty acids interact with nitric oxide, superoxide anion, and transforming growth factor-beta to prevent human essential hypertension. Eur J Clin Nutr, 58: 195-203. Davenport JW, Fernandes ER, Harris LD, Neale GA and Goorha R (1999). The mouse mitotic checkpoint gene bub1b, a novel bub1 family member, is expressed in a cell cycle-dependent manner. Genomics, 55: 113-117.

Bibliography

293

Davies CH, Davia K, Bennett JG, Pepper JR, Poole-Wilson PA and Harding SE (1995). Reduced contraction and altered frequency response of isolated ventricular myocytes from patients with heart failure. Circulation, 92: 2540-2549. Davis LM, Rodefeld ME, Green K, Beyer EC and Saffitz JE (1995). Gap junction protein phenotypes of the human heart and conduction system. J Cardiovasc Electrophysiol, 6: 813–822. Day FL, Rafty LA, Chesterman CN and Khachigian LM (1999). Angiotensin II (ATII)-inducible platelet-derived growth factor A-chain gene expression is p42/44 extracellular signal-regulated kinase-1/2 and Egr-1-dependent and mediated via the ATII type 1 but not type 2 receptor. Induction by ATII antagonized by nitric oxide. J Biol Chem, 274: 23726-23733. de Gasparo M, Husain A, Alexander W, Catt KJ, Chiu AT, Drew M, Goodfriend T, Harding JW, Inagami T and Timmermans PB (1995). Proposed update of angiotensin receptor nomenclature. Hypertension, 25: 924-927. De La Bastie D, Levitsky D, Rappaport L, Mercadier JJ, Marotte F, Wisnewsky C, Brokovich V, Schwartz K and Lompre AM (1990). Function of the sarcoplasmic reticulum and expression of its Ca2+ ATPase gene in pressure overload-induced cardiac hypertrophy in the rat. Circ Res, 66: 554-564. de Launoit Y and Adamski J (1999). Unique multifunctional HSD17B4 gene product: 17beta-hydroxysteroid dehydrogenase 4 and D-3-hydroxyacyl-coenzyme A dehydrogenase/hydratase involved in Zellweger syndrome. J Mol Endocrinol, 22: 227-240. De Mello WC and Altieri PI (1992). The role of the renin-angiotensin system in the control of cell communication in heart; effects of angiotensin II and enalapril. J Cardiovasc Pharmacol, 20: 643-651. De Mello WC (1996). Renin-angiotensin system and cell communication in the failing heart. Hypertension, 27: 1267-1272. De Mello WC (1998). Intracellular angiotensin II regulates the inward calcium current in cardiac myocytes. Hypertension, 32: 976-982. de Simone G and Palmieri V (2002). Left ventricular hypertrophy in hypertension as a predictor of coronary events: relation to geometry. Curr Opin Nephrol Hypertens, 11: 215-220. De Windt LJ, Lim HW, Taigen T, Wencker D, Condorelli G, Dorn GW 2nd, Kitsis RN and Molkentin JD (2000). Calcineurin-mediated hypertrophy protects cardiomyocytes from apoptosis in vitro and in vivo: An apoptosis-independent model of dilated heart failure. Circ Res, 86: 255-263. Debouck C and Goodfellow PN (1999). DNA microarrays in drug discovery and development. Nat Genet, 21(Suppl 1): 48-50.

Bibliography

294

Decker RS, Decker ML, Nakamura S, Zhao YS, Hedjbeli S, Harris KR and Klocke FJ (2002). HSC73-tubulin complex formation during low-flow ischemia in the canine myocardium. Am J Physiol, 283: H1322-H1333. DeFronzo RA (1992). Insulin resistance, hyperinsulinemia, and coronary artery disease: a complex metabolic web. J Cardiovasc Parmacol, 20 (Suppl 11): S1-S16. Delbridge LM, Harris PJ and Morgan TO (1989). Characterization of single heart cell contractility by rapid imaging. Clin Exp Pharmacol Physiol, 16: 179-184. Delbridge LM, Harris PJ, Pringle JT, Dally LJ and Morgan TO (1990). A superfusion bath for single-cell recording with high-precision optical depth control, temperature regulation, and rapid solution switching. Pflűgers Arch, 416: 94-97. Delbridge LM, Morgan TO and Harris PJ (1995). Effects of endothelin-1 on the contractility of cardiomyocytes from the spontaneously hypertensive rats. Clin Exp Pharmacol Physiol, 22: 755-762. Delbridge LM, Satoh H, Yuan W, Bassani JW, Qi M, Ginsburg KS, Samarel AM and Bers DM (1997). Cardiac myocyte volume, Ca2+ fluxes, and sarcoplasmic reticulum loading in pressure-overload hypertrophy. Am J Physiol, 272: H2425-H2435. Denker SP and Barber DL (2002). Cell migration requires both ion translocation and cytoskeletal anchoring by the Na-H exchanger NHE1. J Cell Biol, 159: 1087-1096. Denker SP, Huang DC, Orlowski J, Furthmayr H and Barber DL (2000). Direct binding of the Na--H exchanger NHE1 to ERM proteins regulates the cortical cytoskeleton and cell shape independently of H(+) translocation. Mol Cell, 6: 1425-1436. Depre C, Shipley GL, Chen W, Han Q, Doenst T, Moore ML, Stepkowski S, Davies PJ and Taegtmeyer H (1999a). Unloaded heart in vivo replicates fetal gene expression of cardiac hypertrophy. Nat Med, 4: 1269-1275. Desrois M, Sidell RJ, Gauguier D, King LM, Radda GK and Clarke K (2004). Initial steps of insulin signaling and glucose transport are defective in the type 2 diabetic rat heart. Cardiovasc Res, 61: 288-296. Desvergne B and Wahli W (1999). Peroxisome proliferator-activated receptors. Nuclear control of metabolism. Endocr Rev, 20: 649-688. Dhein S, Polontchouk L, Salameh A and Haefliger JA (2002). Pharmacological modulation and differential regulation of the cardiac gap junction proteins connexin 43 and connexin 40. Biol Cell, 94: 409-422. Diep QN, Amiri F, Touyz RM, Cohn JS, Endemann D, Neves MF and Schiffrin EL (2002). PPARalpha activator effects on Ang II-induced vascular oxidative stress and inflammation. Hypertension, 40: 866-871.

Bibliography

295

Dierks T, Volkmer J, Schlenstedt G, Jung C, Sandholzer U, Zachmann K, Schlotterhose P, Neifer K, Schmidt B and Zimmermann R (1996). A microsomal ATP-binding protein involved in efficient protein transport into the mammalian endoplasmic reticulum. EMBO J, 15: 6931-6942. Dimri GP, Testori A, Acosta M and Campisi J (1996). Replicative senescence, aging and growth-regulatory transcription factors. Biol Signals, 5: 154-162. Dinh DT, Frauman AG, Sourial M, Casley DJ, Johnston CI and Fabiani ME (2001). Identification, distribution and expression of angiotensin II receptors in the normal human prostate and benign prostate hyperplasia. Endocrinology, 142: 1349-1356. Di Nicolantonio R, Kostka V, Chow MZ, Jansa P, Harrap SB. The Lvm-1 locus controlling left ventricular size contains a variant of the angiotensin II type 1b receptor. Rat Genomics and Models Meeting, 2003, Cold Spring Harbour. Dipla K, Mattiello JA, Margulies KB, Jeevanandam V and Houser SR (1999). The sarcoplasmic reticulum and the Na+/Ca2+ exchanger both contribute to the Ca2+ transient of failing human ventricular myocytes. Circ Res, 84: 435-444. Doble BW, Ping P and Kardami E (2000). The epsilon subtype of protein kinase C is required for cardiomyocyte connexion-43 phosphorylation. Circ Res, 86: 293-301. Dodge SM, Beardslee MA, Darrow BJ, Green KG, Beyer EC and Saffitz JE (1998). Effects of angiotensin II on expression of the gap junction channel protein connexin43 in neonatal rat ventricular myocytes. J Am Coll Cardiol, 32: 800-807. Domenighetti AA, Wang Q, Egger M, Richards SM, Pedrazzini T and Delbridge LM (2005). Angiotensin II-mediated phenotypic cardiomyocyte remodeling leads to age-dependent cardiac dysfunction and failure. Hypertension, 46: 426-432. Dorn 2nd GW and Hahn HS (2004). Genetic factors in cardiac hypertrophy. Ann N Y Acad Sci, 1015:225-237. Dorn GW, Robbins J, Ball N and Walsh RA (1994). Myosin heavy chain regulation and myocyte contractile depression after LV hypertrophy in aortic-banded mice. Am J Physiol, 267: H400-H405. Dostal DE (2000). The cardiac renin-angiotensin system: novel signaling mechanisms related to cardiac growth and function. Regul Pept, 91: 1-11. Dostal DE and Baker KM (1998). Angiotensin and endothelin: messengers that couple ventricular stretch to the Na+/H+ exchanger and cardiac hypertrophy. Circ Res, 83: 870-873. Dostal DE and Baker KM (1999). The cardiac renin-angiotensin system: conceptual, or a regulator of cardiac function ? Circ Res, 85: 643-650.

Bibliography

296

Dostal DE, Rothblum KC and Baker KM (1994). An improved method for absolute quantification of mRNA using multiplex polymerase chain reaction: determination of renin and angiotensinogen mRNA levels in various tissues. Anal Biochem, 223: 239-250. Drazner MH, Rame JE, Marino EK, Gottdiener JS, Kitzman DW, Gardin JM, Manolio TA, Dries DL and Siscovick DS (2004). Increased left ventricular mass is a risk factor for the development of a depressed left ventricular ejection fraction within five years: the Cardiovascular Health Study. J Am Coll Cardiol, 43: 2207-2215. Drobyshev A, Mologina N, Shik V, Pobedimskaya D, Yershov G and Mirzabekov A (1997). Sequence analysis by hybridization with oligonucleotide microchip: identification of beta-thalassemia mutations. Gene, 188: 45-52. Duan J, Zhang HY, Adkins SD, Ren BH, Norby FL, Zhan X, Benoit JN, Epstein PN and Ren J (2003). Impaired cardiac function and IGF-I response in myocytes from calmodulin-diabetic mice: role of Akt and RhoA. Am J Physiol Endocrinol Metab, 284: E366-E376. Dudoit S and Fridlyand J (2002). A prediction-based resampling method for estimating the number of clusters in a dataset. Genome Biol, 3: RESEARCH0036. Duggan DJ, Bittner M, Chen Y, Meltzer P and Trent JM (1999). Expression profiling using cDNA microarrays. Nat Genet, 21(Suppl 1):10-14. Dupont E, Matsushita T, Kaba RA, Vozzi C, Coppen SR, Khan N, Kaprielian R, Yacoub MH and Severs NJ (2001). Altered connexin expression in human congestive heart failure. J Mol Cell Cardiol, 33: 359-371. Dutta K, Podolin DA, Davidson MB and Davidoff AJ (2001). Cardiomyocyte dysfunction in sucrose-fed rats is associated with insulin resistance. Diabetes, 50: 1186-1192. Dutta K, Carmody MW, Cala SE and Davidoff AJ (2002). Depressed PKA activity contributes to impaired SERCA function and is linked to the pathogenesis of glucose-induced cardiomyopathy. J Mol Cell Cardiol, 34: 985-996. Dyck JR and Lopaschuk GD (1998). Glucose metabolism, H+ production and Na+/H+-exchanger mRNA levels in ischemic hearts from diabetic rats. Mol Cell Biochem, 180: 85-93 Eckel J and Reinauer H (1990). Insulin action on glucose transport in isolated cardiac myocytes: signalling pathways and diabetes-induced alterations. Biochem Soc Trans, 18: 1125-1127. Egan B, Gleim G and Panish J (2004). Use of losartan in diabetic patients in the primary care setting: review of the results in LIFE and RENAAL. Curr Med Res Opin, 20: 1909-1917.

Bibliography

297

Egeo A, Di Lisi1 R, Sandri C, Mazzocco M, Lapide M, Schiaffino S and Scartezzini P (2000). Developmental expression of the SH3BGR gene, mapping to the Down syndrome heart critical region. Mech Dev, 90: 313-316. Eguchi S, Iwasaki H, Ueno H, Frank GD, Motley ED, Eguchi K, Marumo F, Hirata Y and Inagami T (1999). Intracellular signaling of angiotensin II-induced p70 S6 kinase phosphorylation at Ser(411) in vascular smooth muscle cells. Possible requirement of epidermal growth factor receptor, Ras, extracellular signal-regulated kinase, and Akt. J Biol Chem, 274: 36843-36851. Elhendy A, Schinkel AF, van Domburg RT, Bax JJ and Poldermans D (2004). Incidence and predictors of heart failure during long-term follow-up after stress Tc-99m sestamibi tomography in patients with suspected coronary artery disease. J Nucl Cardiol, 11: 527-533. Elliott P (2000). Cardiomyopathy. Diagnosis and management of dilated cardiomyopathy. Heart 84: 106–112. Eloff BC, Gilat E, Wan X and Rosenbaum DS (2003). Pharmacological modulation of cardiac gap junctions to enhance cardiac conduction: evidence supporting a novel target for antiarrhythmic therapy. Circulation, 108: 3157-3163. Egert S, Nguyen N, Brosius FC 3rd and Schwaiger M (1997). Effects of wortmannin on insulin- and ischemia-induced stimulation of GLUT4 translocation and FDG uptake in perfused rat hearts. Cardiovasc Res, 35: 283-293. Egger M and Niggli E (1999). Regulatory function of Na-Ca exchange in the heart: milestones and outlook. J Membr Biol, 168: 107-130. Eguchi S, Iwasaki H, Ueno H, Frank GD, Motley ED, Eguchi K, Marumo F, Hirata Y and Inagami T (1999). Intracellular signaling of angiotensin II-induced p70 S6 kinase phosphorylation at Ser(411) in vascular smooth muscle cells. Possible requirement of epidermal growth factor receptor, Ras, extracellular signal-regulated kinase, and Akt. J Biol Chem, 274: 36843-36851. Engelhardt S, Hein L, Keller U, Klambt K and Lohse MJ (2002). Inhibition of Na+-H+ exchange prevents hypertrophy, fibrosis, and heart failure in β1-adrenergic receptor transgenic mice. Circ Res, 90: 814-819. Engels W, van Bilsen M, Wolffenbuttel BH, van der Vusse GJ and Glatz JF (1999). Cytochrome P450, peroxysome proliferation and cytoplasmic fatty acid-binding protein content in liver, heart and kidney of the diabetic rat. Mol Cell Biochem, 192: 53-61. Engler MM, Engler MB, Pierson DM, Molteni LB and Molteni A (2003). Effects of docosahexaenoic acid on vascular pathology and reactivity in hypertension. Exp Biol Med (Maywood), 228: 299-307.

Bibliography

298

Er F, Larbig R, Ludwig A, Biel M, Hofmann F, Beuckelmann DJ and Hoppe UC (2003). Dominant-negative suppression of HCN channels markedly reduces the native pacemaker current I(f) and undermines spontaneous beating of neonatal cardiomyocytes. Circulation, 107: 485–489. Esler M, Kaye D, Lambert G, Esler D and Jenning G (1997). Adrenergic nervous system in heart failure. Am J Cardiol, 80 (Suppl): 7L-14L. Everett AD, Fisher A, Tufro-McReddie A and Harris M (1997). Developmental regulation of angiotensin type 1 and 2 receptor gene expression and heart growth. J Mol Cell Cardiol, 29: 141-148. Fadel MP, Szewczenko-Pawlikowski M, Leclerc P, Dziak E, Symonds JM, Blaschuk O, Michalak M and Opas M (2001). Calreticulin affects beta-catenin-associated pathways. J Biol Chem, 276: 27083-27089. Falck S, Paavilainen VO, Wear MA, Grossmann JG, Cooper JA and Lappalainen P (2004). Biological role and structural mechanism of twinfilin-capping protein interaction. EMBO J, 23: 3010-3019. Fambrough D, McClure K, Kazlauskas A and Lander ES (1999). Diverse signaling pathways activated by growth factor receptors induce broadly overlapping, rather than independent, sets of genes. Cell, 97: 727-741. Fang ZY, Prins JB and Marwick TH (2004). Diabetic cardiomyopathy: evidence, mechanisms, and therapeutic implications. Endocr Rev, 25: 543-567. Fatkin D, McConnell BK, Mudd JO, Semsarian C, Moskowitz IG, Schoen FJ, Giewat M, Seidman CE and Seidman JG (2000). An abnormal Ca(2+) response in mutant sarcomere protein-mediated familial hypertrophic cardiomyopathy. J Clin Invest, 106: 1351-1359. Fedida D, Braun AP and Giles WR (1993). Alpha1-Adrenoceptors in myocardium: functional aspects and transmembrane signaling mechanisms. Physiol Rev, 73: 469–487. Fedorova OV, Talan MI, Agalakova NI, Lakatta EG and Bagrov AY (2004). Coordinated shifts in Na/K-ATPase isoforms and their endogenous ligands during cardiac hypertrophy and failure in NaCl-sensitive hypertension. J Hypertens, 22: 389-397. Feldman AM, Weinberg EO, Ray PE and Lorell BH (1993). Selective changes in cardiac gene expression during compensated hypertrophy and the transition to cardiac decompensation in rats with chronic aortic banding. Circ Res, 73: 184-192. Feng YH and Douglas JG (2001). Angiotensin receptors: an overview. In: Epstein,M and Brunner, HR editors, ‘Angiotensin II receptor antagonists’, Philadelphia: Hanley & Belfus, Inc., pp. 29-48.

Bibliography

299

Feng YH, Sun Y and Douglas JG (2002). Gbeta gamma -independent constitutive association of Galpha s with SHP-1 and angiotensin II receptor AT2 is essential in AT2-mediated ITIM-independent activation of SHP-1. Proc Natl Acad Sci U S A, 99: 12049-12054. Feolde E, Vigne P and Frelin C (1993). Angiotensin AT1 receptors mediate a positive inotropic effect of angiotensin II in guinea pig atria. Eur J Pharmacol, 245: 63-66. Ferdinandusse S, Finckh B, de Hingh YC, Stroomer LE, Denis S, Kohlschutter A and Wanders RJ (2003). Evidence for increased oxidative stress in peroxisomal D-bifunctional protein deficiency. Mol Genet Metab, 79: 281-287. Ferencz C, Rubin JD, McCarter RJ, Brenner JI, Neill CA, Perry LW, Hepner SI and Downing JW (1985). Congenital heart disease: prevalence at livebirth. The Baltimore-Washington Infant Study. Am J Epidemiol, 121: 31-36. Fernandes L, Fortes ZB, Nigro D, Tostes RCA, Santos RAS and Carvalho MHC (2001). Potentiation of bradykinin by angiotensin-(1-7) on arterioles of spontaneously hypertensive rats studied in vivo. Hypertension, 37: 703-709. Fernandes R, Girao H and Pereira P (2004). High glucose down-regulates intercellular communication in retinal endothelial cells by enhancing degradation of connexin 43 by a proteasome-dependent mechanism. J Biol Chem, 279: 27219-27224. Fernandez-Real JM and Ricart W (2003). Insulin resistance and chronic cardiovascular inflammatory syndrome. Endocr Rev, 24: 278-301. Fernandez-Velasco M, Goren N, Benito G, Blanco-Rivero J, Bosca L and Delgado C (2003). Regional distribution of hyperpolarization-activated current (If) and hyperpolarization-activated cyclic nucleotide-gated channel mRNA expression in ventricular cells from control and hypertrophied rat hearts. J Physiol, 553: 395-405. Ferrannini E, Vichi S, Bech-Nielsen H, Laakso M, Paolisso G and Smith U (1996). Insulin action and age. Diabetes, 45: 947-953. Ferron L, Capuano V, Ruchon Y, Deroubaix E, Coulombe A and Renaud JF (2003). Angiotensin II signaling pathways mediate expression of cardiac T-type calcium channels. Circ Res, 93: 1241-1248. Fesik SW (2000). Insights into programmed cell death through structural biology. Cell, 103: 273-282. Filipek A, Jastrzebska B, Nowotny M and Kuznicki J (2002). CacyBP/SIP, a calcyclin and Siah-1-interacting protein, binds EF-hand proteins of the S100 family. J Biol Chem, 277: 28848-28852. Filipek A, Wojda U and Lesniak W (1995). Interaction of calcyclin and its cyanogen bromide fragments with annexin II and glyceraldehyde 3-phosphate dehydrogenase. Int J Biochem Cell Biol, 27:1123-1131.

Bibliography

300

Filippatos GS, Tsilias K, Venetsanou K, Karambinos E, Manolatos D, Kranidis A, Antonellis J, Kardaras F, Anthopoulos L and Baltopoulos G (2000). Leptin serum levels in cachectic heart failure patients. Relationship with tumor necrosis factor-alpha system. Int J Cardiol, 76: 117-122. Fiordaliso F, Leri A, Cesselli D, Limana F, Safai B, Nadal-Ginard B, Anversa P and Kajstura J (2001). Hyperglycemia activates p53 and p53-regulated genes leading to myocyte cell death. Diabetes, 50: 2363-2375. Fischer Y, Bottcher U, Eblenkamp M, Thomas J, Jungling E, Rosen P and Kammermeier H (1997a). Glucose transport and glucose transporter GLUT4 are regulated by product(s) of intermediary metabolism in cardiomyocytes. Biochem J, 321: 629-638. Fischer Y, Thomas J, Sevilla L, Munoz P, Becker C, Holman G, Kozka IJ, Palacin M, Testar X, Kammermeier H and Zorzano A (1997b). Insulin-induced recruitment of glucose transporter 4 (GLUT4) and GLUT1 in isolated rat myocytes. J Biol Chem, 272: 7085-7092. Fitzsimons JT (1998). Angiotensin, thirst, and sodium appetite. Physiol Rev, 78: 583-686. Flesch M, Schwinger RH, Schiffer F, Frank K, Sudkamp M, Kuhn-Regnier F, Arnold G and Bohm M (1996). Evidence for functional relevance of an enhanced expression of the Na(+)-Ca2+ exchanger in failing human myocardium. Circulation, 94: 992-1002. Flesch M, Schwinger RH, Schnabel P, Schiffer F, van Gelder I, Bavendiek U, Sudkamp M, Kuhn-Regnier F and Bohm M (1997). Sarcoplasmic reticulum Ca2+ATPase and phospholamban mRNA and protein levels in end-stage heart failure due to ischemic or dilated cardiomyopathy. J Mol Med, 74: 321-332. Fliss H and Gattinger D (1996). Apoptosis in ischemic and reperfused rat myocardium. Circ Res, 79: 949-956. Flucher BE and Franzini-Armstrong C (1996). Formation of junctions involved in excitation-contraction coupling in skeletal and cardiac muscle. Proc Natl Acad Sci U S A, 93: 8101-8106. Fournier A, Fantini E, Sergiel JP, Athias P and Grynberg A (1995). Influence of the phospholipid content in docosahexaenoic acid on electrophysiology and contraction of rat heart muscle cells. Cardioscience, 6: 71-78. Frangogiannis NG, Michael LH and Entman ML (2000). Myofibroblasts in reperfused myocardial infarcts express the embryonic form of smooth muscle myosin heavy chain (SMemb). Cardiovasc Res, 48: 89-100. Franzini-Armstrong C and Jorgensen AO (1994). Structure and development of E-C coupling units in skeletal muscle. Annu Rev Physiol, 56: 509-534.

Bibliography

301

Frasch M (1995). Induction of visceral and cardiac mesoderm by ectodermal Dpp in the early Drosophila embryo. Nature, 374: 464-467. Freeman GL, Little WC and O’Rourke RA (1987). Influence of heart rate on left ventricular performance in conscious dogs. Circ Res, 61: 455-464. Freeman WM, Walker SJ and Vrana KE (1999). Quantitative RT-PCT: pitfalls and potential. Biotechniques, 26:112-125. Frey N and Olson EN (2003). Cardiac hypertrophy: the good, the bad, and the ugly. Annu Rev Physiol, 65: 45-79. Friddle CJ, Koga T, Rubin EM and Bristow J (2000). Expression profiling reveals distinct sets of genes altered during induction and regression of cardiac hypertrophy. Proc Natl Acad Sci U S A, 97: 6745-6750. Fukada SY, Tirapelli CR, de Godoy MA and de Oliveira AM (2005). Mechanisms underlying the endothelium-independent relaxation induced by angiotensin II in rat aorta. J Cardiovasc Pharmacol, 45: 136-143. Furman C, Short SM, Subramanian RR, Zetter BR and Roberts TM (2002). DEF-1/ASAP1 is a GTPase-activating protein (GAP) for ARF1 that enhances cell motility through a GAP-dependent mechanism. J Biol Chem, 277: 7962-7969. Gallinat S, Csikos T, Meffert S, Herdegen T, Stoll M and Unger T (1997). The angiotensin AT2 receptor down-regulates neurofilament M in PC12W cells. Neurosci Lett, 227: 29-32. Galderisi M, Anderson KM, Wilson PWF and Levy D (1991). Echocardiographic evidence for the existence of a distinct diabetic cardiomyopathy (the Framingham Heart Study). Am J Cardiol, 68: 85-89. Gan XT, Chakrabarti S and Karmazyn M (1999). Modulation of Na+/H+ exchange isoform 1 mRNA expression in isolated rat hearts. Am J Physiol, 277: H993-H998. Gao WD, Perez NG and Marban E (1998). Calcium cycling and contractile activation in intact mouse cardiac muscle. J Physiol, 507: 175-184. Garber DW and Neely JR (1983). Decreased myocardial function and myosin ATPase in hearts from diabetic rats. Am J Physiol, 244: H586-H591. Garcia-Martinez V and Schoenwolf GC (1993). Primitive-streak origin of the cardiovascular system in avian embryos. Dev Biol, 159: 706-719. Garvey WT, Hardin T, Juhaszova M and Dominguez JH (1993). Effects of diabetes on myocardial glucose transport system in rats: implications for diabetic cardiomyopathy. Am J Physiol, 264: H837-H844.

Bibliography

302

Geiman TM, Sankpal UT, Robertson AK, Chen Y, Mazumdar M, Heale JT, Schmiesing JA, Kim W, Yokomori K, Zhao Y and Robertson KD (2004). Isolation and characterization of a novel DNA methyltransferase complex linking DNMT3B with components of the mitotic chromosome condensation machinery. Nucleic Acids Res, 32: 2716-2729. Georgakopoulos D and Kass DA (2001). Minimal force-frequency modulation of inotropy and relaxation of in situ murine heart. J Physiol, 534: 535-545. Gerdts E, Oikarinen L, Palmieri V, Otterstad JE, Wachtell K, Boman K, Dahlof B and Devereux RB (2002). Correlates of left atrial size in hypertensive patients with left ventricular hypertrophy: the Losartan Intervention For Endpoint Reduction in Hypertension (LIFE) Study. Hypertension, 39: 739-743. Giancotti FG and Ruoslahti E (1999). Integrin signaling. Science, 285: 1028–1032. Gilbert RE, Rumble JR, Cao Z, Cox AJ, van Eeden P, Allen TJ, Kelly DJ and Cooper ME (2000). Endothelin receptor antagonism ameliorates mast cell infiltration, vascular hypertrophy, and epidermal growth factor expression in experimental diabetes. Circ Res, 86: 158-165. Glatz JF, van Breda E, Keizer HA, de Jong YF, Lakey JR, Rajotte RV, Thompson A, van der Vusse GJ and Lopaschuk GD (1994). Rat heart fatty acid-binding protein content is increased in experimental diabetes. Biochem Biophys Res Commun, 199: 639-646. Go LO, Moschella MC, Watras J, Handa KK, Fyfe BS and Marks AR (1995). Differential regulation of two types of intracellular calcium release channels during end-stage heart failure. J Clin Invest, 95: 888-894. Golfman L, Dixon IM, Takeda N, Lukas A, Dakshinamurti K and Dhalla NS (1998). Cardiac sarcolemmal Na(+)-Ca2+ exchange and Na(+)-K+ ATPase activities and gene expression in alloxan-induced diabetes in rats. Mol Cell Biochem, 188: 91-101. Goto T, Takase H, Toriyama T, Sugiura T, Sato K, Ueda R and Dohi Y (2003). Circulating concentrations of cardiac proteins indicate the severity of congestive heart failure. Heart, 89: 1303-1307. Goodwin GW, Taylor CS and Taegtmeyer H (1998). Regulation of energy metabolism of the heart during acute increase in heart work. J Biol Chem, 273: 29530-29539. Goodwin GW and Taegtmeyer H (2000). Improved energy homeostasis of the heart in the metabolic state of exercise. Am J Physiol Heart Circ Physiol, 279: H1490-H1501. Gordon JR (2000). TGFbeta1 and TNFalpha secreted by mast cells stimulated via the FcepsilonRI activate fibroblasts for high-level production of monocyte chemoattractant protein-1 (MCP-1). Cell Immunol, 201: 42-49. Gottlieb RA, Burleson KO, Kloner RA, Babior BM and Engler RL (1994). Reperfusion injury induces apoptosis in rabbit cardiomyocytes. J Clin Invest, 94: 1621-1628.

Bibliography

303

Grace AA, Metcalfe JC, Weissberg PL, Bethell HW and Vandenberg JI (1996). Angiotensin II stimulates sodium-dependent proton extrusion in perfused ferret heart. Am J Physiol, 270: C1687-C1694. Graf K, Do YS, Ashizawa N, Meehan WP, Giachelli CM, Marboe CC, Fleck E and Hsueh WA (1997). Myocardial osteopontin expression is associated with left ventricular hypertrophy. Circulation, 96: 3063-3071. Grinstein S, Rotin D and Mason MJ (1989). Na+/H+ exchange and growth factor-induced cytosolic pH changes. Role in cellular proliferation. Biochim Biophys Acta, 988: 73-97. Gross V, Obst M, Kiss E, Janke J, Mazak I, Shagdarsuren E, Muller DN, Langenickel TH, Grone HJ and Luft FC (2004). Cardiac hypertrophy and fibrosis in chronic L-NAME-treated AT2 receptor-deficient mice. J Hypertens, 22: 997-1005. Grundy SM, Benjamin IJ, Burke GL, Chait A, Eckel RH, Howard BV, Mitch W, Smith SC Jr and Sowers JR (1999). Diabetes and cardiovascular disease: a statement for healthcare professionals from the American Heart Association. Circulation, 100: 1134-1146. Grünig E, Tasman JA, Kücherer H, Franz W, Kübler W and Katus HA (1998). Frequency and phenotypes of familial dilated cardiomyopathy. J Am Coll Cardiol, 31: 186–194. Gu JW, Anand V, Shek EW, Moore MC, Brady AL, Kelly WC and Adair TH (1998). Sodium induces hypertrophy of cultured myocardial myoblasts and vascular smooth muscle cells. Hypertension, 31: 1083-1087. Guimaraes S, Pinheiro (2005). Functional evidence that in the cardiovascular system AT1 angiotensin II receptors are AT1b prejunctionally and AT1A postjunctionally. Cardiovasc Res, 67: 208-215. Gunja-Smith Z, Morales AR, Romanelli R and Woessner Jr. JF (1996). Remodeling of human myocardial collagen in idiopathic dilated cardiomyopathy. Role of metalloproteinases and pyridinoline cross-links. Am J Pathol, 148: 1639-1648. Guo X, Chapman D and Dhalla NS (2003). Partial prevention of changes in SR gene expression in congestive heart failure due to myocardial infarction by enalapril or losartan. Mol Cell Biochem, 254: 163-172. Gustafsson F and Holstein-Rathlou NH (1999). Angiotensin II modulates conducted vasoconstriction to norepinephrine and local electrical stimulation in rat mesenteric arterioles. Cardiovasc Res, 44: 176-184. Gutstein DE, Morley GE, Tamaddon H, Vaidya D, Schneider MD, Chen J, Chien KR, Stuhlmann H and Fishman GI (2001). Conduction slowing and sudden arrhythmic death in mice with cardiac-restricted inactivation of connexin43. Circ Res, 88: 333-339.

Bibliography

304

Guy LG, Kothary R and Wall L (1997). Position effects in mice carrying a lacZ transgene in cis with the beta-globin LCR can be explained by a graded model. Nucleic Acids Res, 25: 4400-4407. Guyton AC and Hall JE (1996). Renin-angiotensin system: its role in pressure control and hypertension. In: Saunders, WB, editor, ‘Textbook of medical physiology’, Philadelphia, pp. 227-233. Gwathmey JK, Slawsky MT, Hajjar RJ, Briggs GM and Morgan JP (1990). Role of intracellular calcium handling in force-interval relationships of human ventricular myocardium. J Clin Invest, 85: 1599-1613. Hacia JG (1999). Resequencing and mutational analysis using oligonucleotide microarrays. Nat Genet, 21 (Suppl 1): 42-47. Hammerer-Lercher A, Mair J, Bonatti J, Watzka SB, Puschendorf B and Dirnhofer S (2001). Hypoxia induces heat shock protein expression in human coronary artery bypass grafts. Cardiovasc Res, 50: 115-124. Hao L, Du M, Lopez-Campistrous A and Fernandez-Patron C (2004). Agonist-induced activation of matrix metalloproteinase-7 promotes vasoconstriction through the epidermal growth factor-receptor pathway. Circ Res, 94: 68-76. Haq S, Michael A, Andreucci M, Bhattacharya K, Dotto P, Walters B, Woodgett J, Kilter H and Force T (2003). Stabilization of beta-catenin by a Wnt-independent mechanism regulates cardiomyocyte growth. Proc Natl Acad Sci U S A, 100: 4610-4615. Harada K, Komuro I, Shiojima I, Hayashi D, Kudoh S, Mizuno T, Kijima K, Matsubara H, Sugaya T, Murakami K and Yazaki Y (1998). Pressure overload induces cardiac hypertrophy in angiotensin II type 1A receptor knockout mice. Circulation, 97: 1952-1959. Harada K, Komuro I, Zou Y, Kudoh S, Kijima K, Matsubara H, Sugaya T, Murakami K and Yazaki Y (1998). Acute pressure overload could induce hypertrophic responses in the heart of angiotensin II type 1a knockout mice. Circ Res, 82: 779-785. Harada K, Sugaya T, Murakami K, Yazaki Y and Komuro I (1999). Angiotensin II type 1A receptor knockout mice display less left ventricular remodeling and improved survival after myocardial infarction. Circulation, 100: 2093-2099. Harada M, Saito Y, Nakagawa O, Miyamoto Y, Ishikawa M, Kuwahara K, Ogawa E, Nakayama M, Kamitani S, Hamanaka I, Kajiyama N, Masuda I, Itoh H, Tanaka I and Nakao K (1997). Role of cardiac nonmyocytes in cyclic mechanical stretch-induced myocyte hypertrophy. Heart Vessels, 12 (Suppl):198-200. Harris PJ, Steward D, Cullinan MC, Delbridger LM, Dally L and Grinwald P (1987). Rapid measurement of isolated cardiac muscle cell length using a line-scan camera. IEEE Trans Biomed Eng, 34: 463-467.

Bibliography

305

Hart G (1994). Cellular electrophysiology in cardiac hypertrophy and failure. Cardiovasc Res, 28: 933-946. Hartzell HC and Duchatelle-Gourdon I (1997). Regulation of the cardiac delayed rectifier K current by neurotransmitters and magnesium. Cardiovasc Drugs Ther, 7 (Suppl 3): 547-554. Hasenfuss G and Pieske B (2002). Calcium cycling in congestive heart failure. J Mol Cell Cardiol, 34: 951-969. Hasenfuss G, Meyer M, Schillinger W, Preuss M, Pieske B and Just H (1997). Calcium handling proteins in the failing human heart. Basic Res Cardiol, 92 (Suppl 1): 87-93. Hasenfuss G, Mulieri LA, Blanchard EM, Holubarsch C, Leavitt BJ, Ittleman F and Alpert NR (1991). Energetics of isometric force development in control and volume-overload human myocardium. Comparison with animal species. Circ Res, 68: 836-846. Hasenfuss G, Reinecke H, Studer R, Pieske B, Meyer M, Drexler H and Just H (1996). Calcium cycling proteins and force-frequency relationship in heart failure. Basic Res Cardiol, 91 (Suppl 2): 17-22. Hasenfuss G, Schillinger W, Lehnart SE, Preuss M, Pieske B, Maier LS, Prestle J, Minami K and Just H (1999). Relationship between Na+-Ca2+-exchanger protein levels and diastolic function of failing human myocardium. Circulation, 99: 641-648. Hashida H, Hamada M and Hiwada K (1999). Serial changes in sarcoplasmic reticulum gene expression in volume-overloaded cardiac hypertrophy in the rat: effect of an angiotensin II receptor antagonist. Clin Sci, 96: 387-395. Hattori Y, Matsuda N, Kimura J, Ishitani T, Tamada A, Gando S, Kemmotsu O and Kanno M (2000). Diminished function and expression of the cardiac Na+-Ca2+ exchanger in diabetic rats: implication in Ca2+ overload. J Physiol, 527: 85-94. Haworth RS, Yasutake M, Brooks G and Avkiran M (1997). Cardiac Na+-H+ exchanger during postnatal development in the rat: changes in mRNA expression and sarcolemmal activity. J Mol Cell Cardiol, 29: 321-332. Hayasaki-Kajiwara Y, Kitano Y, Iwasaki T, Shimamura T, Naya N, Iwaki K and Nakajima M (1999). Na(+)influx via Na(+)/H(+)exchange activates protein kinase C isozymes delta and epsilon in cultured neonatal rat cardiac myocytes. J Mol Cell Cardiol, 31: 1559-1572. Hayat SA, Patel B, Khattar RS and Malik RA (2004). Diabetic cardiomyopathy: mechanisms, diagnosis and treatment. Clin Sci (Lond), 107: 539-557. He J, Conklin MW, Foell JD, Wolff MR, Haworth RA, Coronado R and Kamp TJ (2001). Reduction in density of transverse tubules and L-type Ca(2+) channels in canine tachycardia-induced heart failure. Cardiovasc Res, 49: 298-307.

Bibliography

306

Hein L, Dzau VJ and Barsh GS (2005). Linkage mapping of the angiotensin AT2 receptor gene (Agtr2) to the mouse X chromosome. Genomics, 30: 369-371. Hein L, Stevens ME, Barsh GS, Pratt RE, Kobilka BK and Dzau VJ (1997). Overexpression of angiotensin AT1 receptor transgene in the mouse myocardium produces a lethal phenotype associated with myocyte hyperplasia and heart block. Proc Natl Acad Sci U S A, 94: 6391-6396. Henegariu O, Heerema NA, Dlouhy SR, Vance GH and Vogt PH (1997). Multiplex PCR: critical parameters and step-by-step protocol. Biotechniques, 23:504-511. Hileeto D, Cukiernik M, Mukherjee S, Evans T, Barbin Y, Downey D, Karmazyn M and Chakrabarti S (2002). Contributions of endothelin-1 and sodium hydrogen exchanger-1 in the diabetic myocardium. Diabetes Metab Res Rev, 18: 386-394. Hittinger L, Ghaleh B, Chen J, Edwards JG, Kudej RK, Iwase M, Ki SJ, Vatner SF and Vatner DE (1999). Reduced subendocardial ryanodine receptors and consequent effects on cardiac function in conscious dogs with left ventricular hypertrophy. Circ Res, 84: 999-1006. Hiyoshi H, Yayama K, Takano M and Okamoto H (2005). Angiotensin type 2 receptor-mediated phosphorylation of eNOS in the aortas of mice with 2-kidney, 1-clip hypertension. Hypertension, 45: 967-973. Hobai IA and O’Rourke B (2000). Enhanced Ca2+-activated Na+-Ca2+ exchange activity in canine pacing-induced heart failure. Circ Res, 87: 690-698. Hochstrasser M (1996). Protein degradation or regulation: Ub the judge. Cell, 84: 813-815. Hofmann PA, Menon V and Gannaway KF (1995). Effects of diabetes on isometric tension as a function of [Ca2+] and pH in rat skinned cardiac myocytes. Am J Physiol, 269: H1656-H1663. Hoffmann R, Seidl T and Dugas M (2002). Profound effect of normalization on detection of differentially expressed genes in oligonucleotide microarray data analysis. Genome Biol, 3: RESEARCH0033. Hoffmann S, Krause T, van Geel PP, Willenbrock R, Pagel I, Pinto YM, Buikema H, van Gilst WH, Lindschau C, Paul M, Inagami T, Ganten D and Urata H (2001). Overexpression of the human angiotensin II type 1 receptor in the rat heart augments load induced cardiac hypertrophy. J Mol Med, 79: 601-608. Hoit BD, Ball N and Walsh RA (1997). Invasive hemodynamics and force-frequency relationships in open- versus closed-chest mice. Am J Physiol, 273: H2528-H2533. Honen BN, Saint DA and Laver DR (2003). Suppression of calcium sparks in rat ventricular myocytes and direct inhibition of sheep cardiac RyR channels by EPA, DHA and oleic acid. J Membr Biol, 196: 95-103.

Bibliography

307

Hong-Brown LQ and Deschepper CF (1992). Effects of thyroid hormones on angiotensinogen gene expression in rat liver, brain and cultured cells. Endocrinology, 130: 1231-1237. Hoppe UC, Jansen E, Südkamp M and Beuckelmann DJ (1998). A hyperpolarization-activated inward current (If) in ventricular myocytes from normal and failing human hearts. Circulation, 97: 55–65. Horackova M and Armour JA (1997). Ang II modifies cardiomyocyte function via extracardiac and intracardiac neurons: in situ and in vitro studies. Am J Physiol, 272: R766-R775. Hori M, Nakatsubo N, Kagiya T, Iwai K, Sato H, Iwakura K, Kitabatake A and Kamada T (1990). The role of Na+/H+ exchange in norepinephrine-induced protein synthesis in neonatal cultured rat cardiomyocytes. Jpn Circ J, 54: 535-539. Horinaka S, Kobayashi N, Mori Y, Yagi H, Onoda M, Matsuoka H (2003). Expression of inducible nitric oxide synthase, left ventricular function and remodeling in Dahl salt-sensitive hypertensive rats. Int J Cardiol, 91: 25-35. Hotamisligil GS and Spiegelman BM (1994). Tumor necrosis factor-(alpha): a key component of the obesity-diabetes link. Diabetes, 43: 1271-1278. Houiller P, Chambrey R, Achard JM, Froissart M, Poggioli J and Paillard M (1996). Signaling pathways in the biphasic effect of angiotensin II on apical Na/H antiport activity in proximal tubule. Kidney Int, 50: 1496-1505. Hsieh JC, Lee L, Zhang L, Wefer S, Brown K, DeRossi C, Wines ME, Rosenquist T and Holdener BC (2003). Mesd encodes an LRP5/6 chaperone essential for specification of mouse embryonic polarity. Cell, 112: 355-367. Hsueh WA, Law RE and Do YS (1998). Integrins, adhesion, and cardiac remodeling. Hypertension, 31: 176-180. Hu E, Kim JB, Sarraf P and Spiegelman BM (1996). Inhibition of adipogenesis through MAP kinase-mediated phosphorylation of PPARγ. Science, 274: 2100-2103. Hu J and Cotgreave IA (1997). Differential regulation of gap junctions by proinflammatory mediators in vitro. J Clin Invest, 99: 2312-2316. Huggins CE, Domenighetti AA, Pedrazzini T, Pepe S and Delbridge LM (2003). Elevated intracardiac angiotensin II leads to cardiac hypertrophy and mechanical dysfunction in normotensive mice. J Renin Angiotensin Aldosterone Syst, 4: 186-190. Huggins CE, Kalil N, Proietto J, Pepe S, Delbridge LMD (2004). Pyruvate supplementation ameliorates ex vivo cardiac dysfunction in GLUT4 deficient mice. J Mol Cell Cardiol, 37: C79 (abstract).

Bibliography

308

Hughes TR, Mao M, Jones AR, Burchard J, Marton MJ, Shannon KW, Lefkowitz SM, Ziman M, Schelter JM, Meyer MR, Kobayashi S, Davis C, Dai H, He YD, Stephaniants SB, Cavet G, Walker WL, West A, Coffey E, Shoemaker DD, Stoughton R, Blanchard AP, Friend SH and Linsley PS (2001). Expression profiling using microarrays fabricated by an ink-jet oligonucleotide synthesizer. Nat Biotechnol, 19: 342-347. Hung CC, Ichimura T, Stevens JL and Bonventre JV (2003). Protection of renal epithelial cells against oxidative injury by endoplasmic reticulum stress preconditioning is mediated by ERK1/2 activation. J Biol Chem, 278: 29317-29326. Hunter JJ and Chien KR (1999). Signaling pathways for cardiac hypertrophy and failure. New Engl J Med, 341:1276-1283. Hwang JJ, Allen PD, Tseng GC, Lam CW, Fananapazir L, Dzau VJ and Liew CC (2002). Microarray gene expression profiles in dilated and hypertrophic cardiomyopathic end-stage heart failure. Physiol Genomics, 10: 31-44. Hynes RO (1999). Cell adhesion: old and new questions. Trends Cell Biol, 9: M33–M37. Ichiki T, Labosky PA, Shiota C, Okuyama S, Imagawa Y, Fogo A, Niimura F, Ichikawa I, Hogan BL and Inagami T (1995). Effects on blood pressure and exploratory behaviour of mice lacking angiotensin II type-2 receptor. Nature, 377: 748-750. Ichiyanagi O, Ishii K and Endoh M (2002). Angiotensin II increases L-type Ca2+ current in gramicidin D-perforated adult rabbit ventricular myocytes: comparison with conventional patch-clamp method. Pflugers Arch, 444: 107-116. Ikenouchi H, Barry WH, Bridge JH, Weinberg EO, Apstein CS and Lorell BH (1994). Effects of angiotensin II in intracellular Ca2+ and pH in isolated beating rabbit hearts and myocytes loaded with indicator indo-1. J Physiol, 480: 203-215. Imahashi K, Hashimoto K, Yamaguchi H, Nishimura T and Kusuoka H (1998). Alteration of intracellular Na+ during ischemia in diabetic rat hearts: the role of reduced activity in Na+/H+ exchange against stunning. J Mol Cell Cardiol, 30: 509-517. Imanaka-Yoshida K, Amitani A, Ioshii SO, Koyabu S, Yamakado T and Yoshida T (1996). Alterations of expression and distribution of the Ca(2+)-storing proteins in endo/sarcoplasmic reticulum during differentiation of rat cardiomyocytes. J Mol Cell Cardiol, 28: 553-562. Imanishi K, Nonoguchi H, Nakayama Y, Machida K, Ikebe M and Tomita K (2003). Type 1A angiotensin II receptor is regulated differently in proximal and distal nephron segments. Hypertens Res, 26: 405-411. Imperatore G, Pinsky LE, Motulsky A, Reyes M, Bradley LA and Burke W (2003). Hereditary hemochromatosis: perspectives of public health, medical genetics, and primary care. Genet Med, 5: 1-8.

Bibliography

309

Inoguchi T, Yu HY, Imamura M, Kakimoto M, Kuroki T, Maruyama T and Nawata H (2001). Altered gap junction activity in cardiovascular tissues of diabetes. Med Electron Microsc, 34: 86-91. Ishihata A and Endoh M (1995). Species-related differences in inotropic effects of angiotensin II in mammalian ventricular muscle: receptors, subtypes and phosphoinositide hydrolysis. Br J Pharmacol, 114: 447-453. Ishikawa T, Kajiwara H and Kurihara S (1999). Alterations in contractile properties and Ca2+ handling in streptozotocin-induced diabetic rat myocardium. Am J Physiol, 277: H2185-H2194. Ito H, Hiroe M, Hirata Y, Fujisaki H, Adachi S, Akimoto H, Ohta Y, and Marumo F (1994). Endothelin ETA receptor antagonist blocks cardiac hypertrophy provoked by hemodynamic overload. Circulation, 89: 2198-2203. Ito N, Kagaya Y, Weinberg EO, Barry WH and Lorell BH (1997). Endothelin and angiotensin II stimulation of Na+-H+ exchange is impaired in cardiac hypertrophy. J Clin Invest, 99: 125-135. Ito Y, Suko J and Chidsey CA (1974). Intracellular calcium and myocardial contractility. V. Calcium uptake of sarcoplasmic reticulum fractions in hypertrophied and failing rabbit hearts. J Mol Cell Cardiol, 6: 237-247. Iwai N, Shimoike H and Kinoshita M (1995). Cardiac renin-angiotensin system in the hypertrophied heart. Circulation, 92: 2690-2696. Iwanaga Y, Aoyama T, Kihara Y, Onozawa Y, Yoneda T and Sasayama S (2002). Excessive activation of matrix metalloproteinases coincides with left ventricular remodeling during transition from hypertrophy to heart failure in hypertensive rats. J Am Coll Cardiol, 39: 1384-1391. Janczewski AM, Zahid M, Lemster BH, Frye CS, Gibson G, Higuchi Y, Kranias EG, Feldman AM and McTiernan CF (2004). Phospholamban gene ablation improves calcium transients but not cardiac function in a heart failure model. Cardiovasc Res, 62: 468-480. Janicki JS, Brower GL, Gardner JD, Chancey AL and Stewart JA Jr (2004). The dynamic interaction between matrix metalloproteinase activity and adverse myocardial remodeling. Heart Fail Rev, 9: 33-42. Janse MJ (2004). Electrophysiological changes in heart failure and their relationship to arrhythmogenesis. Cardiovasc Res, 61: 208-217. Jayaraman T and Marks AR (2000). Calcineurin is downstream of the inositol 1,4,5-trisphosphate receptor in the apoptotic and cell growth pathways. J Biol Chem, 275: 6417-6420.

Bibliography

310

Jethmalani SM and Henle KJ (1998). Calreticulin associates with stress proteins: implications for chaperone function during heat stress. J Cell Biochem, 69: 30-43. Jeunemaitre X, Soubrier F, Kotelevtsev YV, Lifton RP, Williams CS, Charru A, Hunt SC, Hopkins PN, Williams RR, Lalouel JM and Corvol P (1992). Molecular basis of human hypertension: role of angiotensinogen. Cell, 71: 7-20. Jortani SA and Valdes Jr R (2001). Mammalian cardenolides as biomarkers in congestive heart failure. Cardiovasc Toxicol, 1: 165-170. Jost-Vu E, Horton R and Antonipillai I (1992). Altered regulation of renin secretion by insulinlike growth factors and angiotensin II in diabetic rats. Diabetes, 41: 1100-1105. Jourdon P and Feuvray D (1993). Calcium and potassium currents in ventricular myocytes isolated from diabetic hearts. J Physiol, 470: 411-429. Ju H, Scammel-La Fleur T and Dixon IM (1996). Altered mRNA abundance of calcium transport genes in cardiac myocytes induced by angiotensin II. J Mol Cell Cardiol, 28: 1119-1128. Jude S, Bedut S, Roger S, Pinault M, Champeroux P, White E and Le Guennec JY (2003). Peroxidation of docosahexaenoic acid is responsible for its effects on I TO and I SS in rat ventricular myocytes. Br J Pharmacol, 139: 816-822. Julien J (1997). Cardiac complications in non-insulin-dependent diabetes mellitus. J Diabetes Complications, 11: 123-130. Kaab S, Nuss HB, Chiamvimonvat N, O'Rourke B, Pak PH, Kass DA, Marban E and Tomaselli GF (1996). Ionic mechanism of action potential prolongation in ventricular myocytes from dogs with pacing-induced heart failure. Circ Res, 78: 262-273. Kaczmarczyk SJ, Andrikopoulos S, Favaloro J, Domenighetti AA, Dunn A, Ernst M, Grail D, Fodero-Tavoletti M, Huggins CE, Delbridge LM, Zajac JD and Proietto J (2003). Threshold effects of glucose transporter-4 (GLUT4) deficiency on cardiac glucose uptake and development of hypertrophy. J Mol Endocrinol, 31: 449-459. Kadambi VJ, Ball N, Kranias EG, Walsh RA and Hoit BD (1999). Modulation of force-frequency relation by phospholamban in genetically engineered mice. Am J Physiol, 276: H2245-H2250. Kagaya Y, Hajjar RJ, Gwathmey JK, Barry WH and Lorell BH (1996). Long-term angiotensin-converting enzyme inhibition with fosinopril improves depressed responsiveness to Ca2+ in myocytes from aortic-banded rats. Circulation, 94: 2915-2922. Kahn BB (1996). Glucose transport: pivotal step in insulin action. Diabetes, 45: 1644-1654.

Bibliography

311

Kahn BB (2000). Targeted disruption of the glucose transporter 4 selectively in muscle causes insulin resistance and glucose intolerance. Nat Med, 6: 924-928. Kaibara M, Mitarai S, Yano K and Kameyama M (1994). Involvement of Na+-H+ antiporter in regulation of L-type Ca2+ channel current by angiotensin II in rabbit ventricular myocytes. Circ Res, 75: 1121-1125. Kajstura J, Fiordaliso F, Andreoli AM, Li B, Chimenti S, Medow MS, Limana F, Nadal-Ginard B, Leri A and Anversa P (2001). IGF-1 overexpression inhibits the development of diabetic cardiomyopathy and angiotensin II-mediated oxidative stress. Diabetes, 50: 1414-1424. Kambayashi Y, Bardhan S, Takahashi K, Tsuzuki S, Inui H, Hamakubo T and Inagami T (1993). Molecular cloning of a novel angiotensin II receptor isoform involved in phosphotyrosine phosphatase inhibition. J Biol Chem, 268: 24543-24546. Kannel WB and McGee DL (1979). Diabetes and cardiovascular disease: the Framingham study. JAMA, 241: 2035-2038. Kannel WB and Belanger AJ (1991). Epidemiology of heart failure. Am Heart J, 121: 951. Kang N, Walther T, Tian XL, Bohlender J, Fukamizu A, Ganten D and Bader M (2002). Reduced hypertension-induced end-organ damage in mice lacking cardiac and renal angiotensinogen synthesis. J Mol Med, 80: 359-366. Kapus A, Grinstein S, Wasan S, Kandasamy R and Orlowski J (1994). Functional characterization of three isoforms of the Na+/H+ exchanger stably expressed in Chinese hamster ovary cells. ATP dependence, osmotic sensitivity, and role in cell proliferation. J Biol Chem, 269: 23544-23552. Karmazyn M, Gan XT, Humphreys RA, Yoshida H and Kusumoto K (1999). The myocardial Na+/H+ exchange: structure, regulation, and its role in heart disease. Circ Res, 85: 777-786. Kasahara H, Ueyama T, Wakimoto H, Liu MK, Maguire CT, Converso KL, Kang PM, Manning WJ, Lawitts J, Paul DL, Berul CI and Izumo S (2003). Nkx2.5 homeoprotein regulates expression of gap junction protein connexin 43 and sarcomere organization in postnatal cardiomyocytes. J Mol Cell Cardiol, 35: 243-256. Kashihara H, Shi ZQ, Yu JZ, McNeill JH and Tibbits GF (2000). Effects of diabetes and hypertension on myocardial Na+-Ca2+ exchange. Can J Physiol Pharmacol, 78: 12-19. Kasper EK, Agema WR, Hutchins GM, Deckers JW, Hare JM and Baughman KL (1994). The causes of dilated cardiomyopathy: a clinicopathologic review of 673 consecutive patients. J Am Coll Cardiol, 23: 586-590.

Bibliography

312

Katsuki A, Sumida Y, Gabazza EC, Murashima S, Urakawa H, Morioka K, Kitagawa N, Tanaka T, Araki-Sasaki R, Hori Y, Nakatani K, Yano Y and Adachi Y (2002). Acute hyperinsulinemia is associated with increased circulating levels of adrenomedullin in patients with type 2 diabetes mellitus. Eur J Endocrinol, 147: 71-75. Katsuya T, Koike G, Yee TW, Sharpe N, Jackson R, Norton R, Horiuchi M, Pratt RE, Dzau VJ and MacMahon S (1995). Association of angiotensinogen gene T235 variant with increased risk of coronary heart disease. Lancet, 345: 1600-1603. Katz AM (2002). Maladaptive growth in the failing heart: the cardiomyopathy of overload. Cardiovasc Drugs Ther, 16: 245-249. Katz EB, Burcelin R, Tsao TS, Stenbit AE and Charron MJ (1996). The metabolic consequences of altered glucose transporter expression in transgenic mice. J Mol Med, 74: 639-652. Katz EB, Stenbit AE, Hatton K, Depinho R and Charron MJ (1995). Cardiac and adipose tissue abnormalities but not diabetes in mice deficient in GLUT4. Nature, 377: 151-155. Kearns M, Preis J, McDonald M, Morris C and Whitelaw E (2000). Complex patterns of inheritance of an imprinted murine transgene suggest incomplete germline erasure. Nucleic Acids Res, 28: 3301-3309. Khaled AR, Moor AN, Li A, Kim K, Ferris DK, Muegge K, Fisher RJ, Fliegel L and Durum SK (2001). Trophic factor withdrawal: p38 mitogen-activated protein kinase activates NHE1, which induces intracellular alkalinization. Mol Cell Biol, 21: 7545-7557. Kim HW, Ch YS, Lee HR, Park SY and Kim YH (2001). Diabetic alterations in cardiac sarcoplasmic reticulum Ca2+-ATPase and phospholamban protein expression. Life Sci, 70: 367-379. Kim S, Ohta K, Hamaguchi A, Yukimura T, Miura K and Iwao H (1995). Angiotensin II induces cardiac phenotypic modulation and remodeling in vivo in rats. Hypertension, 25: 1252-1259. King H, Aubert RE and Herman WH (1998). Global burden of diabetes 1995-2025. Diabetes Care, 21: 1414-1431. Kiss E, Ball NA, Kranias EG and Walsh RA (1995). Differential changes in cardiac phospholamban and sarcoplasmic reticular Ca2+-ATPase protein levels. Circ Res, 77: 759-764. Kita Y, Shimizu M, Shibayama S, Yoshio H, Ino H and Mabuchi H (1996). Correlation between myocardial dysfunction and changes in myosin isoenzymes in diabetic rat hearts. J Diabetes Complications, 10: 38-42.

Bibliography

313

Kitamura H, Ohnishi Y, Yoshida A, Okajima K, Azumi H, Ishida A, Galeano EJ, Kubo S, Hayashi Y, Itoh H and Yokoyama M (2002). Heterogeneous loss of connexin43 protein in nonischemic dilated cardiomyopathy with ventricular tachycardia. J Cardiovasc Electrophysiol, 13: 865-870. Klein F (1995). Hyperglycemia and microvascular and macrovascular disease in diabetes. Diabetes Care, 18: 258-268. Klett CP, Printz MP, Bader M, Ganten D and Eggena P (1996). Angiotensinogen messenger RNA stabilization by angiotensin II. J Hypertens, 15 (Suppl): S25-S36. Klionsky DJ and Emr SD (2000). Autophagy as a regulated pathway of cellular degradation. Science, 290: 1717-1721. Knaapen MW, Davies MJ, De Bie M, Haven AJ, Martinet W and Kockx MM (2001). Apoptotic versus autophagic cell death in heart failure. Cardiovasc Res, 51: 304-312. Kobayashi M, Furukawa Y and Chiba S (1978). Positive chronotropic and inotropic effects of angiotensin II in the dog heart. Eur J Pharmacol, 50: 17-25. Koch-Weser J (1965). Nature of the inotropic action of angiotensin on ventricular myocardium. Circ Res, 16: 230-237. Koegl M, Hoppe T, Schlenker S, Ulrich HD, Mayer TU and Jentsch S (1999). A novel ubiquitination factor, E4, is involved in multiubiquitin chain assembly. Cell, 96: 635-644. Kono T, Ikeda F, Oseko F, Ohmori Y, Nakano R, Muranaka H, Taniguchi A, Imura H, Khosla MC and Bumpus FM (1982). Biological activity of des-asp1-des-arg2-angiotensin II in man. Acta Endocrinol (Copenh), 99: 577-584. Kono T, Oseko F, Shimpo S, Nanno M and Endo J (1975). Biological activity of des-asp1-angiotensin II (angiotensin III) in man. J Clin Endocrinol Metab, 41: 1174-1177. Kook H, Lepore JJ, Gitler AD, Lu MM, Wing-Man Yung W, Mackay J, Zhou R, Ferrari V, Gruber P and Epstein JA (2003). Cardiac hypertrophy and histone deacetylase-dependent transcriptional repression mediated by the atypical homeodomain protein Hop. J Clin Invest, 112: 863-871. Koren MJ, Devereux RB, Casale PN, Savage DD and Laragh JH (1991). Relation of left ventricular mass and geometry to morbidity and mortality in uncomplicated essential hypertension. Ann Intern Med, 114: 345-352. Kostin S, Dammer S, Hein S, Klovekorn WP, Bauer EP and Schaper J (2004). Connexin 43 expression and distribution in compensated and decompensated cardiac hypertrophy in patients with aortic stenosis. Cardiovasc Res, 62: 426-436.

Bibliography

314

Kothary RK, Allen ND, Barton SC, Norris ML and Surani MA (1992). Factors affecting cellular mosaicism in the expression of a lacZ transgene in two-cell stage mouse embryos. Biochem Cell Biol, 70: 1097-1104. Kotsanas G, Delbridge LMD and Wendt IR (2000). Stimulus interval-dependent differences in Ca2+ transients and contractile responses of diabetic rat cardiomyocytes. Cardiovasc Res, 46: 450-462. Krattenmacher R, Knauthe R, Parczyk K, Walker A, Hilgenfeldt U and Fritzemeier KH (1994). Estrogen action on hepatic synthesis of angiotensinogen and IGF-I: direct and indirect estrogen effects. J Steroid Biochem Mol Biol, 48: 207-214. Krizanova O, Orlicky J, Masanova C, Juhaszova M and Hudecova S (1997). Angiotensin I modulates Ca-transport systems in the rat heart through angiotensin II. J Mol Cell Cardiol, 29: 1739-1746. Kulan K, Ural D, Komsuoglu B, Agacdiken A, Goldeli O and Komsuoglu SS (1998). Significance of QTc prolongation on ventricular arrhythmias in patients with left ventricular hypertrophy secondary to essential hypertension. Int J Cardiol, 64: 179-184. Kumar D and Jugdutt BI (2003). Apoptosis and oxidants in the heart. J Lab Clin Med, 142: 288-297. Kurisu S, Ozono R, Oshima T, Kambe M, Ishida T, Sugino H, Matsuura H, Chayama K, Teranishi Y, Iba O, Amano K and Matsubara H (2003). Cardiac angiotensin II type 2 receptor activates the kinin/NO system and inhibits fibrosis. Hypertension, 41: 99-107. Kuroki T, Inoguchi T, Umeda F, Ueda F and Nawata H (1998). High glucose induces alteration of gap junction permeability and phosphorylation of connexin-43 in cultured aortic smooth muscle cells. Diabetes, 47: 931-936. Kurtz A and Wagner C (1999). Regulation of renin secretion by angiotensin II-AT1 receptors. J Am Soc Nephrol, 10 (Suppl 11): S162-S168. Kuznicki J, Filipek A, Heimann P, Kaczmarek L and Kaminska B (1989). Tissue specific distribution of calcyclin--10.5 kDa Ca2+-binding protein. FEBS Lett, 254: 141-144. Kuznicki J, Kordowska J, Puzianowska M and Wozniewicz BM (1992). Calcyclin as a marker of human epithelial cells and fibroblasts. Exp Cell Res, 200: 425-430. Lagadic-Gossmann D, Buckler KJ, Le Prigent K and Feuvray D (1996). Altered Ca2+ handling in ventricular myocytes isolated from diabetic rats. Am J Physiol, 270: H1529-H1537.

Bibliography

315

Larkin JE, Frank BC, Gaspard RM, Duka I, Gavras H and Quackenbush F (2004). Cardiac transcriptional response to acute and chronic angiotensin II treatment. Physiol Genomics, 18: 152-166. Larsen LA, Christiansen M, Vuust J and Andersen PS (2001). Recent developments in high-throughput mutation screening. Pharmacogenomics, 2: 387-399. Larsen N, Samuelsson T and Zwieb C (1998). The Signal Recognition Particle Database (SRPDB). Nucleic Acids Res, 26: 177-178. Lau S, Jardine K and McBurney MW (1999). DNA methylation pattern of a tandemly repeated LacZ transgene indicates that most copies are silent. Dev Dyn, 215: 126-138. Lavoie C, Mercier JF, Salahpour A, Umapathy D, Breit A, Villeneuve LR, Zhu WZ, Xiao RP, Lakatta EG, Bouvier M and Hebert TE (2002). Beta 1/beta 2-adrenergic receptor heterodimerization regulates beta 2-adrenergic receptor internalization and ERK signaling efficacy. J Biol Chem, 277: 35402-35410. Le Prigent K, Lagadic-Gossmann D and Feuvray D (1997). Modulation by pH0 and intracellular Ca2+ of Na+-H+ exchange in diabetic rat isolated ventricular myocytes. Circ Res, 80: 253-260. Lear W, Ruzicka M and Leenen FHH (1997). ACE inhibitors and cardiac ACE mRNA in volume overload-induced cardiac hypertrophy. Am J Physiol, 273: H641-H646. Lecarpentier Y, Bugaisky LB, Chemla D, Mercadier JJ, Schwartz K, Whalen RG and Martin JL (1987). Coordinated changes in myocardial contractility, energetics and myosin isozyme pattern following thoracic aortic stenosis in the young rat. Am J Physiol, 252: H275-H282. Lee AA, Dillmann WH, McCulloch AD and Villarreal FJ (1995). Angiotensin II stimulates the autocrine production of transforming growth factor-beta 1 in adult rat cardiac fibroblasts. J Mol Cell Cardiol, 27: 2347-2357. Lee DS, Johansen H, Gong Y, Hall RE, Tu JV, Cox JL and Canadian Cardiovascular Outcomes Research Team (2004). Regional outcomes of heart failure in Canada. Can J Cardiol, 20: 599-607. Lee WL, Chen JW, Ting CT, Ishiwata T, Lin SJ, Korc M and Wang PH (1999). Insulin-like growth factor I improves cardiovascular function and suppresses apoptosis of cardiomyocytes in dilated cardiomyopathy. Endocrinology, 140: 4831-4840. Leenders F, Husen B, Thole HH and Adamski J (1994). The sequence of porcine 80 kDa 17 beta-estradiol dehydrogenase reveals similarities to the short chain alcohol dehydrogenase family, to actin binding motifs and to sterol carrier protein 2. Mol Cell Endocrinol, 104: 127-131.

Bibliography

316

Lefroy DC, Crake T, Del Monte F, Vescovo G, Dalla Libera L, Harding S and Poole-Wilson PA (1996). Angiotensin II and contraction of isolated myocytes from human, guinea pig, and infarcted rat hearts. Am J Physiol, 270: H2060-H2069. Leri A, Fiordaliso F, Setoguchi M, Limana F, Bishopric NH, Kajstura J, Webster K and Anversa P (2000). Inhibition of p53 function prevents renin-angiotensin system activation and stretch-mediated myocyte apoptosis. Am J Pathol, 157: 843-857. Levy D, Garrison RJ, Savage DD, Kannel WB and Castelli WP (1990). Prognostic implications of echocardiographically determined left ventricular mass in the Framingham Heart Study. N Engl J Med, 322: 1561-1566. Levy D, Larson MG, Vasan RS, Kannel WB and Ho KK (1996). The progression from hypertension to congestive heart failure. JAMA, 275: 1557-1562. Li D, Fareh S, Leung TK and Nattel S (1999). Promotion of atrial fibrillation by heart failure in dogs: atrial remodeling of a different sort. Circulation, 100: 87-95. Li GR, Lau CP, Ducharme A, Tardif JC and Nattel S (2002). Transmural action potential and ionic current remodeling in ventricles of failing canine hearts. Am J Physiol, 283: H1031-H1041. Li P, Sonnenblick EH, Anversa P and Capasso JM (1994). Length-dependent modulation of Ang II inotropism in rat myocardium: effects of myocardial infarction. Am J Physiol, 266: H779-H786. Li X, Alvarez B, Casey JR, Reithmeier RA and Fliegel L (2002). Carbonic anhydrase II binds to and enhances activity of the Na+/H+ exchanger. J Biol Chem, 277: 36085-36091. Li Z, Iwai M, Wu L, Shiuchi T, Jinno T, Cui TX and Horiuchi M (2003). Role of AT2 receptor in the brain in regulation of blood pressure and water intake. Am J Physiol, 284: H116-H121. Liao P, Wang SQ, Wang S, Zheng M, Zheng M, Zhang SJ, Cheng H, Wang Y and Xiao RP (2002). p38 Mitogen-activated protein kinase mediates a negative inotropic effect in cardiac myocytes. Circ Res, 90: 190-196. Liedtke AJ, DeMaison L, Eggleston AM, Cohen LM and Nellis SH (1988). Changes in substrate metabolism and effects of excess fatty acids in reperfused myocardium. Circ Res, 62: 535-542. Lim CC, Liao R, Varma N and Apstein CS (1999). Impaired lusitropy-frequency in the aging mouse: role of Ca2+-handling proteins and effects of isoproterenol. Am J Physiol, 274: H2083-H2090. Lim CC, Apstein CS, Colucci WS and Liao R (2000). Impaired cell shortening and relengthening with increased pacing frequency are intrinsic to the senescent mouse cardiomyocyte. J Mol Cell Cardiol, 32: 2075-2082.

Bibliography

317

Lints TJ, Parsons LM, Hartley L, Lyons I and Harvey RP (1993). Nkx-2.5: a novel murine homeobox gene expressed in early heart progenitor cells and their myogenic descendants. Development, 119: 419-431. Lipshutz RJ, Fodor SP, Gingeras TR and Lockhart DJ (1999). High density synthetic oligonucleotide arrays. Nat Genet, 21 (Suppl 1): 20-24. Liu C, Li Y, Semenov M, Han C, Baeg GH, Tan Y, Zhang Z, Lin X and He X (2002). Control of beta-catenin phosphorylation/degradation by a dual-kinase mechanism. Cell, 108: 837-847. Liu CP, Ting CT, Lawrence W, Maughan WL, Chang MS and Kass DA (1993). Diminished contractile response to increased heart rate in intact human left ventricular hypertrophy: systolic versus diastolic determinants. Circulation, 88: 1893-1906. Liu J, Masurekar MR, Vatner DE, Jyothirmayi GN, Regan TJ, Vatner SF, Meggs LG and Malhotra A (2003). Glycation end-product cross-link breaker reduces collagen and improves cardiac function in aging diabetic heart. Am J Physiol, 285: H2587-H2591. Liu X, Sentex E, Golfman L, Takeda S, Osada M and Dhalla NS (1999). Modification of cardiac subcellular remodeling due to pressure overload by captopril and losartan. Clin Exp Hypertens, 21: 145-156. Liu ZP and Olson EN (2002). Suppression of proliferation and cardiomyocyte hypertrophy by CHAMP, a cardiac-specific RNA helicase. Proc Natl Acad Sci, 99: 2043-2048. Llewellyn DH and Roderick HL (1998). Overexpression of calreticulin fails to abolish its induction by perturbation of normal ER function. Biochem Cell Biol, 76: 875-880. Lockhart DJ, Dong H, Byrne MC, Follettie MT, Gallo MV, Chee MS, Mittmann M, Wang C, Kobayashi M, Horton H and Brown EL (1996). Expression monitoring by hybridization to high-density oligonucleotide arrays. Nat Biotechnol, 14: 1675-1680. Lockwood EA and Bailey E (1970). Fatty acid utilization during development of the rat. Biochem J, 120: 49-54. Loiselle FB, Morgan PE, Alvarez BV and Casey JR (2004). Regulation of the human NBC3 Na+/HCO3- cotransporter by carbonic anhydrase II and PKA. Am J Physiol, 286: C1423-C1433. Lompre AM, Nadal-Ginard B and Mahdavi V (1984). Expression of the cardiac ventricular alpha- and beta-myosin heavy chain is developmentally and hormonally regulated. J Biol Chem, 259: 6437-6446. Lönnstedt I and Britton T (2005). Hierarchical Bayes models for cDNA microarray gene expression. Biostatistics, 6: 279-291.

Bibliography

318

Loot AE, Roks AJM, Henning RH, Tio RA, Suurmeijer AJH, Boomsma F and van Gilst WH (2002). Angiotensin-(1-7) attenuates the development of heart failure after myocardial infarction in rats. Circulation, 105: 1548-1550. Lopes CMB, Gallagher PG, Buck ME, Butler MH and Goldstein SAN (2000). Proton block and voltage-gating are potassium-dependent in the cardiac leak channel Kcnk3. J Biol Chem, 275: 16969-16978. Lou XJ, Schena M, Horrigan FT, Lawn RM and Davis RW (2001). Expression monitoring using cDNA microarrays. A general protocol. Methods Mol Biol, 175: 323-340. Lough J and Sugi Y (2000). Endoderm and heart development. Dev Dyn, 217: 327-342. Lucius R, Gallinat S, Rosenstiel P, Herdegen T, Sievers J and Unger T (1998). The angiotensin II type 2 (AT2) receptor promotes axonal regeneration in the optic nerve of adult rats. J Exp Med, 188: 661-670. Luiken JJ, Turkotte LP and Bonen A (1999). Protein-mediated palmitate uptake and expression of fatty acid transport proteins in heart giant vesicles. J Lipid Res, 40: 1007-1016. MacGregor DP, Murone C, Song K, Allen AM, Paxinos G and Mendelsohn FA (1995). Angiotensin II receptor subtypes in the human central nervous system. Brain Res, 675: 231-240. Mackenzie L, Bootman MD, Laine M, Berridge MJ, Thuring J, Holmes A, Li WH and Lipp P (2002). The role of inositol 1,4,5-trisphosphate receptors in Ca(2+) signalling and the generation of arrhythmias in rat atrial myocytes. J Physiol 541: 395-409. Maier LS, Bers DM and Pieske B (2000). Differences in Ca(2+)-handling and sarcoplasmic reticulum Ca(2+)-content in isolated rat and rabbit myocardium. J Mol Cell Cardiol, 32: 2249-2258. Maisel AS, McCord J, Nowak RM, Hollander JE, Wu AH, Duc P, Omland T, Storrow AB, Krishnaswamy P, Abraham WT, Clopton P, Steg G, Aumont MC, Westheim A, Knudsen CW, Perez A, Kamin R, Kazanegra R, Herrmann HC, McCullough PA and Breathing Not Properly Multinational Study Investigators (2003). Bedside B-Type natriuretic peptide in the emergency diagnosis of heart failure with reduced or preserved ejection fraction. Results from the Breathing Not Properly Multinational Study. J Am Coll Cardiol, 41: 2010-2017. Majno G and Joris I (1995). Apoptosis, oncosis, and necrosis. An overview of cell death. Am J Pathol, 146: 3-15. Malhotra A, Mordes JP, McDermott L and Schaible TF (1985). Abnormal cardiac biochemistry in spontaneously diabetic Bio-Breeding/Worcester rat. Am J Physiol, 249: H1051-H1055.

Bibliography

319

Malhotra A, Reich D, Reich D, Nakouzi A, Sanghi V, Geenen DL and Buttrick PM (1997). Experimental diabetes is associated with functional activation of protein kinase C epsilon and phosphorylation of troponin I in the heart, which are prevented by angiotensin II receptor blockade. Circ Res, 81: 1027-1033. Malmqvist K, Wallen HN, Held C and Kahan T (2002). Soluble cell adhesion molecules in hypertensive concentric left ventricular hypertrophy. J Hypertens, 20: 1563-1569. Maly K, Strese K, Kampfer S, Ueberall F, Baier G, Ghaffari-Tabrizi N, Grunicke HH and Leitges M (2002). Critical role of protein kinase C alpha and calcium in growth factor induced activation of the Na+/H+ exchanger NHE1. FEBS Lett, 521: 205-210. Manchester J, Kong X, Nerbonne J, Lowry OH and Lawrence Jr JC (1994). Glucose transport and phosphorylation in single cardiac myocytes: rate-limiting steps in glucose metabolism. Am J Physiol, 266: E326-E333. Mancina R, Susini T, Renzetti A, Forti G, Razzoli E, Serio M and Maggi M (1996). Sex steroid modulation of AT2 receptors in human myometrium. J Clin Endocrinol Metab, 81: 1753-1757. Marano G, Formigari R and Vergari A (1997). Effects of angiotensin II on myocardial contractility during short-term pressor responses to angiotensin II. J Hypertens, 15: 1019-1025. Marks AR (2001). Ryanodine receptors/calcium release channels in heart failure and sudden cardiac death. J Mol Cell Cardiol, 33: 615-624. Marks AR, Marx SO and Reiken S (2002). Regulation of ryanodine receptors via macromolecular complexes: a novel role for leucine/isoleucine zippers. Trends Cardiovasc Med, 12: 166-170. Maron BJ, Poliac LC and Roberts WO (1996). Risk for sudden cardiac death associated with marathon running. J Am Coll Cardiol, 28: 428-431. Maron BJ (2003). Sudden death in young athletes. N Engl J Med, 349: 1064-1075. Marvin MJ, Di Rocco G, Gardiner A, Bush SM and Lassar AB (2001). Inhibition of Wnt activity induces heart formation from posterior mesoderm. Genes Dev, 15: 316-327. Masaki H, Kurihara T, Yamaki A, Inomata N, Nozawa Y, Mori Y, Murasawa S, Kizima K, Maruyama K, Horiuchi M, Dzau VJ, Takahashi H, Iwasaka T, Inada M and Matsubara H (1998). Cardiac-specific overexpression of angiotensin II AT2 receptor causes attenuated response to AT1 receptor-mediated pressor and chronotropic effects. J Clin Invest, 101: 527-535.

Bibliography

320

Masseroli M, Martucci D and Pinciroli F (2004). GFINDer: Genome Function INtegrated Discoverer through dynamic annotation, statistical analysis, and mining. Nucleic Acids Res, 32 (Web Server issue): W293-W300. Mathew J, Sleight P, Lonn E, Johnstone D, Pogue J, Yi Q, Bosch J, Sussex B, Probstfield J, Yusuf S and Heart Outcomes Prevention Evaluation (HOPE) Investigators (2001). Reduction of cardiovascular risk by regression of electrocardiographic markers of left ventricular hypertrophy by the angiotensin-converting enzyme inhibitor ramipril. Circulation, 104: 1615-1621. Mattiello JA, Margulies KB, Jeevanandam V and Houser SR (1988). Contribution of reverse-mode sodium-calcium exchange to contractions in failing human left ventricular myocytes. Cardiovasc Res, 37: 424-431. Matsui H, Barry WH, Livsey C and Spitzer KW (1995). Angiotensin II stimulates sodium-hydrogen exchange in adult rabbit ventricular myocytes. Cardiovasc Res, 29: 215-221. Mattiazzi A, Perez NG, Vila-Petroff MG, Alvarez B, Camilion de Hurtado MC and Cingolani HE (1997). Dissociation between positive inotropic and alkalinizing effects of angiotensin II in feline myocardium. Am J Physiol, 272: H1131-H1136. Mazzocco M, Maffei M, Egeo A, Vergano A, Arrigo P, Di Lisi R, Ghiotto F and Scartezzini P (2002). The identification of a novel human homologue of the SH3 binding glutamic acid-rich (SH3BGR) gene establishes a new family of highly conserved small proteins related to Thioredoxin Superfamily. Gene, 291: 233-239. Mazzolai L, Nussberger J, Aubert JF, Brunner DB, Gabbiani G, Brunner HR and Pedrazzini T (1998). Blood pressure-independent cardiac hypertrophy induced by locally activated renin-angiotensin system. Hypertension, 31: 1324-1330. Mazzolai L, Pedrazzini T, Nicoud F, Gabbiani G, Brunner HR and Nussberger J (2000). Increased cardiac angiotensin II levels induce right and left ventricular hypertrophy in normotensive mice. Hypertension, 35: 985-991. McIntyre H and Fry CH (1997). Abnormal action potential conduction in isolated human hypertrophied left ventricular myocardium. J Cardiovasc Electrophysiol, 8: 887–894. Meacham GC, Patterson C, Zhang W, Younger JM and Cyr DM (2001). The Hsc70 co-chaperone CHIP targets immature CFTR for proteasomal degradation. Nat Cell Biol, 3: 100-105. Medugorac I (1980). Collagen content in different areas of normal and hypertrophied rat myocardium. Cardiovasc Res, 14: 551-554. Meissner A, Min JY and Simon R (1998). Effects of angiotensin II on inotropy and intracellular Ca2+ handling in normal and hypertrophied rat myocardium. J Mol Cell Cardiol, 30: 2507-2518.

Bibliography

321

Menard J, Bouhnik J, Clauser E, Richoux JP and Corvol P (1983). Biochemistry and regulation of angiotensinogen. Clin Exp Hypertens, 5: 1005-1019. Menard J, El Amrani AIK, Savoie F and Bouhnik (1991). Angiotensinogen: an attractive and underrated participant in hypertension and inflammation. Hypertension, 18: 705-707. Meulemans AL, Andries LJ and Brutsaert DL (1990). Does endocardial endothelium mediate positive inotropic response to angiotensin I and angiotensin II? Circ Res, 66: 1591-1601. Michael LF, Wu Z, Cheatham RB, Puigserver P, Adelmant G, Lehman JJ, Kelly DP and Spiegelman BM (2001). Restoration of insulin-sensitive glucose transporter (GLUT4) gene expression in muscle cells by the transcriptional coactivator PGC-1. Proc Natl Acad Sci U S A, 98: 3820-3825. Michels G, Er F, Khan I, Sudkamp M, Herzig S and Hoppe UC (2005). Single-channel properties support a potential contribution of hyperpolarization-activated cyclic nucleotide-gated channels and If to cardiac arrhythmias. Circulation, 111: 399-404. Miller TB (1983). Altered regulation of cardiac glycogen metabolism in spontaneously diabetic rats. Am J Physiol, 245: E379-E383. Mills AA, Mills MJ, Gardiner DM, Bryant SV and Stanbridge EJ (1999). Analysis of the pattern of QM expression during mouse development. Differentiation, 64: 161-171. Milnes JT and MacLeod KT (2001). Reduced ryanodine receptor to dihydropyridine receptor ratio may underlie slowed contraction in a rabbit model of left ventricular cardiac hypertrophy. J Mol Cell Cardiol, 33: 473-485. Min LJ, Cui TX, Yahata Y, Yamasaki K, Shiuchi T, Liu HW, Chen R, Li JM, Okumura M, Jinno T, Wu L, Iwai M, Nahmias C, Hashimoto K and Horiuchi M (2004). Regulation of collagen synthesis in mouse skin fibroblasts by distinct angiotensin II receptor subtypes. Endocrinology, 145: 253-260. Minamino T, Yujiri T, Papst PJ, Chan ED, Johnson GL and Terada N (1999). MEKK1 suppresses oxidative stress-induced apoptosis of embryonic stem cell-derived cardiac myocytes. Proc Natl Acad Sci U S A, 96: 15127-15132. Miniou P, Tiziano D, Frugier T, Roblot N, Le Meur M and Melki J (1999). Gene targeting restricted to mouse striated muscle lineage. Nucleic Acids Res, 27: e27. Mirotsou M, Watanabe CM, Schultz PG, Pratt RE and Dzau VJ (2003). Elucidating the molecular mechanism of cardiac remodeling using a comparative genomic approach. Physiol Genomics, 15: 115-126. Missov ED and De Marco T (1999). Clinical insights on the use of highly sensitive cardiac troponin assays. Clin Chim Acta, 284: 175-185.

Bibliography

322

Mitarai S, Reed TD and Yatani A (2000). Changes in ionic currents and beta-adrenergic receptor signaling in hypertrophied myocytes overexpressing G alpha(q). Am J Physiol, 279: H139-H148. Miyashita T and Reed JC (1995). Tumor suppressor p53 is a direct transcriptional activator of the human bax gene. Cell, 80: 293-299. Mizushige K, Yao L, Noma T, Kiyomoto H, Yu Y, Hosomi N, Ohmori K and Matsuo H (2000). Alteration in left ventricular diastolic filling and accumulation of myocardial collagen at insulin-resistant prediabetic stage of a type II diabetic rat model. Circulation, 101: 899-907. Mochizuki T, Eberli FR, Apstein CS and Lorell BH (1992). Exacerbation of ischemic dysfunction by angiotensin II in red cell-perfused rabbit hearts. Effects on coronary flow, contractility, and high-energy phosphate metabolism. J Clin Invest, 89: 490-498. Modesti PA, Vanni S, Bertolozzi I, Cecioni I, Polidori G, Paniccia R, Bandinelli B, Perna A, Liguori P, Boddi M, Galanti G, Serneri GG (2000). Early sequence of cardiac adaptations and growth factor formation in pressure- and volume-overload hypertrophy. Am J Physiol, 279: H976-H985. Molkentin JD and Olson EN (1997). GATA4: a novel transcriptional regulator of cardiac hypertrophy? Circulation, 96: 3833-3835. Molkentin JD, Lu JR, Antos CL, Markham B, Richardson J, Robbins J, Grant SR, and Olson EN (1998). A calcineurin-dependent transcriptional pathway for cardiac hypertrophy. Cell, 93: 215-228. Moller G, van Grunsven EG, Wanders RJn and Adamski J (2001). Molecular basis of D-bifunctional protein deficiency. Mol Cell Endocrinol, 171: 61-70. Montiel M, Barker S, Vinson GP and Jimenez E (1993). Angiotensin II receptor isoforms in the rat adrenal gland: studies with selective subtype antagonists DuP 753 and CGP41221A. J Mol Endocrinol, 11: 69-75. Moor AN and Fliegel L (1999). Protein kinase-mediated regulation of the Na(+)/H(+) exchanger in the rat myocardium by mitogen-activated protein kinase-dependent pathways. J Biol Chem, 274: 22985-22992. Moorman AF and Christoffels VM (2003). Cardiac chamber formation: development, genes, and evolution. Physiol Rev, 83: 1223-1267. Moravec CS, Schluchter MD, Paranandi L, Czerska B, Stewart RW, Rosenkranz E and Bond M (1990). Inotropic effects of angiotensin II on human cardiac muscle in vitro. Circulation, 82: 1973-1984. Morgan TO and Delbridge LMD (1999). Angiotensin blocking drugs and the heart beyond 2000. J Am Soc Nephrol, 10: 243-247.

Bibliography

323

Morii I, Kihara Y, Inoko M and Sasayama S (1998). Myocardial contractile efficiency and oxygen cost of contractility are preserved during transition from compensated hypertrophy to failure in rats with salt-sensitive hypertension. Hypertension, 31: 949-960. Mukherjee R and Spinale FG (1998). L-type calcium channel abundance and function with cardiac hypertrophy and failure: a review. J Mol Cell Cardiol, 30: 1899-1916. Mukherjee R, Hewett KW and Spinale FG (1995). Myocyte electrophysiological properties following the development of supraventricular tachycardia-induced cardiomyopathy. J Mol Cell Cardiol, 27: 1333-1348. Mukoyama M, Nakajima M, Horiuchi M, Sasamura H, Pratt RE and Dzau VJ (1993). Expression cloning of type 2 angiotensin II receptor reveals a unique class of seven-transmembrane receptors. J Biol Chem, 268: 24539-24542. Mullins JJ, Peters J and Ganten D (1990). Fulminant hypertension in transgenic rats harbouring the mouse Ren-2 gene. Nature, 344: 541-544. Murakami K, Yumoto F, Ohki SY, Yasunaga T, Tanokura M and Wakabayashi T (2005). Structural basis for Ca2+-regulated muscle relaxation at interaction sites of troponin with actin and tropomyosin. J Mol Biol, 352: 178-201. Murray DB, Gardner JD, Brower GL and Janicki JS (2004). Endothelin-1 mediates cardiac mast cell degranulation, matrix metalloproteinase activation, and myocardial remodeling in rats. Am J Physiol, 287: H2295-H2299. Muscat GE and Kedes L (1987). Multiple 5'-flanking regions of the human alpha-skeletal actin gene synergistically modulate muscle-specific expression. Mol Cell Biol, 7: 4089-4099. Muscella A, Marsigliante S, Vilella S, Jimenez E and Storelli C (1999). Angiotensin II stimulates Na+/H+ exchanger in human umbilical vein endothelial cells via AT1 receptor. Life Sci, 65: 2385-2394. Miyamoto S, Teramoto H, Gutkind JS and Yamada KM (1996). Integrins can collaborate with growth factors for phosphorylation of receptor tyrosine kinases and MAP kinase activation: roles of integrin aggregation and occupancy of receptors. J Cell Biol, 135: 1633–1642. Miyashita T and Reed JC (1995). Tumor suppressor p53 is a direct transcriptional activator of the human bax gene. Cell, 80: 293-299. Myers ML, Farhangkhoee P and Karmazyn M (1998). Hydrogen peroxide induced impairment of post-ischemic ventricular function is prevented by the sodium-hydrogen exchange inhibitor HOE 642 (cariporide). Cardiovasc Res, 40: 290-296.

Bibliography

324

Nagano M, Higaki J, Nakamura F, Higashimori K, Nagano N, Mikami H and Ogihara T (1992). Role of cardiac angiotensin II in isoproterenol-induced left ventricular hypertrophy. Hypertension, 19: 708-712. Nagao M, Parimoo B and Tanaka K (1993). Developmental, nutritional, and hormonal regulation of tissue-specific expression of the genes encoding various acyl-CoA dehydrogenases and a-subunit of electron transfer flavoprotein in rat. J Biol Chem, 268: 24114-24124. Nagata M, Tanimoto K, Fukamizu A, Kon Y, Sugiyama F, Yagami K, Murakami K and Watanabe T (1996). Nephrogenesis and renovascular development in angiotensinogen-deficient mice. Lab Invest, 75: 745-753. Nagy A (2000). Cre recombinase: the universal reagent for genome tailoring. Genesis, 26: 99-109. Naudin V, Oliviero P, Rannou F, Sainte Beuve C and Charlemagne D (1991). The density of ryanodine receptors decreases with pressure overload-induced rat cardiac hypertrophy. FEBS Lett, 285: 135-138. Neely JR, Rovetto MJ and Oram JF (1972). Myocardial utilization of carbohydrate and lipids. Prog Cardiovasc Dis, 15: 289-329. Neri Serneri GG, Boddi M, Modesti PA, Cecioni I, Coppo M, Padeletti L, Michelucci A, Colella A and Galanti G (2001). Increased cardiac sympathetic activity and insulin-like growth factor-I formation are associated with physiological hypertrophy in athletes. Circ Res, 89: 977-982. Negretti N, O'Neill SC and Eisner DA (1993). The relative contributions of different intracellular and sarcolemmal systems to relaxation in rat ventricular myocytes. Cardiovasc Res, 27: 1826-1830. Neyses L and Vetter H (1989). Action of atrial natriuretic peptide and angiotensin II on the myocardium: studies in isolated rat ventricular cardiomyocytes. Biochem Biophys Res Commun, 163: 1435-1443. Nigam SK, Goldberg AL, Ho S, Rohde MF, Bush KT and Sherman My (1994). A set of endoplasmic reticulum proteins possessing properties of molecular chaperones includes Ca(2+)-binding proteins and members of the thioredoxin superfamily. J Biol Chem, 269: 1744-1749. Niggli E and Egger M (2002). Calcium quarks. Front Biosci, 7: d1288-d1297. Nishida M, Futami S, Morita I, Maekawa K and Murota SI (2000). Hypoxia-reoxygenation inhibits gap junctional communication in cultured human umbilical vein endothelial cells. Endothelium, 7: 279-286.

Bibliography

325

Nishimoto S, Tawara J, Toyoda H, Kitamura K and Komurasaki T (2003). A novel homocysteine-responsive gene, smap8, modulates mitogenesis in rat vascular smooth muscle cells. Eur J Biochem, 270: 2521-2531. Nishizawa K, Freund C, Li J, Wagner G and Reinherz EL (1998). Identification of a proline-binding motif regulating CD2-triggered T lymphocyte activation. Proc Natl Acad Sci U S A, 95: 14897-14902. Noda N, Hayashi H, Miyata H, Suzuki S, Kobayashi A and Yamazaki N (1992). Cytosolic Ca2+ concentration and pH of diabetic rat myocytes during metabolic inhibition. J Mol Cell Cardiol, 24: 435-446. Noda N, Hayashi H, Satoh H, Terada H, Hirano M, Kobayashi A and Yamazaki N (1993). Ca2+ transients and cell shortening in diabetic rat ventricular myocytes. Jpn Circ J, 57: 449-457. Noguchi T, Kihara Y, Begin KJ, Gorga JA, Palmiter KA, LeWinter MM and VanBuren P (2003). Altered myocardial thin-filament function in the failing Dahl salt-sensitive rat heart: amelioration by endothelin blockade. Circulation, 107: 630-635. Nowotny M, Spiechowicz M, Jastrzebska B, Filipek A, Kitagawa K and Kuznicki J (2003). Calcium-regulated interaction of Sgt1 with S100A6 (calcyclin) and other S100 proteins. J Biol Chem, 278: 26923-26928. Nuedling S, Kahlert S, Loebbert K, Meyer R, Vetter H and Grohe C (1999). Differential effects of 17beta-estradiol on mitogen-activated protein kinase pathways in rat cardiomyocytes. FEBS Lett, 454: 271-276. Nuss HB, Johns DC, Kaab S, Tomaselli GF, Kass D, Lawrence JH and Marban E (1996). Reversal of potassium channel deficiency in cells from failing hearts by adenoviral gene transfer: a prototype for gene therapy for disorders of cardiac excitability and contractility. Gene Ther, 3: 900-912. Oh HS, Kwon H, Sun SK and Yang CH (2002). QM, a putative tumor suppressor, regulates proto-oncogene c-yes. J Biol Chem, 277: 36489-36498. Oikarinen L, Nieminen MS, Viitasalo M, Toivonen L, Wachtell K, Papademetriou V, Jern S, Dahlof B, Devereux RB and Okin PM (2001). Relation of QT interval and QT dispersion to echocardiographic left ventricular hypertrophy and geometric pattern in hypertensive patients. The LIFE study. The Losartan Intervention For Endpoint Reduction. J Hypertens, 19: 1883-1891. Okruhlicova L, Tribulova N, Misejkova M, Kucka M, Stetka R, Slezak J and Manoach M (2002). Gap junction remodelling is involved in the susceptibility of diabetic rats to hypokalemia-induced ventricular fibrillation. Acta Histochem, 104: 387-391.

Bibliography

326

Okumura H, Nagaya N, Itoh T, Okano I, Hino J, Mori K, Tsukamoto Y, Ishibashi-Ueda H, Miwa S, Tambara K, Toyokuni S, Yutani C and Kangawa K (2004). Adrenomedullin infusion attenuates myocardial ischemia/reperfusion injury through the phosphatidylinositol 3-kinase/Akt-dependent pathway. Circulation, 109: 242-248. Okumura M, Iwai M, Ide A, Mogi M, Ito M and Horiuchi M (2005). Sex difference in vascular injury and the vasoprotective effect of valsartan are related to differential AT2 receptor expression. Hypertension, 46: 577-583. Olichon A, Baricault L, Gas N, Guillou E, Valette A, Belenguer P and Lenaers G (2003). Loss of OPA1 perturbates the mitochondrial inner membrane structure and integrity, leading to cytochrome c release and apoptosis. J Biol Chem, 278: 7743-7746. Oliverio MI, Kim HS, Ito M, Le T, Audoly L, Best CF, Hiller S, Kluckman K, Maeda N, Smithies O and Coffman TM (1998). Reduced growth, abnormal kidney structure, and type 2 (AT2) angiotensin receptor-mediated blood pressure regulation in mice lacking both AT1A and AT1B receptors for angiotensin II. Proc Natl Acad Sci U S A, 5: 15496-15501. Olivetti G, Abbi R, Quaini F, Kajstura J, Cheng W and Nitahara JA (1997). Apoptosis in the failing human heart. N Engl J Med, 336: 1131-1141. Olivetti G, Capasso JM, Sonnenblick EH and Anversa P (1990). Side-to-side slippage of myocytes in ventricular wall remodeling acutely after myocardial infarction in rats. Circ Res, 67: 23-24. Olson EN and Srivastava D (1996). Molecular pathways controlling heart development. Science, 272: 671-676. Oltvai ZN, Milliman CL and Korsmeyer SJ (1993). Bcl-2 heterodimerizes in vivo with a conserved homolog, Bax, that accelerates programmed cell death. Cell, 74: 609-619. Opie LH (1968). Metabolism of the heart in health and disease. Am Heart J, 76:685-689. Orchard CH and Lakatta EG (1986). Intracellular calcium transients and developed tension in rat heart muscle. A mechanism for the negative interval-strength relationship. J Gen Physiol, 86: 637-651. Orlic D, Kajstura J, Chimenti S, Jakoniuk I, Anderson SM, Li B, Pickel J, McKay R, Nadal-Ginard B, Bodine DM, Leri A and Anversa P (2001). Bone marrow cells regenerate infarcted myocardium. Nature, 410: 240-243. Orlowski J and Grinstein S (2003). In M. Karmazyn, M. Avkiran, & L. Fliegel (Eds.), The Na+/H+ exchanger, from molecular to its role in disease (pp. 17–34). Boston, Dordrecht, London: Kluwer Academic Publishers. Oudit GY, Crackower MA, Backx PH and Penninger JM (2003). The role of ACE2 in cardiovascular physiology. Trends Cardiovasc Med, 13: 93-101.

Bibliography

327

Oudit GY, Sun H, Trivieri MG, Koch SE, Dawood F, Ackerley C, Yazdanpanah M, Wilson GJ, Schwartz A, Liu PP and Backx PH (2003). L-type Ca2+ channels provide a major pathway for iron entry into cardiomyocytes in iron-overload cardiomyopathy. Nat Med, 9: 1187-1194. Pagliassotti MJ, Prach PA, Koppenhafer TA and Pan DA (1996). Changes in insulin action, triglycerides, and lipid composition during sucrose feeding in rats. Am J Physiol, 271: R1319-R1326. Pahor M, Bernabei R, Sgadari A, Gambassi G Jr, Lo Giudice P, Pacifici L, Ramacci MT, Lagrasta C, Olivetti G and Carbonin P (1991). Enalapril prevents cardiac fibrosis and arrhythmias in hypertensive rats. Hypertension, 18: 148-157. Palakodeti V, Oh S, Oh BH, Mao L, Hongo M, Peterson KL and Ross J (1997). Force-frequency effect is a powerful determinant of myocardial contractility in the mouse. Am J Physiol, 273: H1283-H1290. Palatini P and Julius S (2004). Elevated heart rate: a major risk factor for cardiovascular disease. Clin Exp Hypertens, 26: 637-644. Paolisso G and Giugliano D (1996). Oxidative stress and insulin action: is there a relationship ? Diabetologia, 39: 357-363. Paolisso G, De Riu S, Marrazzo G, Verza M, Varricchio M and D’Onofrio F (1991). Insulin resistance and hyperinsulinemia in patients with chronic congestive heart failure. Metabolism, 40: 972-977. Paradis P, Dali-Youcef N, Paradis FW, Thibault G and Nemer M (2000). Overexpression of angiotensin II type I receptor in cardiomyocytes induces cardiac hypertrophy and remodeling. Proc Natl Acad Sci U S A, 97: 931-936. Parissis JT, Adamopoulos SN, Venetsanou KF, Karas SM and Kremastinos DT (2003). Elevated plasma amylase levels in advanced chronic heart failure secondary to ischemic or idiopathic dilated cardiomyopathy: correlation with circulating interleukin-6 activity. J Interferon Cytokine Res, 23: 329-333. Parkes JG, Hussain RA, Olivieri NF and Templeton DM (1993). Effects of iron loading on uptake, speciation, and chelation of iron in cultured myocardial cells. J Lab Clin Med, 122: 36-47. Passier R, Zeng H, Frey N, Naya FJ, Nicol RL, McKinsey TA, Overbeek P, Richardson JA, Grant SR and Olson EN (2000). CaM kinase signaling induces cardiac hypertrophy and activates the MEF2 transcription factor in vivo. J Clin Invest, 105: 1395-1406. Patel JM, Martens JR, Li YD, Gelband CH, Raizada MK and Block ER (1998). Angiotensin IV receptor-mediated activation of lung endothelial NOS is associated with vasorelaxation. Am J Physiol, 275: L1061-L1068.

Bibliography

328

Patton EE, Willems AR and Tyers M (1998). Combinatorial control in ubiquitin-dependent proteolysis: don't Skp the F-box hypothesis. Trends Genet, 14: 236-243. Paul M, Wagner J and Dzau VJ (1993). Gene expression of the renin-angiotensin system in human tissue. Quantitative analysis by the polymerase chain reaction. J Clin Invest, 91: 2058-2064. Paula RD, Lima CV, Khosla MC and Santos RAS (1995). Angiotensin-(1-7) potentiates the hypotensive effect of bradykinin in conscious rats. Hypertension, 26: 1154-1159. Paulson DJ and Crass MF (1982). Endogenous triacylglycerol metabolism in diabetic heart. Am J Physiol, 242: H1084-H1094. Pauschinger M, Knopf D, Petschauer S, Doerner A, Poller W, Schwimmbeck PL, Kühl U and Schultheiss HP (1999). Dilated cardiomyopathy is associated with significant changes in collagen type I/III ratio. Circulation, 99: 2750-2756. Pelliccia A, Maron BJ, De Luca R, Di Paolo FM, Spataro A and Culasso F (2002) Remodeling of left ventricular hypertrophy in elite athletes after long-term deconditioning. Circulation, 105: 944-949. Pellieux C, Sauthier T, Aubert JF, Brunner HR and Pedrazzini T (2000). Angiotensin II-induced cardiac hypertrophy is associated with different mitogen-activated protein kinase activation in normotensive and hypertensive mice. J Hypertens, 18: 1307-1317. Pelsers MM, Lutgerink JT, Nieuwenhoven FA, Tandon NN, van der Vusse GJ, Arends JW, Hoogenboom HR and Glatz JF (1999). A sensitive immunoassay for rat fatty acid translocase (CD 36) using phage antibodies selected on cell transfectants: abundant presence of fatty acid translocase/CD 36 in cardiac and red skeletal muscle and upregulation in diabetes. Biochem J, 337: 407-414. Peng CF, Wei Y, Levsky JM, McDonald TV, Childs G and Kitsis RN (2002). Microarray analysis of global changes in gene expression during cardiac myocyte differentiation. Physiol Genomics, 9: 145-155. Peng J, Gurantz D, Tran V, Cowling RT and Greenberg BH (2002). Tumor necrosis factor-alpha-induced AT1 receptor upregulation enhances angiotensin II-mediated cardiac fibroblast responses that favor fibrosis. Circ Res, 91: 1119-1126. Perego L, Pizzocri P, Corradi D, Maisano F, Paganelli M, Fiorina P, Barbieri M, Morabito A, Paolisso G, Folli F and Pontiroli AE (2005). Circulating leptin correlates with left ventricular mass in morbid (grade III) obesity before and after weigth loss induced by LAGB. A potential role for leptin in mediating human left ventricular hypertrophy. J Clin Endocrinol Metab, Apr 26 [Electronic pub. ahead of print]. Pessin JE and Bell GI (1992). Mammalian facilitative glucose-transporter family: structure and molecular regulation. Annu Rev Physiol, 54: 911-930.

Bibliography

329

Pestova TV, Borukhov SI and Hellen CU (1998). Eukaryotic ribosomes require initiation factors 1 and 1A to locate initiation codons. Nature, 394: 854-859. Peters NS, Coromilas J, Severs NJ and Wit AL (1997). Disturbed connexin43 gap junction distribution correlates with the location of reentrant circuits in the epicardial border zone of healing canine infarcts that cause ventricular tachycardia. Circulation, 95: 988-996. Peters NS, Green CR, Poole-Wilson PA and Severs NJ (1993). Reduced content of connexin43 gap junctions in ventricular myocardium from hypertrophied and ischemic human hearts. Circulation. 1993, 88: 864-875. Petersen KF and Shulman GI (2002). Pathogenesis of skeletal muscle insulin resistance in type 2 diabetes mellitus. Am J Cardiol, 90: 11G-18G. Petrich BG, Gong X, Lerner DL, Wang X, Brown JH, Saffitz JE and Wang Y (2002). c-Jun N-terminal kinase activation mediates downregulation of connexin43 in cardiomyocytes. Circ Res, 91: 640-647. Petroff MG, Aiello EA, Palomeque J, Salas MA and Mattiazzi A (2000). Subcellular mechanisms of the positive inotropic effect of angiotensin II in cat myocardium. J Physiol, 529: 189-203. Philip-Couderc P, Pathak A, Smih F, Dambrin C, Harmancey R, Buys S, Galinier M, Massabuau P, Roncalli J, Senard JM and Rouet P (2004). Uncomplicated human obesity is associated with a specific cardiac transcriptome: involvement of the Wnt pathway. FASEB J, 18: 1539-1540. Pierce GN, Lockwood MK and Eckhert CD (1989). Cardiac contractile protein ATPase activity in a diet induced model of noninsulin dependent diabetes mellitus. Can J Cardiol, 5: 117-120. Pierce GN, Ramjiawan B, Dhalla NS and Ferrari R (1990). Na+-H+ exchange in cardiac sarcolemmal vesicles isolated from diabetic rats. Am J Physiol, 258: H255-H261. Pierzchalski P, Reiss K, Cheng W, Cirielli C, Kajstura J, Nitahara JA, Rizk M, Capogrossi MC and Anversa P (1997). p53 Induces myocyte apoptosis via the activation of the renin-angiotensin system. Exp Cell Res, 234: 57-65. Pieske B, Kretschmann B, Meyer M, Holubarsch C, Weirich J, Posival H, Minami K, Just H and Hasenfuss G (1995). Alterations in intracellular calcium handling associated with the inverse force-frequency relation in human dilated cardiomyopathy. Circulation, 92: 1169-1178. Pieske B, Maier LS, Bers DM and Hasenfuss G (1999). Ca2+ handling and sarcoplasmic reticulum Ca2+ content in isolated failing and nonfailing human myocardium. Circ Res, 85: 38-46.

Bibliography

330

Pieske B, Sutterlin M, Schmidt-Schweda S, Minami K, Meyer M, Olschewski M, Holubarsch C, Just H and Hasenfuss G (1996). Diminished post-rest potentiation of contractile force in human dilated cardiomyopathy. Functional evidence for alterations in intracellular Ca2+ handling. J Clin Invest, 98: 764-776. Pinto YM, de Smet BG, van Gilst WH, Scholtens E, Monnink S, de Graeff PA and Wesseling H (1993). Selective and time related activation of the cardiac renin-angiotensin system after experimental heart failure: relation to ventricular function and morphology. Cardiovasc Res, 27: 1933-1938. Pinto YM, Pinto-Sietsma SJ, Philipp T, Engler S, Kossamehl P, Hocher B, Marquardt H, Sethmann S, Lauster R, Merker HJ and Paul M (2000). Reduction in left ventricular messenger RNA for transforming growth factor b1 attenuates left ventricular fibrosis and improves survival without lowering blood pressure in the hypertensive TGR(mRen2)27 rat. Hypertension, 36: 747-754. Poelzing S and Rosenbaum DS (2004). Altered connexin43 expression produces arrhythmia substrate in heart failure. Am J Physiol, 287: H1762-H1770. Polontchouk L, Ebelt B, Jackels M and Dhein S (2002). Chronic effects of endothelin 1 and angiotensin II on gap junctions and intercellular communication in cardiac cells. FASEB J, 16: 87-89. Pogwizd SM and Bers DM (2004). Cellular basis of triggered arrhythmias in heart failure. Trends Cardiovasc Med, 14: 61-66. Pravenec M, Simonet L, Kren V, Kunes J, Levan G, Szpirer J, Szpirer C and Kurtz T (1991). The rat renin gene: assignment to chromosome 13 and linkage to the regulation of blood pressure. Genomics, 9: 466-472. Privratsky JR, Wold LE, Sowers JR, Quinn MT and Ren J (2003). AT1 blockade prevents glucose-induced cardiac dysfunction in ventricular myocytes: role of the AT1 receptor and NADPH oxidase. Hypertension, 42: 206-12. Quackenbush J (2001). Computational analysis of microarray data. Nat Rev Genet, 2: 418-427. Raimondi L, De Paoli P, Mannucci E, Lonardo G, Sartiani L, Banchelli G, Pirisino R, Mugelli A and Cerbai E (2004). Restoration of cardiomyocyte functional properties by angiotensin II receptor blockade in diabetic rats. Diabetes, 53: 1927-1933. Ramaraj P, Kessler SP, Colmenares C and Sen GC (1998). Selective restoration of male fertility in mice lacking angiotensin-converting enzymes by sperm-specific expression of the testicular isozyme. J Clin Invest, 102: 371-378. Ramasamy R and Schaefer S (1999). Inhibition of Na+-H+ exchanger protects diabetic and non-diabetic hearts from ischemic injury: insight into altered susceptibility of diabetic hearts to ischemic injury. J Mol Cell Cardiol, 31: 785-797.

Bibliography

331

Randhawa AK and Singal PK (1992). Pressure overload-induced cardiac hypertrophy with and without dilation. J Am Coll Cardiol, 20: 1569-1575. Rannou F, Dambrin G, Marty I, Carre F, Trouve P, Lompre AM and Charlemagne D (1996). Expression of the cardiac ryanodine receptor in the compensated phase of hypertrophy in rat heart. Cardiovasc Res, 32: 258-265. Ranu HK, Terracciano CM, Davia K, Bernobich E, Chaudhri B, Robinson SE, Bin Kang Z, Hajjar RJ, MacLeod KT and Harding SE (2002). Effects of Na(+)/Ca(2+)-exchanger overexpression on excitation-contraction coupling in adult rabbit ventricular myocytes. J Mol Cell Cardiol, 34: 389-400. Rapp JP, Wang SM and Dene H (1989). A genetic polymorphism in the renin gene of Dahl rats cosegregates with blood pressure. Science, 243: 542-544. Reaven GM, Hollenbeck C, Jeng CY, Wu MS and Chen YD (1988). Measurement of plasma glucose, free fatty acid, lactate and insulin for 24 h in patients in NIDDM. Diabetes, 37: 1020-1024. Reaven GM (1991). Insulin resistance, hyperinsulinaemia, hypertriglyceridaemia, and hypertension: parallels between human disease and rodent models. Diabetes Care, 14: 195-202. Reddy RK, Mao C, Baumeister P, Austin RC, Kaufman RJ and Lee AS (2003). Endoplasmic reticulum chaperone protein GRP78 protects cells from apoptosis induced by topoisomerase inhibitors: role of ATP binding site in suppression of caspase-7 activation. J Biol Chem, 278: 20915-20924. Reinecke H, Vetter R and Drexler H (1997). Effects of alpha-adrenergic stimulation on the sarcolemmal Na+/Ca2+-exchanger in adult rat ventricular cardiocytes. Cardiovasc Res, 36: 216-222. Remme W, Boccanelli A, Cline C, Cohen-Solal A, Dietz R, Hobbs R, Keukelaar K, Sendon JL, Macarie C, McMurray J, Rauch B, Ruzyllo W, Zannad F and SHAPE Study (2004). Increasing awareness and perception of heart failure in Europe and improving care--rationale and design of the SHAPE Study. Cardiovasc Drugs Ther, 18: 153-159. Ren J and Davidoff AJ (1997). Diabetes rapidly induces contractile dysfunctions in isolated ventricular myocytes. Am J Physiol, 272: H148-H158. Ren J, Walsh MF, Hamaty M, Sowers JR and Brown RA (1999). Augmentation of the inotropic response to insulin in diabetic rat hearts. Life Sci, 65: 369-380. Reuter H, Henderson SA, Han T, Ross RS, Goldhaber JI and Philipson KD (2002). The Na+-Ca2+ exchanger is essential for the action of cardiac glycosides. Circ Res, 90: 305-308.

Bibliography

332

Richter-Cook NJ, Dever TE, Hensold JO, Merrick WC (1998). Purification and characterization of a new eukaryotic protein translation factor. Eukaryotic initiation factor 4H. J Biol Chem, 273: 7579-7587. Robillon JF, Sadoul JL, Benmerabet S, Joly-Lemoine L, Fredenrich A and Canivet B (1999). Assessment of cardiac arrhythmic risk in diabetic patients using QT dispersion abnormalities. Diabetes Metab, 25: 419-423. Robinson MJ and Cobb MH (1997). Mitogen-activated protein kinase pathways. Curr Opin Cell Biol, 9: 180-186. Rodrigues B, Cam MC, McNeill JH (1998). Metabolic disturbances in diabetic cardiomyopathy. Mol Cell Biochem, 180: 53-57. Rodrigues B and McNeill JH (1992). The diabetic heart: causes for the development of a cardiomyopathy. Cardiovasc Res, 26: 913-922. Romppanen H, Marttila M, Magga J, Vuolteenaho O, Kinnunen P, Szokodi I and Ruskoaho H (1997). Adrenomedullin gene expression in the rat heart is stimulated by acute pressure overload: blunted effect in experimental hypertension. Endocrinology, 138: 2636-2639. Ron D, Brasier A and Habener J (1990). Transcriptional regulation of hepatic angiotensinogen gene expression by the acute-phase response. Mol Cell Endocrinol, 74: C97-C104. Rosenblatt-Velin N, Lepore MG, Cartoni C, Beermann F and Pedrazzini T (2005). FGF-2 controls the differentiation of resident cardiac precursors into functional cardiomyocytes. J Clin Invest, 115: 1724-1733. Rosenkranz S (2004). TGF-beta1 and angiotensin networking in cardiac remodeling. Cardiovasc Res, 3: 423-432. Roskes EJ, Boughman JA, Schwartz S and Cohen MM (1990). Congenital cardiovascular malformations (CCVM) and structural chromosome abnormalities: a report of 9 cases and literature review. Clin Genet, 38: 198-210. Ruiz J, Blanche H, Cohen N, Velho G, Cambien F, Cohen D, Passa P and Froguel P (1994). Insertion/deletion polymorphism of the angiotensin-converting enzyme gene is strongly associated with coronary heart disease in non-insulin-dependent diabetes mellitus. Proc Natl Acad Sci U S A, 91: 3662-3665. Russell LK, Finck BN and Kelly DP (2005). Mouse models of mitochondrial dysfunction and heart failure. J Mol Cell Cardiol, 38: 81-91. Rysa J, Leskinen H, Ilves M and Ruskoaho H (2004). Distinct upregulation of extracellular matrix genes in transition from hypertrophy to hypertensive heart failure. Hypertension, 45: 927-933.

Bibliography

333

Saavedra JM (2005). Brain angiotensin II: new developments, unanswered questions and therapeutic opportunities. Cell Mol Neurobiol, 25: 485-512. Saavedra JM, Viswanathan M and Shigematsu K (1993). Localization of angiotensin AT1 receptors in the rat heart conduction system. Eur J Pharmacol, 235: 301-303. Sadoshima J and Izumo S (1993a). Mechanotransduction in stretch-induced hypertrophy of cardiac myocytes. J Recept Res, 13: 777–794. Sadoshima J, Xu Y, Slayter HS and Izumo S (1993). Autocrine release of angiotensin II mediates stretch-induced hypertrophy of cardiac myocytes in vitro. Cell 75 (1993), pp. 977–984. Saez JC, Berthoud VM, Branes MC, Martinez AD and Beyer EC (2003). Plasma membrane channels formed by connexins: their regulation and functions. Physiol Rev, 83: 1359-1400. Saito S, Hiroi Y, Zou Y, Aikawa R, Toko H, Shibasaki F, Yazaki Y, Nagai R and Komuro I (2000). beta-Adrenergic pathway induces apoptosis through calcineurin activation in cardiac myocytes. J Biol Chem, 275: 34528-34533. Sakata Y, Hoit BD, Liggett SB, Walsh RA and Dorn GW 2nd (1998). Decompensation of pressure-overload hypertrophy in G alpha q-overexpressing mice. Circulation, 97: 1488-1495. Sakurai K, Norota I, Tanaka H, Kubota I, Tomoike H and Endoh M (2002). Negative inotropic effects of angiotensin II, endothelin-1 and phenylephrine in indo-1 loaded adult mouse ventricular myocytes. Life Sci, 70: 1173-1184. Salinovich O and Montelaro RC (1986). Reversible staining and peptide mapping of proteins transferred to nitrocellulose after separation by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Anal Biochem, 156: 341-347. Sampaio WO, Nascimento AA and Santos RAS (2003). Systemic and regional hemodynamic effects of angiotensin-(1-7) in rats. Am J Physiol, 284: H1985-H1994. Samyn ME, Petershack JA, Bedell KA, Mathews MS and Segar JL (1998). Ontogeny and regulation of cardiac angiotensin types 1 and 2 receptors during fetal life in sheep. Pediatr Res, 44: 323-329. Sandberg K, Ji H, Clark AJ, Shapira H and Catt KJ (1992). Cloning and expression of a novel angiotensin II receptor subtype. J Biol Chem, 267: 9455-9458. Sanders J, Brandsma M, Janssen GM, Dijk J and Moller W (1996). Immunofluorescence studies of human fibroblasts demonstrate the presence of the complex of elongation factor-1 beta gamma delta in the endoplasmic reticulum. J Cell Sci, 109: 1113-1117.

Bibliography

334

Sandmann S, Yu M, Kaschina E, Blume A, Bouzinova E, Aalkjaer C and Unger T (2001). Differential effects of angiotensin AT1 and AT2 receptors on the expression translation and function of the Na+-H+ exchanger and Na+-HCO3- symporter in the rat heart after myocardial infarction. J Am Coll Cardiol, 37: 2154-2165. Sandri C, Di Lisi R, Picard A, Argentini C, Calabria E, Myklak K, Scartezzini P and Schiaffino S (2004). Heart morphogenesis is not affected by overexpression of the Sh3bgr gene mapping to the Down syndrome heart critical region. Hum Genet, 114: 517-519. Sanghani SP, Quinney SK, Fredenburg TB, Davis WI, Murry DJ and Bosron WF (2004). Hydrolysis of irinotecan and its oxidative metabolites, 7-ethyl-10-[4-N-(5-aminopentanoic acid)-1-piperidino] carbonyloxycamptothecin and 7-ethyl-10-[4-(1-piperidino)-1-amino]-carbonyloxycamptothecin, by human carboxylesterases CES1A1, CES2, and a newly expressed carboxylesterase isoenzyme, CES3. Drug Metab Dispos, 32: 505-511. Sato T, Haimovici R, Kao R, Li AF and Roy S (2002). Downregulation of connexin 43 expression by high glucose reduces gap junction activity in microvascular endothelial cells. Diabetes, 51: 1565-1571. Schachter F, Faure-Delanef L, Guenot F, Rouger H, Froguel P, Lesueur-Ginot L and Cohen D (1994). Genetic associations with human longevity at the APOE and ACE loci. Nat Genet, 6: 29-32. Schaffer JE and Lodish HF (1994). Expression cloning and characterization of a novel adipocite long-chain fatty acid transport protein. Cell, 79: 427-436. Schaffer SW, Ballard-Croft C, Boerth S and Allo SN (1997). Mechanisms underlying depressed Na+/Ca2+ exchanger activity in the diabetic heart. Cardiovasc Res, 34: 129-136. Schaffer SW, Tan BH and Wilson GL (1985). Development of a cardiomyopathy in a model of noninsulin-dependent diabetes. Am J Physiol, 248: H179-H185. Schaffer SW and Wilson GL (1993). Insulin resistance and mechanical dysfunction in hearts of Wistar rats with streptozotocin-induced non-insulin-dependent diabetes mellitus. Diabetologia, 36: 195-199. Schaub MC, Hefti MA, Harder BA and Eppenberger HM (1997). Various hypertrophic stimuli induce distinct phenotypes in cardiomyocytes. J Mol Med, 75: 901-920. Schena M, Shalon D, Davis RW and Brown PO (1995). Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science, 270: 467-70. Schillaci G, Vaudo G, Pasqualini L, Reboldi G, Porcellati C and Verdecchia P (2002). Left ventricular mass and systolic dysfunction in essential hypertension. J Hum Hypertens, 16: 117-122.

Bibliography

335

Schlange T, Andree B, Arnold HH and Brand T (2000). BMP2 is required for early heart development during a distinct time period. Mech Dev, 91: 259-270. Schlotthauer K and Bers DM (2000). Sarcoplasmic reticulum Ca(2+) release causes myocyte depolarization. Underlying mechanism and threshold for triggered action potentials. Circ Res, 87: 774-780. Schmieder RE and Messerli FH (2000). Hypertension and the heart. J Hum Hypertens, 14: 597-604. Schotten U, Koenigs B, Rueppel M, Schoendube F, Boknik P, Schmitz W and Hanrath P (1999). Reduced myocardial sarcoplasmic reticulum Ca2+-ATPase protein expression in compensated primary and secondary human cardiac hypertrophy. J Mol Cell Cardiol, 31: 1483-1494. Schunkert H, Dzau VJ, Tang SS, Hirsch AT, Apstein CS and Lorell BH (1990). Increased rat cardiac angiotensin converting enzyme activity and mRNA expression in pressure overload left ventricular hypertrophy. J Clin Invest, 86: 1913-1920. Schunkert H, Hense HW, Holmer SR, Stender M, Perz S, Keil U, Lorell BH and Riegger GAJ (1994). Association between a deletion polymorphism of the angiotensin-converting-enzyme gene and left ventricular hypertrophy. New Eng J Med, 330: 1634-1638. Schweizer L and Varmus H (2003). Wnt/Wingless signaling through beta-catenin requires the function of both LRP/Arrow and frizzled classes of receptors. BMC Cell Biol, 4: 4. Scott AL, Chang RS, Lotti VJ and Siegl PK (1992). Cardiac angiotensin receptors: effects of selective angiotensin II receptor antagonists, DUP 753 and PD 121981, in rabbit heart. J Pharmacol Exp Ther, 261: 931-935. Sechi LA, Griffin CA, Grady EF, Kalinyak JE and Schambelan M (1992). Characterization of angiotensin II receptor subtypes in rat heart. Circ Res, 71: 1482-1489. Sechi LA, Griffin CA and Schambelan M (1994). The cardiac renin-angiotensin system in STZ-induced diabetes. Diabetes, 43: 1180-1184. Seedorf U, Raabe M, Ellinghaus P, Kannenberg F, Fobker M, Engel T, Denis S, Wouters F, Wirtz KW, Wanders RJ, Maeda N and Assmann G (1998). Defective peroxisomal catabolism of branched fatty acyl coenzyme A in mice lacking the sterol carrier protein-2/sterol carrier protein-x gene function. Genes Dev, 12: 1189-1201. Sehl PD, Tai JT, Hillan KJ, Brown LA, Goddard A, Yang R, Jin H and Lowe DG (2000). Application of cDNA microarrays in determining molecular phenotype in cardiac growth, development, and response to injury. Circulation, 101: 1990-1999.

Bibliography

336

Seidman JG and Seidman C (2001). The genetic basis for cardiomyopathy: from mutation identification to mechanistic paradigms. Cell, 104: 557–567. Sekine T, Kusano H, Nishimaru K, Tanaka Y, Tanaka H and Shigenobu K (1999). Developmental conversion of inotropism by endothelin I and angiotensin II from positive to negative in mice. Eur J Pharmacol, 374: 411-415. Seltzer A, Bregonzio C, Armando I, Baiardi G and Saavedra JM (2004). Oral administration of an AT1 receptor antagonist prevents the central effects of angiotensin II in spontaneously hypertensive rats. Brain Res, 1028: 9-18. Semeniuk LM, Kryski AJ and Severson DL (2002). Echocardiographic assessment of cardiac function in diabetic db/db and transgenic db/db-hGLUT4 mice. Am J Physiol, 283: H976-H982. Senzaki H, Gluzband YA, Pak PH, Crow MT, Janicki JS and Kass DA (1998). Synergistic exacerbation of diastolic stiffness from short-term tachycardia-induced cardiodepression and angiotensin II. Circ Res, 82: 503-512. Serneri GG, Boddi M, Cecioni I, Vanni S, Coppo M, Papa ML, Bandinelli B, Bertolozzi I, Polidori G, Toscano T, Maccherini M and Modesti PA (2001). Cardiac angiotensin II formation in the clinical course of heart failure and its relationship with left ventricular function. Circ Res, 88: 961-968. Shah BH, Yesilkaya A, Olivares-Reyes JA, Chen HD, Hunyady L and Catt KJ (2004). Differential pathways of angiotensin II-induced extracellularly regulated kinase 1/2 phosphorylation in specific cell types: role of heparin-binding epidermal growth factor. Mol Endocrinol, 18: 2035-2048. Shao Q, Ren B, Zarain-Herzberg A, Ganguly PK and Dhalla NS (1999). Captopril treatment improves the sarcoplasmic reticular Ca2+ transport in heart failure due to myocardial infarction. J Mol Cell Cardiol, 31: 1663-1672. She P, Shiota M, Shelton KD, Chalkley R, Postic C and Magnuson MA (2000). Phosphoenolpyruvate carboxykinase is necessary for the integration of hepatic energy metabolism. Mol Cell Biol, 20: 6508-6517. Shehadeh A and Regan TJ (1995). Cardiac consequences of diabetes mellitus. Clin Cardiol, 18: 301-305. Sheikh MS, Fernandez-Salas E, Yu M, Hussain A, Dinman JD, Peltz SW, Huang Y and Fornace AJ Jr (1999). Cloning and characterization of a human genotoxic and endoplasmic reticulum stress-inducible cDNA that encodes translation initiation factor 1(eIF1(A121/SUI1)). J Biol Chem, 274: 16487-16493. Shimizu M, Umeda K, Sugihara N, Yoshio H, Ino H, Takeda R, Okada Y and Nakanishi I (1993). Collagen remodeling in myocardia of patients with diabetes. J Clin Pathol, 46: 32-36.

Bibliography

337

Shoemaker CA, Pungliya M, Sao Pedro MA, Ruiz C, Alvarez SA, Ward M, Ryder EF and Krushkal J (2001). Computational methods for single-point and multipoint analysis of genetic variants associated with a simulated complex disorder in a general population. Genet Epidemiol, 21 (Suppl 1): S738-S745. Shyu KG, Chen CC, Wang BW, Kuan P (2001). Angiotensin II receptor antagonist blocks the expression of connexin43 induced by cyclical mechanical stretch in cultured neonatal rat cardiac myocytes. J Mol Cell Cardiol, 33: 691-698. Singh CR, He H, Ii M, Yamamoto Y and Asano K (2004). Efficient incorporation of eukaryotic initiation factor 1 into the multifactor complex is critical for formation of functional ribosomal preinitiation complexes in vivo. J Biol Chem, 279: 31910-31920. Singh K, Sirokman G, Communal C, Robinson KG, Conrad CH, Brooks WW, Bing OH and Colucci WS (1999). Myocardial osteopontin expression coincides with the development of heart failure. Hypertension, 33: 663-670. Sipido KR, Volders PG, Schoenmakers M, De Groot SH, Verdonck F and Vos MA (2002). Role of the Na/Ca exchanger in arrhythmias in compensated hypertrophy. Ann N Y Acad Sci, 976: 438-445. Skaletz-Rorowski A, Pinkernell K, Sindermann JR, Schriever C, Muller JG, Eschert H and Breithardt G (2004). Angiotensin AT1 receptor upregulates expression of basic fibroblast growth factor, basic fibroblast growth factor receptor and coreceptor in human coronary smooth muscle cells. Basic Res Cardiol, 99: 272-278. Skolnick RL, Litwin SE, Barry WH and Spitzer KW (1998). Effect of Ang II on pHi, [Ca2+]I, and contraction in rabbit ventricular myocytes from infarcted hearts. Am J Physiol, 275: H1788-H1797. Skrabal CA, Thompson LO, Southard RE, Joyce DL, Noon GP, Loebe M and Youker KA (2004). Interaction between isolated human myocardial mast cells and cultured fibroblasts. J Surg Res, 118: 66-70. Skurat AV and Dietrich AD (2004). Phosphorylation of Ser640 in muscle glycogen synthase by DYRK family protein kinases. J Biol Chem, 279:2490-2498. Smith WM (1985). Epidemiology of congestive heart failure. Am J Cardiol, 55: 3A-8A. Smyth GK, Yang YH and Speed T (2003). Statistical issues in cDNA microarray data analysis. Methods Mol Biol, 224: 111-136. Snabaitis AK, Hearse DJ and Avkiran M (2002). Regulation of sarcolemmal Na(+)/H(+) exchange by hydrogen peroxide in adult rat ventricular myocytes. Cardiovasc Res, 53: 470-480. Song K, Allen AM, Paxinos G and Mendelsohn FA (1992). Mapping of angiotensin II receptor subtype heterogeneity in rat brain. J Comp Neurol, 316: 467-484.

Bibliography

338

Spedding M and Mir AK (1987). Direct activation of Ca2+ channels by palmitoyl carnitine, a putative endogenous ligand. Br J Pharmacol, 92: 457-468. Spinale FG, Coker ML, Bond BR and Zellner JL (2000). Myocardial matrix degradation and metalloproteinase activation in the failing heart: a potential therapeutic target. Cardiovasc Res, 46: 225-238. Spirito P, Seidman CE, McKenna WJ and Maron BJ (1997). The management of hypertrophic cardiomyopathy. N Engl J Med, 336: 775-785. Stamler J, Vaccaro O, Neaton JD and Wentworth D (1993). Diabetes, other risk factors, and 12-yr cardiovascular mortality for men screened in the Multiple Risk Factor Intervention Trial. Diabetes Care, 16: 434-444. Stanton LW, Garrard LJ, Damm D, Garrick BL, Lam A, Kapoun AM, Zheng Q, Protter AA, Schreiner GF and White RT (2000). Altered patterns of gene expression in response to myocardial infarction. Circ Res, 86: 939-945. Stawowy P, Margeta C, Kallisch H, Seidah NG, Chretien M, Fleck E and Graf K (2004). Regulation of matrix metalloproteinase MT1-MMP/MMP-2 in cardiac fibroblasts by TGF-beta1 involves furin-convertase. Cardiovasc Res, 63: 87-97. Stenbit AE, Katz EB, Chatham JC, Geenen DL, Factor SM, Weiss RG, Tsao TS, Malhotra A, Chacko VP, Ocampo C, Jelicks LA and Charron MJ (2000). Preservation of glucose metabolism in hypertrophic GLUT4-null hearts. Am J Physiol, 279: H313-H318. Stenbit AE, Tsao TS, Li J, Burcelin R, Geenen DL, Factor SM, Houseknecht K, Katz EB and Charron MJ (1997). GLUT4 heterozygous knockout mice develop muscle insulin resistance and diabetes. Nat Med, 3: 1096-1101. Stewart S, MacIntyre K, Capewell S and McMurray JJ (2003). Heart failure and the aging population: an increasing burden in the 21st century? Heart, 89: 49-53. Stottmann RW, Choi M, Mishina Y, Meyers EN and Klingensmith J (2004). BMP receptor IA is required in mammalian neural crest cells for development of the cardiac outflow tract and ventricular myocardium. Development, 131: 2205-2218. Stronetta RL, Hawelu-Johnson CL, Guyenet PG and Lynch KR (1988). Astrocytes synthesize angiotensinogen in brain. Science, 242: 1444-1446. Studer R, Reinecke H, Bilger J, Eschenhagen T, Bohm M, Hasenfuss G, Just H, Holtz J and Drexler H (1994). Gene expression of the cardiac Na+/Ca2+ exchanger in end-stage human heart failure. Circ Res, 75: 443-453. Stull LB, Leppo MK, Marban E and Janssen PML (2002). Physiological determinants of contractile force generation and calcium handling in mouse myocardium. J Mol Cell Cardiol, 34: 1367-1376.

Bibliography

339

Su HM, Moser AB, Moser HW and Watkins PA (2001). Peroxisomal straight-chain Acyl-CoA oxidase and D-bifunctional protein are essential for the retroconversion step in docosahexaenoic acid synthesis. J Biol Chem, 276: 38115-38120. Sugaya T, Nishimatsu S, Tanimoto K, Takimoto E, Yamagishi T, Imamura K, Goto S, Imaizumi K, Hisada Y, Otsuka A, et al. (1995). Angiotensin II type 1a receptor-deficient mice with hypotension and hyperreninemia. J Biol Chem, 270: 18719-18722. Sugi Y and Markwald RR (2003). Endodermal growth factors promote endocardial precursor cell formation from precardiac mesoderm. Dev Biol, 263: 35-49. Sugino H, Ozono R, Kurisu S, Matsuura H, Ishida M, Oshima T, Kambe M, Teranishi Y, Masaki H and Matsubara H (2001). Apoptosis is not increased in myocardium overexpressing type 2 angiotensin II receptor in transgenic mice. Hypertension, 37: 1394-1398. Suko J, Vogel JHK and Chidsey CA (1970). Intracellular calcium and myocardial contractility. III. Reduced calcium uptake and ATPase of the sarcoplasmic reticulum fraction prepared from chronically failing calf hearts. Circ Res, 27: 235-247. Sumida Y, Umemura S, Tamura K, Kihara M, Kobayashi S, Ishigami T, Yabana M, Nyui N, Ochiai H, Fukamizu A, Miyazaki H, Murakami K and Ishii M (1998). Increased cardiac angiotensin II receptors in angiotensinogen-deficient mice. Hypertension, 31: 45-49. Sun D, Nguyen N, Degrado TR, Schwaiger M and Brosius FC (1994). Ischemia induces translocation of the insulin-responsive glucose transporter GLUT4 to the plasma membrane of cardiac myocytes. Circulation, 89: 793-798. Sun HY, Wang NP, Halkos ME, Kerendi F, Kin H, Wang RX, Guyton RA and Zhao ZQ (2004). Involvement of Na+/H+ exchanger in hypoxia/re-oxygenation-induced neonatal rat cardiomyocyte apoptosis. Eur J Pharmacol, 486: 121-131. Sussman MA, Welch S, Cambon N, Klevitsky R, Hewett TE, Price R, Witt SA and Kimball TR (1998a). Myofibril degeneration caused by tropomodulin overexpression leads to dilated cardiomyopathy in juvenile mice. J Clin Invest, 101: 51-61. Sussman MA, Lim HW, Gude N, Taigen T, Olson EN, Robbins J, Colbert MC, Gualberto A, Wieczorek DF and Molkentin JD (1998b). Prevention of cardiac hypertrophy in mice by calcineurin inhibition. Science, 281: 1690-1693. Sussman MA, Welch S, Walker A, Klevitsky R, Hewett TE, Witt SA, Kimball TR, Price R, Lim HW and Molkentin JD (2000). Hypertrophic defect unmasked by calcineurin expression in asymptomatic tropomodulin overexpressing transgenic mice. Cardiovasc Res, 46: 90-101.

Bibliography

340

Suurmeijer AJ, Clement S, Francesconi A, Bocchi L, Angelini A, Van Veldhuisen DJ, Spagnoli LG, Gabbiani G and Orlandi A (2003). Alpha-actin isoform distribution in normal and failing human heart: a morphological, morphometric, and biochemical study. J Pathol, 199: 387-397. Swanson GN, Hanesworth JM, Sardinia MF, Coleman JK, Wright JW, Hall KL, Miller-Wing AV, Stobb JW, Cook VI, Harding EC, et al. (1992). Discovery of a distinct binding site for angiotensin II (3-8), a putative angiotensin IV receptor. Regul Pept, 40: 409-419. Swynghedauw B (1999). Molecular mechanisms of myocardial remodeling. Physiol Rev, 79: 215-62. Tabuchi M, Yoshimori T, Yamaguchi K, Yoshida T and Kishi F (2000). Human NRAMP2/DMT1, which mediates iron transport across endosomal membranes, is localized to late endosomes and lysosomes in HEp-2 cells. J Biol Chem, 275: 22220-22228. Taegtmeyer H and Overturf L (1988). Effects of moderate hypertension on cardiac function and metabolism in the rabbit. Hypertension, 11: 416-426. Tagawa H, Koide M, Sato H, Zile MR, Carabello BA and Cooper G 4th (1998). Cytoskeletal role in the transition from compensated to decompensated hypertrophy during adult canine left ventricular pressure overloading. Circ Res, 82: 751-761. Takahashi E, Abe J, Gallis B, Aebersold R, Spring DJ, Krebs EG and Berk BC (1999). p90(RSK) is a serum-stimulated Na+/H+ exchanger isoform-1 kinase. Regulatory phosphorylation of serine 703 of Na+/H+ exchanger isoform-1. J Biol Chem, 274: 20206-20214. Takewaki S, Kuro-o M, Hiroi Y, Yamazaki T, Noguchi T, Miyagishi A, Nakahara K, Aikawa M, Manabe I, Yazaki Y et al. (1995). Activation of Na+-H+ antiporter (NHE-1) gene expression during growth, hypertrophy and proliferation of the rabbit cardiovascular system. J Mol Cell Cardiol, 27: 729-742. Takeyama D, Kagaya Y, Yamane Y, Shiba N, Chida M, Takahashi T, Ido T, Ishide N and Takishima T (1995). Effects of chronic right ventricular pressure overload on myocardial glucose and free fatty acid metabolism in the conscious rat. Cardiovasc Res, 29: 763-767. Talukder MA and Endoh M (1997). Pharmacological differentiation of synergistic contribution of L-type Ca2+ channels and Na+/H+ exchange to the positive inotropic effect of phenylephrine, endothelin-3 and angiotensin II in rabbit ventricular myocardium. Naunyn Schmiedebergs Arch Pharmacol, 355: 87-96. Tamaddon HS, Vaidya D, Simon AM, Paul DL, Jalife J and Morley GE (2000). High-resolution optical mapping of the right bundle branch in connexin40 knockout mice reveals slow conduction in the specialized conduction system. Circ Res, 87: 929-936.

Bibliography

341

Tambara K, Fujita M, Miyamoto S, Doi K, Nishimura K and Komeda M (2004). Pericardial fluid level of heart-type cytoplasmic fatty acid-binding protein (H-FABP) is an indicator of severe myocardial ischemia. Int J Cardiol, 93: 281-284. Tanabe A, Naruse M, Arai K, Naruse K, Yoshimoto T, Seki T, Imaki T, Miyazaki H, Zeng ZP, Demura R and Demura H (1998). Gene expression and roles of angiotensin II type 1 and type 2 receptors in human adrenals. Horm Metab Res, 30: 490-495. Tanaka M, Ito H, Adachi S, Akimoto H, Nishikawa T, Kasajima T, Marumo F and Hiroe M (1994). Hypoxia induces apoptosis with enhanced expression of Fas antigen messenger RNA in cultured neonatal rat cardiomyocytes. Circ Res, 75: 426-433. Tanimoto K, Sugiyama F, Goto Y, Ishida J, Takimoto E, Yagami K, Fukamizu A and Murakami K (1994). Angiotensinogen-deficient mice with hypotension. J Biol Chem, 269: 31334-31337. Tanonaka K, Yoshida H, Toga W, Furuhama K and Takeo S (2001). Myocardial heat shock proteins during the development of heart failure. Biochem Biophys Res Commun, 283: 520-525. Tavaria M, Gabriele T, Anderson RL, Mirault ME, Baker E, Sutherland G and Kola I (1995). Localization of the gene encoding the human heat shock cognate protein, HSP73, to chromosome 11. Genomics, 29: 266-268. Terracciano CMN, Philipson KD and MacLeod KT (2001). Overexpression of the Na+/Ca2+ exchanger and inhibition of the sarcoplasmic reticulum Ca2+-ATPase in ventricular myocytes from transgenic mice. Cardiovasc Res, 49: 38-47. Terracio L, Rubin K, Gullberg D, Balog E, Carver W, Jyring R and Borg TK (1991). Expression of collagen binding integrins during cardiac development and hypertrophy. Circ Res, 68: 734-744. Terrenoire C, Clancy CE, Cormier JW, Sampson KJ and Kass RS (2005). Autonomic control of cardiac action potentials: role of potassium channel kinetics in response to sympathetic stimulation. Circ Res, 96: e25-e34. Teshima Y, Takahashi N, Saikawa T, Hara M, Yasunaga S, Hidaka S and Sakata T (2000). Diminished expression of sarcoplasmic reticulum Ca2+-ATPase and ryanodine sensitive Ca2+ Channel mRNA in streptozotocin-induced diabetic rat heart. J Mol Cell Cardiol, 32: 655-664. Tewksbury DA and Dart RA (1982). High molecular weight angiotensinogen levels in hypertensive pregnant women. Hypertension, 4: 729-734. Thai MV, Guruswamy S, Cao KT, Pessin JE and Olson AL (1998). Myocyte enhancer factor 2 (MEF2)-binding site is required for GLUT4 gene expression in transgenic mice. Regulation of MEF2 DNA binding activity in insulin-deficient diabetes. J Biol Chem, 273: 14285-14292.

Bibliography

342

Tham DM, Martin-McNulty B, Wang YX, Wilson DW, Vergona R, Sullivan ME, Dole W and Rutledge JC (2002). Angiotensin II is associated with activation of NF-kappaB-mediated genes and downregulation of PPARs. Physiol Genomics, 11: 21-30. Thomas WG, Brandenburger Y, Auteliano DJ, Pham T, Qian H and Hannan RD (2002). Adenoviral-directed expression of the type 1A angiotensin receptor promotes cardiomyocyte hypertrophy via transactivation of the epidermal growth factor receptor. Circ Res, 90: 135-142. Tian R, Pham M and Abel ED (1999). Cardiac glucose transporter deficiency increases the susceptibility of the heart to ischemic injury. Diabetes, 100 (Suppl I): I-118 (Abstract). Tiemann K, Weyer D, Djoufack PC, Ghanem A, Lewalter T, Dreiner U, Meyer R, Grohe C and Fink KB (2003). Increasing myocardial contraction and blood pressure in C57BL/6 mice during early postnatal development. Am J Physiol, 284: H464-H474. Till M, Kolter T and Eckel J (1997). Molecular mechanisms of contraction-induced translocation of GLUT4 in isolated cardiomyocytes. Am J Cardiol, 80: 85A-89A. Tingleff J, Munch M, Jakobsen TJ, Torp-Pedersen C, Olsen ME, Jensen KH, Jorgensen T and Kirchoff M (1996). Prevalence of left ventricular hypertrophy in a hypertensive population. Eur Heart J, 17: 143-149. Tipnis SR, Hooper NM, Hyde R, Karran E, Christie G and Turner AJ (2000). A human homolog of angiotensin-converting enzyme: cloning and functional expression as a captopril-insensitive carboxypeptidase. J Biol Chem, 275: 33238-33243. Todd JA (1999). From genome to aetiology in a multifactorial disease, type I diabetes. Bioessays, 21: 164-174. Tom B, de Vries R, Saxena PR and Danser AH (2001). Bradykinin potentiation by angiotensin-(1-7) and ACE inhibitors correlates with ACE C- and N-domain blockade. Hypertension, 38: 95-99. Touret N, Furuya W, Forbes J, Gros P and Grinstein S (2003). Dynamic traffic through the recycling compartment couples the metal transporter Nramp2 (DMT1) with the transferrin receptor. J Biol Chem, 278: 25548-25557. Tournier C, Hess P, Yang DD, Xu J, Turner TK, Nimnual A, Bar-Sagi D, Jones SN, Flavell RA and Davis RJ (2000). Requirement of JNK for stress-induced activation of the cytochrome c-mediated death pathway. Science, 288: 870-874. Touyz RM, Fareh J, Thibault G, Tolloczko B, Lariviere R and Schiffrin EL (1996). Modulation of Ca2+ transients in neonatal and adult rat cardiomyocytes by angiotensin II and endothelin-1. Am J Physiol, 270: H857-H868.

Bibliography

343

Touyz RM, Deng LY, He G, Wu XH and Schiffrin EL (1999). Angiotensin II stimulates DNA and protein synthesis in vascular smooth muscle cells from human arteries: role of extracellular signal-regulated kinases. J Hypertens, 17: 907-916. Touyz RM and Schiffrin EL (2001). Increased generation of superoxide by angiotensin II in smooth muscle cells from resistance arteries of hypertensive patients: role of phospholipase D-dependent NAD(P)H oxidase-sensitive pathways. J Hypertens, 19: 1245-1254. Trongtorsak P, Morgan TO and Delbridge LM (2003). Combined renin-angiotensin system blockade and dietary sodium restriction impairs cardiomyocyte contractility. J Renin Angiotensin Aldosterone Syst, 4: 213-219. Tronik D, Dreyfus M, Babinet C and Rougeon F (1987). Regulated expression of the Ren-2 gene in transgenic mice derived from parental strains carrying only the Ren-1 gene. EMBO J, 6: 983-987. Troughton RW, Lewis LK, Yandle TG, Richards AM and Nicholls MG (2000). Hemodynamic, hormone, and urinary effects of adrenomedullin infusion in essential hypertension. Hypertension, 36: 588-593. Tse J, Huang MW, Leone RJ, Weiss HR, He YQ and Scholz PM (2000). Down regulation of myocardial beta1-adrenoceptor signal transduction system in pacing-induced failure in dogs with aortic stenosis-induced left ventricular hypertrophy. Mol Cell Biochem, 205: 67-73. Tsioufis C, Stefanadis C, Toutouza M, Kallikazaros I, Toutouzas K, Tousoulis D, Pitsavos C, Papademetriou V and Toutouzas P (2002). Microalbuminuria is associated with unfavourable cardiac geometric adaptations in essential hypertensive subjects. J Hum Hypertens, 16: 249-254. Tsuchida K, Watajima H and Otomo S (1994). Calcium current in rat diabetic ventricular myocytes. Am J Physiol, 267: H2280-H2289. Tsuchida K and Watajima H (1997). Potassium currents in ventricular myocytes from genetically diabetic rats. Am J Physiol, 273: E695-E700. Tsuruda T, Jougasaki M, Boerrigter G, Costello-Boerrigter LC, Cataliotti A, Lee SC, Salz-Gilman L, Nordstrom LJ, McGregor CG and Burnett JC (2003). Ventricular adrenomedullin is associated with myocyte hypertrophy in human transplanted heart. Regul Pept, 112: 161-166. Ueno S, Ohki R, Hashimoto T, Takizawa T, Takeuchi K, Yamashita Y, Ota J, Choi YL, Wada T, Koinuma K, Yamamoto K, Ikeda U, Shimada K and Mano H (2003). DNA microarray analysis of in vivo progression mechanism of heart failure. Biochem Biophys Res Commun, 307: 771-777. Unger T (2001). Inhibiting renin-angiotensin in the brain: the possible therapeutic implications. Blood Press Suppl, 1: 12-6.

Bibliography

344

Vakili BA, Okin PM and Devereux RB (2001). Prognostic implications of left ventricular hypertrophy. Am Heart J, 141: 334-341. Vallega GA, Canessa ML, Berk BC, Brock TA and Alexander RW (1988). Vascular smooth muscle Na+-H+ exchanger kinetics and its activation by angiotensin II. Am J Physiol, 254: C751-C758. van Bilsen M and Chien KR (1993). Growth and hypertrophy of the heart: towards an understanding of cardiac specific and inducible gene expression. Cardiovasc Res, 27: 1140-1149. Van Cruchten S and Van Den Broeck W (2002). Morphological and biochemical aspects of apoptosis, oncosis and necrosis. Anat Histol Embryol, 31: 214-223. van den Hoff MJ, van den Eijnde SM, Viragh S and Moorman AF (2000). Programmed cell death in the developing heart. Cardiovasc Res, 45: 603-620. Van Der Vusse GJ, Van Bilsen M and Glatz JFC (2000) Cardiac fatty acid uptake and transport in health and disease. Cardiovasc Res, 45: 279-293. van Eickels M, Grohe C, Cleutjens JP, Janssen BJ, Wellens HJ and Doevendans PA (2001). 17beta-estradiol attenuates the development of pressure-overload hypertrophy. Circulation, 104: 1419-1423. van Ginneken AC and Giles W (1991). Voltage clamp measurements of the hyperpolarization-activated inward current I(f) in single cells from rabbit sino-atrial node. J Physiol, 434: 57-83. van Kats JP, Danser AH, van Meegen JR, Sassen LM, Verdouw PD and Schalekamp MA (1998). Angiotensin production by the heart: a quantitative study in pigs with the use of radiolabeled angiotensin infusions. Circulation, 98: 73-81. van Kats JP, Methot D, Paradis P, Silversides DW and Reudelhuber TL (2001). Use of a biological peptide pump to study chronic peptide hormone action in transgenic mice. Direct and indirect effects of angiotensin II on the heart. J Biol Chem, 276: 44012-44017. van Rijen HV, van Kempen MJ, Postma S and Jongsma HJ (1998). Tumour necrosis factor alpha alters the expression of connexin43, connexin40, and connexin37 in human umbilical vein endothelial cells. Cytokine, 10: 258-264. Vanucci SJ, Rutherford T, Wilkie MB, Simpson IA and Lauder JM (2000). Prenatal expression of the GLUT4 glucose transporter in the mouse. Dev Neurosci, 22: 274-282. Vartiainen M, Ojala PJ, Auvinen P, Peranen J and Lappalainen P (2000). Mouse A6/twinfilin is an actin monomer-binding protein that localizes to the regions of rapid actin dynamics. Mol Cell Biol, 20: 1772-1783.

Bibliography

345

von Lutterotti N, Catanzaro DF, Sealey JE and Laragh JH (1994). Renin is not synthesized by cardiac and extrarenal vascular tissues: a review of experimental evidence. Circulation, 89: 458-470. Wachtell K, Lehto M, Gerdts E, Olsen MH, Hornestam B, Dahlof B, Ibsen H, Julius S, Kjeldsen SE, Lindholm LH, Nieminen MS and Devereux RB (2005). Angiotensin II receptor blockade reduces new-onset atrial fibrillation and subsequent stroke compared to atenolol: the Losartan Intervention For End Point Reduction in Hypertension (LIFE) study. J Am Coll Cardiol, 45: 712-719. Wang H, Singh D and Fliegel L (1997). The Na+/H+ antiporter potentiates growth and retinoic acid-induced differentiation of P19 embryonal carcinoma cells. J Biol Chem, 272: 26545-26549. Wang L, Eberhard M and Erne P (1995). Stimulation of DNA and RNA synthesis in cultured rabbit cardiac fibroblasts by angiotensin IV. Clin Sci (Lond), 88: 557-562. Wang Q, Chen Q and Towbin JA (1998). Genetics, molecular mechanisms and management of long QT syndrome. Annals of Medicine, 30: 58-65. Wang QD and Nygren E (2003). Various inotropic effects of angiotensin II in post-ischaemic rat hearts depending on ischaemic time with possible involvement of protein kinase C. Acta Physiol Scand, 178: 189-196. Wang Z, Nolan B, Kutschke W and Hill JA (2001). Na+-Ca2+ exchanger remodeling in pressure-overload cardiac hypertrophy. J Biol Chem, 276: 17706-17711. Watanabe A and Endoh M (1998). Relationship between the increase in Ca2+ transient and contractile force induced by angiotensin II in aequorin-loaded rabbit ventricular myocardium. Cardiovasc Res, 37:524-531. Weaver D, Skinner S, Walker L and Sangster M (1991). Phenotypic inhibition of the renin-angiotensin system, emergence of the Ren-2 gene, and adaptive radiation of mice. Gen Comp Endocr, 83: 306-315. Weber KT and Brilla CG (1991). Pathological hypertrophy and cardiac interstitium. Fibrosis and rennin-angiotensin-aldosterone system. Circulation, 83: 1849-1865. Wei JY, Spurgeon A and Lakatta EG (1984). Excitation-contraction in rat myocardium: alterations with adult aging. Am J Physiol, 246: H784-H791. Wei M, Gaskill SP, Haffner SM and Stern MP (1998). Effects of diabetes and level of glycemia on all-cause and cardiovascular mortality. The San Antonio Heart Study. Diabetes Care, 21: 1167-1172. Weichenrieder O, Stehlin C, Kapp U, Birse DE, Timmins PA, Strub K and Cusack S (2001). Hierarchical assembly of the Alu domain of the mammalian signal recognition particle. RNA, 7: 731-740.

Bibliography

346

Wencker D, Chandra M, Nguyen K, Miao W, Garantziotis S, Factor SM, Shirani J, Armstrong RC and Kitsis RN (2003). A mechanistic role for cardiac myocyte apoptosis in heart failure. J Clin Invest, 111: 1497-1504. Wenzel S, Taimor G, Piper HM and Schlüter KD (2001). Redox-sensitive intermediates mediate angiotensin II-induced p38 MAP kinase activation, AP-1 binding activity, and TGF-beta expression in adult ventricular cardiomyocytes. FASEB J, 15: 2291–2293. Wettschureck N, Rutten H, Zywietz A, Gehring D, Wilkie TM, Chen J, Chien KR and Offermanns S (2001). Absence of pressure overload induced myocardial hypertrophy after conditional inactivation of Galphaq/Galpha11 in cardiomyocytes. Nat Med, 7: 1236-1240. Wickenden AD, Kaprielian R, Kassiri Z, Tsoporis JN, Tsushima R, Fishman GI and Backx PH (1998). The role of action potential prolongation and altered intracellular calcium handling in the pathogenesis of heart failure. Cardiovasc Res, 37: 312-323. Wiendl H, Mitsdoerffer M, Schneider D, Melms A, Lochmuller H, Hohlfeld R and Weller M (2003). Muscle fibres and cultured muscle cells express the B7.1/2-related inducible co-stimulatory molecule, ICOSL: implications for the pathogenesis of inflammatory myopathies. Brain,126:1026-1035. Wikman-Coffelt J, Wu ST, Parmley WW and Mason DT (1991). Angiotensin II and phorbol esters depress cardiac performance and decrease diastolic and systolic [Ca2+]i in isolated perfused rat hearts. Am Heart J, 122: 786-794. Willecke K, Eiberger J, Degen J, Eckardt D, Romualdi A, Guldenagel M, Deutsch U, and Sohl G (2002). Structural and functional diversity of connexin genes in the mouse and human genome. Biol Chem, 383: 725–737. Wollert KC and Drexler H (1999). The renin-angiotensin system and experimental heart failure. Cardiovasc Res, 43:838-849. Xiao XH and Allen DG (2000). Activity of the Na(+)/H(+) exchanger is critical to reperfusion damage and preconditioning in the isolated rat heart. Cardiovasc Res, 48: 244-253. Yamazaki T, Komuro I, Kudoh S, Zou Y, Nagai R, Aikawa R, Uozumi H, Yazaki Y (1998). Role of ion channels and exchangers in mechanical stretch-induced cardiomyocyte hypertrophy. Circ Res, 82: 430-437. Yang G, Merrill DC, Thompson MW, Robillard JE and Sigmund CD (1994). Functional expression of the human angiotensinogen gene in transgenic mice. J Biol Chem, 269: 32497-32502. Yang J, Gillingham AK, Hodel A, Koumanov F, Woodward B and Holman GD (2002). Insulin-stimulated cytosol alkalinization facilitates optimal activation of glucose transport in cardiomyocytes. Am J Physiol, 283: E1299-E1307.

Bibliography

347

Yasukawa H, Hoshijima M, Gu Y, Nakamura T, Pradervand S, Hanada T, Hanakawa Y, Yoshimura A, Ross J Jr, Chien KR (2001). Suppressor of cytokine signaling-3 is a biomechanical stress-inducible gene that suppresses gp130-mediated cardiac myocyte hypertrophy and survival pathways. J Clin Invest, 108: 1459-1467. Yatani A, Frank K, Sako H, Kranias EG and Dorn GW 2nd (1999). Cardiac-specific overexpression of Galphaq alters excitation-contraction coupling in isolated cardiac myocytes. J Mol Cell Cardiol, 31: 1327-1336. Ye MQ and Healy DP (1992). Characterization of an angiotensin type-1 receptor partial cDNA from rat kidney: evidence for a novel AT1B receptor subtype. Biochem Biophys Res Commun, 185: 204-210. Yoshida H and Karmazyn M (2000). Na+–H+ exchange inhibition attenuates hypertrophy and heart failure in 1-week postinfarction rat myocardium. Am J Physiol, 278: H300–H304. Young ME, Laws FA, Goodwin GW and Taegtmeyer H (2001). Reactivation of peroxisome proliferator-activated receptor alpha is associated with contractile dysfunction in hypertrophied rat heart. J Biol Chem, 276: 44390-44395. Yu Z and Mcneill JH (1991). Force-interval relationship and its response to ryanodine in streptozotocin-induced diabetic rats. Can J Physiol Pharmacol, 69: 1268-1276. Yu Z, Tibbits GF and McNeill JH (1994). Cellular functions of diabetic cardiomyocytes: contractility, rapid-cooling contracture, and ryanodine binding. Am J Physiol, 266: H2082-H2089. Yun J, Zuscik MJ, Gonzalez-Cabrera P, McCune DF, Ross SA, Gaivin R, Piascik MT and Perez DM (2003). Gene expression profiling of alpha(1b)-adrenergic receptor-induced cardiac hypertrophy by oligonucleotide arrays. Cardiovasc Res, 57: 443-455. Yussman MG, Toyokawa T, Odley A, Lynch RA, Wu G, Colbert MC, Aronow BJ, Lorenz JN and Dorn GW 2nd (2002). Mitochondrial death protein Nix is induced in cardiac hypertrophy and triggers apoptotic cardiomyopathy. Nat Med, 8: 725-730. Zapala MA, Lockhart DJ, Pankratz DG, Garcia AJ, Barlow C and Lockhart DJ (2002). Software and methods for oligonucleotide and cDNA array data analysis. Genome Biol, 3: SOFTWARE0001. Zarich SW and Nesto RW (1989). Diabetic cardiomyopathy. Am Heart J, 118: 1000-1012. Zeng FY, Gerke V and Gabius HJ (1993). Identification of annexin II, annexin VI and glyceraldehyde-3-phosphate dehydrogenase as calcyclin-binding proteins in bovine heart. Int J Biochem, 25: 1019-1027.

Bibliography

348

Zhang JX, Braakman I, Matlack KE and Helenius A (1997). Quality control in the secretory pathway: the role of calreticulin, calnexin and BiP in the retention of glycoproteins with C-terminal truncations. Mol Biol Cell, 8: 1943-1954. Zhang W, Kowal RC, Rusnak F, Sikkink RA, Olson EN and Victor RG (1999). Failure of calcineurin inhibitors to prevent pressure-overload left ventricular hypertrophy in rats. Circ Res, 84: 722-728. Zhong Y, Ahmed S, Grupp IL and Matlib MA (2001). Altered SR protein expression associated with contractile dysfunction in diabetic rat hearts. Am J Physiol, 281: H1137-H1147. Zhu YZ, Zhu YC, Stoll M and Unger T (2000). Identification of regulated genes in rat heart after myocardial infarction by means of differential mRNA display. Jpn Heart J, 41: 59-66. Zhuo J, Alcorn D, Allen AM and Mendelsohn FA (1992). High resolution localization of angiotensin II receptors in rat renal medulla. Kidney Int, 42: 1372-1380. Zisman A, Peroni OD, Abel ED, Michael MD, Mauvais-Jarvis F, Lowell BB, Wojtaszewski JF, Hirshman MF, Virkamaki A, Goodyear LJ, Kahn CR and Kahn BB (2000). Targeted disruption of the glucose transporter 4 selectively in muscle causes insulin resistance and glucose intolerance. Nat Med, 6: 924-928. Zorn-Pauly K, Schaffer P, Pelzmann B, Lang P, Machler H, Rigler B and Koidl B (2004). If in left human atrium: a potential contributor to atrial ectopy. Cardiovasc Res, 64: 250-259.