<<

REGULATION OF ATP-SENSITIVE POTASSIUM CHANNELS IN THE HEART

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree

Doctor of Pharmacy

in the Graduate School of The Ohio State University

By

Vivek Garg, M.Pharm

*********

The Ohio State University

2009

Dissertation Committee: Approved by

Professor Keli Hu, Advisor

Professor Terry S. Elton Professor Lane J. Wallace ------Advisor Professor Dale G. Hoyt Graduate Program in Parmacy

Copyright by

Vivek Garg

2009

ABSTRACT

ATP-sensitive potassium channels (KATP channels) link the cellular

energy levels to potential and excitability in various cell types.

They control many important functions like insulin secretion in pancreatic β- cells; vascular tone in vascular smooth muscle cells; and duration in cardiac myocytes and neurons under ischemic conditions. In cardiac myocytes, it has been shown that KATP channels on plasma

membrane are composed of Kir6.2 and SUR2A subunits in 4:4 stoichiometry.

Though many regulators of KATP channels and signal transduction

mechanisms regulating opening or closing of KATP channels have been identified, much less is known about the subcellular localization of KATP channels which can have a profound impact on the temporal and spatial regulation by its signaling modulators. Many studies have localized KATP channels to nucleus and mitochondria, besides cellular plasma membrane.

However, it not known if these are the same cell surface KATP channels, which

are also targeted to mitochondria, or some other isoform of KATP channel. The

plasma membrane itself is not homogenous through out. It is interspersed by

sub-domains rich in sterols and glycosphingolipids. In majority of cases, this sub-domain organization is orchestrated by special proteins called as

ii caveolins, which are the main structural components of caveolae. Caveolae

are small (50 to 100 nm), and sphingolipid enriched “cave”-like

invaginations of the surface membrane. These specialized lipid microdomains have the ability to selectively compartmentalize many signaling molecules, including many modulators of KATP channels. Since little is known about the subcellular locations of KATP channel protein, we designed our studies to

characterize the localization of KATP channel protein in cardiac myocytes.

We first endeavored to look at KATP channel localization along the plasma membrane in rat cardiac myocytes. Using a variety of different

techniques on isolated murine cardiomyocytes, we found that majority of

cardiac KATP channels are localized to caveolin-enriched membrane

microdomains. Further, whole-cell voltage-clamp recording in both adult and

neonatal cardiac myocytes confirmed our hypothesis that caveolae integrity is essential for activation of KATP channel by its modulator (Chapter

1). Adenosine released from ischemic myocardium is a very important

modulator of KATP channels. These findings have significant implications for cardioprotective role of KATP channels during ischemic conditions.

A signaling function for caveolins either on their own (direct) or by

acting as scaffolding proteins (indirect) has also been described. To test its

relevance for KATP channels, we employed HEK293T cells transfected with

recombinant cardiac KATP channels (Kir6.2/SUR2A) with or without caveolin-3

(Cav-3, a muscle-specific caveolin isoform). We found that Cav-3 has

significant inhibitory effect on KATP channel current density which can be

iii reversed by a scaffolding domain peptide from caveolin-3 protein sequence

(CSD) (Chapter 2). This crucial experiment indicates a very interesting, dual

regulation of KATP channels by caveolins and caveolae. Though caveolae

structure ensures that KATP channel modulators are close to the channel proteins for efficient signal transduction, caveolin-3 protein through direct or

indirect interaction with channel proteins makes sure that they remain

inhibited until required.

Whether Kir6.2 containing KATP channel are present in mitochondria or

not is controversial; nevertheless no one has ever studied the trafficking

aspect of KATP channel to mitochondria. In our effort to elucidate the sub-

cellular localization of cardiac KATP channels, we hypothesized that

localization of Kir6.2-containing KATP in mitochondria can be increased by

activation of (PKC). Utilizing KATP-deficient COS-7 cells, we

reported a novel finding that a specific protein kinase C (PKC) isoform, PKCε,

promotes mitochondrial import of Kir6.2-containing KATP channels from cytosol. These findings were further corroborated with functional data using mitochondrial potential measurement studies.

Collectively, our data demonstrated that besides bulk plasma

membrane, Kir6.2 containing cardiac KATP channels are localized to two

distinct sub-cellular locations, namely caveolae and mitochondria. Their

localization can be modulated by specific regulatory pathways, and

furthermore

iv furthermore their regulation can be affected by their sub-cellular localization.

This provides valuable insight into the mechanisms regulating KATP channels,

which have been implicated for cardioprotection under ischemic conditions.

v

Dedicated to My parents & family

vi

Acknowledgements

I would like to pay a tribute to all my teachers (past and present) through this thesis. There are so many people to thank for who have contributed directly or indirectly to this work by, influencing my thinking, discussions, advice and of course blessings and wishes. My deepest gratitude is due to my advisor Dr. Keli Hu for her continued support, guidance and encouragement though out my graduate work; for allowing me to work independently on research areas of my interest. Her attention to detail, focus and fruitful suggestions made my work worthy of presentation. I wish to thank Dr. Jundong Jiao, who started me on patch clamp studies. I also thank Arun Sridhar, Veronique Lacombe for helpful discussions on perforated-patch and voltage-clamp studies. My thanks are due to Dr. Douglas Pfeiffer for discussions regarding mitochondrial studies and allowing me to work in his lab for some of the experiments. I would like to thank my committee members Dr. Lane Wallace, Dr. Terry Elton, Dr. Dale G. Hoyt for their time, effort and sharing their valuable opinions regarding this work. I also thank the faculty members of the pharmacology division especially Dr. Lakhu Keshvara, Dr. Popat N. Patil, Dr. Kari Hoyt, Dr. Cynthia Carnes for their encouragement, conversations, and time to time intellectual stimulation. My training as a research scientist is greatly enriched with your interactions. I thank all of my friends and wonderful colleagues in the division Vaibhav, Zhaogang, Tongzheng, Brandon, Ryan, Rachel, Raeann, Sarmistha who helped me in different ways, be it reagents or discussions regarding individual experiments. I will never forget my friends at OSU. You guys made

vii my stay here truly enjoyable. Thanks for sharing times of happiness and celebration; and emotional support in times of stress. I am just humbled by all kinds of contributions and sacrifices made by my parents Dr. Pawan Kumar Garg, Mrs. Kamla Garg and family members Priya, Rishubh, Upma, Anupma, Sanjeev, Tarun, Rohan, Honey, and Aryan. I would have never reached at the stage where I am now without them. Though far, you are always in my thoughts as a perpetual source of inspiration. I wish to thank you all for your unconditional love, blessings and happiness. Last but not the least, I am highly grateful to my wife Priyanka for being supportive and infinitely patient in tolerating my business. Her support and encouragement was in the end what made this dissertation possible in time. She was always there cheering me up in all my endeavors.

viii

VITA

July 18, 1977……………………...Born – Punjab, India

1995 – 1999.………………………Bachelor of Pharmacy

Punjab University, India

2000 – 2002………………………..Master of Pharmcy

Punjab University, India

2004 – Present……………………Graduate Teaching Associate

The Ohio State University, Columbus, Ohio

PUBLICATIONS

1. Garg V, Jiao JD, Hu K. (2009) ATP-sensitive K+ channels are

regulated by caveolin-enriched microdomains in cardiac myocytes.

Cardiovasc Res 82: 51-58.

2. Jiao J, Garg V, Yang B, Elton TS, Hu K. (2008) PKC-epsilon induces

caveolin-dependent internalization of vascular ATP-sensitive K+

Channels. Hypertension 52: 499-506.

ix 3. Jiao JD*, Garg V*, Yang B, Hu K (2008) Novel functional role of heat

shock protein 90 in ATP-sensitive K+ channel-mediated hypoxic

preconditioning. Cardiovasc Res 77: 126-33 (*equal contribution).

4. Garg V, Hu K (2007) Protein kinase C isoform-dependent modulation

of ATP-sensitive K+ channels in mitochondrial inner membrane. Am J

Physiol Heart Circ Physiol 293: H322-332.

5. Singh A, Garg V, Gupta S, Kulkarni SK (2002) Role of antioxidants in

chronic fatigue syndrome in mice. Indian J Exp Biol 40: 1240-1244.

FIELDS OF STUDY

Major Field: Pharmacy

x

TABLE OF CONTENTS

PAGE

Abstract……………………………………………………………….…………..ii

Dedication…………………………………………………………….………….vi

Acknowledgements………………………………………………….………….vii

Vitae…………………………………………………………………….………..ix

List of Figures………………………………………………………….………..xiv

Abbreviations………………………………………………………….………...xvi

CHAPTER 1. INTRODUCTION………………………………..………..1

1. ATP-sensitive potassium channels……………………………………...1

1.1. Molecular basis……………………………………………...………...... 3

1.2. Biophysical properties………………………………………...…………6

1.3. Regulation……………………………………………………………...... 7

1.3.1. regulation………………………………...... ….……7

1.3.2. Pharmacological regulation………………………...……………9

1.3.3. Regulation by signaling molecules……………………………..11

1.3.3.1. Phosphoinsotides (PIP2)…………………………:...... 13

1.3.3.2. Protein kinase C ……………………..…………...……14

1.3.3.3. Subsarcolemmal ATP………..………………....…...... 17

xi 1.3.3.4. Nitric oxide ……………………………………………….21

1.3.3.5. G-protein coupled receptors (GPCRs)……………...... 22

1.4. Functional role………………………………………………...... …..…....30

2. Caveolae……………………………………….…………………………...... 38

2.1. Caveolae and caveolins……………………………….………...………38

2.2. Functional role of caveolae/caveolins……………….……………..…..41

2.3. Compartmentalization of ion channels…………..………..…………...44

2.3.1. Cardiac ion channels……………………………………….…….45

2.3.2. KATP channels……………………………………………………..48

3. Objectives……………………………………………………………...……..49

CHAPTER 2. REGULATION OF ATP-SENSITIVE POTASSIUM

CHANNELS BY CAVEOLIN-ENRICHED

MICRODOMAINS IN CARDIAC MYOCYTES………..51

1. Introduction……………………………………………...…….…...... 53

2. Materials and Methods……………………………………………....54

3. Results……………………………………………...…….…….…...... 59

4. Discussion…………………………………………….....…..………..81

5. Acknowledgements…………………………………….…….……....85

CHAPTER 3. CAVEOLIN-3 NEGATIVELY REGULATES

RECOMBINANT CARDIAC ATP-SENSITIVE

POTASSIUM CHANNELS IN HEK293T CELLS….....86

xii 1. Introduction………………………………….……………...….…...... 88

2. Materials and Methods…………………….………………………...89

3. Results……………………………………..……….…….…………...93

4. Discussion……………………………………………..……….……..105

5. Acknowledgements………………………...…………….….……....108

CHAPTER 4. PROTEIN KINASE C ISOFORM-DEPENDENT

MODULATION OF ATP-SENSITIVE POTASSIUM

CHANNELS IN MITOCHONDRIAL INNER

MEMBRANE………………………………...…………...109

1. Introduction…………………………………………...... ….……...111

2. Materials and Methods……………………………...………….…...113

3. Results……………………………………………...... ….…………...119

4. Discussion……………………………………………....….………...141

5. Acknowledgements…………………………………...…….…..…...147

CHAPTER 5. CONCLUSIONS AND FUTURE DIRECTIONS……....148

BIBLIOGRAPHY…………………………………………….…………………...154

xiii

LIST OF FIGURES

Figure Page

1.1. Topology and assembly of KATP channels……………………..……..3

1.2. Comparison of biophysical and pharmacological properties

of different isoforms of KATP channels………………………….……..4

1.3. Topology and predicted membrane insertion of Cav-1………….….39

2.1. Kir6.2 is enriched in caveolar plasma membrane in

adult rat cardiomyocytes…………………………………………...…..62

2.2. KATP channels are associated with caveolin-3 (Cav-3) in

adult rat ventricular myocytes……………………………………..…..65

2.3. Colocalization of Kir6.2 and caveolin-3 (Cav-3)……………….…….67

2.4. Caveolar disruption with MβCD eliminates adenosine

-mediated stimulation of KATP channels…………….………71

2.5. Knockdown of caveolin-3 expression with siRNA prevents

-mediated activation of KATP channels…………75

2.6. Immunoblot analysis of samples from neonatal rat hearts

immunoprecipitated with anti-caveolin-3 (Cav-3) or

anti-Kir6.2 antibodies……………………………………………...……78

xiv 2.7. Effect of MβCD on Kir6.2, adenosine A1 receptors (Ade A1)

and caveolin-3 (Cav-3) coprecipitation………………………………..80

3.1. Modulation of spontaneous KATP currents by caveolin-3………….…95

3.2. Modulation of -activated KATP currents by caveolin-3………98

3.3. Effect of caveolin-3 scaffolding domain peptide (SDP) on

KATP currents……………………………………………………………..102

3.4. Co-immunoprecipitation and colocalization of Kir6.2/SUR2A

channels with caveolin-3………………………………………………..104

4.1. Immunofluorescence microscopy of mitochondrial localization

of Kir6.2 in mitoplasts isolated from COS-7 cells…………………….123

4.2. Proximity of Kir6.2-cyan fluorescent protein (CFP) to

Mito-yellow fluorescent protein (YFP) in mitochondrial matrix

yielded fluorescence resonance energy transfer (FRET) signals…..127

4.3. Immunoblot analysis of Kir6.2 protein in mitochondria……………....131

4.4. PMA-induced KATP-dependent changes in mitochondrial

membrane potential (∆ψm)…………………………………………..….136

4.5. Immunofluorescence microscopy of mitochondrial localization of

Kir6.2 in mitochondria isolated from rat adult cardiomyocytes….…..140

5.1 Hypothetical model for the targeting of KATP channels to plasma

membrane, caveolar microdomains and mitochondria……….……...153

xv

ABBREVIATIONS

5-HD 5-Hydroxy decanoate

ABC ATP-binding cassette

AC Adenylate cyclase

Ach Acetylcholine

APD Action potential duration

ATP

[ATP]I Intracellular ATP

BKca -activated

Cav-1 Caveolin-1

Cav-2 Caveolin-2

Cav-3 Caveolin-3

CBD Caveolin binding domain

CFTR Cystic fibrosis transmembrane conductance

regulator channel

CK Creatine kinase

Cr Creatine

CRP Caveolae resident protein

CSD Caveolin scaffolding domain

xvi DAG Diacylglycerol

DEAE Diethylaminoethyl cellulose

E-C Excitation-contraction eNOS Endothelial nitric oxide synthase

Erev Reversal potential

FRET Fluorescence resonance energy transfer

GAPDH Glyceraldehyde phosphate dehydrogenase

GIRKs G-protein activated inwardly rectifying K+ channels

GPCR G-protein coupled receptor

HCN4 Hyperpolarization-activated cyclic nucleotide gated

HEK Human embryonic kidney

ICa,L L-type current

If ‘funny’ current (mediated by HCN channels)

IKACh Muscarinic-gated potassium channel

+ IKr Rapidly activating delayed rectifier K channel

current

IMM Inner membrane of mitochondria

INa Voltage-activated current

IP3 triphosphate

IPC Ischemic preconditioning

Ito Transient outward potassium current

ITP triphosphate

KATP ATP-sensitive potassium channel

xvii KCOs Potassium channel openers

Kir Inwardly rectifier potassium channel

KO Knock out

Kv Voltage-activated potassium channel

M-LDH Muscle isoform of Lactate dehydrogenase

MβCD Methyl-β-cyclodextrin

Na/K-ATPase Sodium-potassium ATPase

NBD Nucleotide binding domain

NBF1 Nucleotide-binding fold-1

NBF2 Nucleotide-binding fold-2

NCX Sodium-calcium exchanger

NDP diphosphate

PC preconditioning

PFK Phosphofructokinase

PI Phosphoinositide

PIP2 Phosphatidylinositol-4,5-bisphosphate

PKA Protein kinase A

PKC Protein kinase C

PLC Phospholipase C

PMA Phorbol-12-myristate-13-acetate

PTX Pertussis toxin

RACK Receptor for activated C kinase

ROS Reactive oxygen species

xviii SarcKATP Sarcolemmal KATP channel siRNA Small interfering RNA

SUR Sulphonylurea receptor

TIM Translocases of the inner membrane of

mitochondria

TOM Translocases of the outer membrane of

mitochondria

TRP Transient receptor potential channel

UDP diphosphate

UTP Uridine-5'-triphosphate

xix

CHAPTER 1 INTRODUCTION

1. ATP-sensitive potassium channels

Ion channels are specialized membrane proteins which provide

hydrophilic conduction pathways for various ions across the hydrophobic

interior of lipid bilayers. They exhibit ion-selectivity based on size, charge and water of hydration. Based on ion-selectivity, they are classified broadly as sodium channels, potassium channels, calcium channels and chloride channels. Typically, they are composed of multi-subunits which are arranged in a circular fashion around a water-filled tunnel/pore which has openings on both extracellular and intracellular side. The pore-forming subunits are called as α-

subunits while the accessory subunits are called as β-, γ-subunits and so on. In

many ion channels, the opening or closing of the pore is governed by a ‘gate’

which is controlled by various signals like chemical, electrical or mechanical

signals. In the heart most of the ion channels are voltage gated while some are

(chemical) gated. For the purposes of a review relevant to the content of

this work, I will mainly focus on a ligand gated potassium channel called as

ATP-sensitive potassium channels.

1 Originally discovered in the heart, ATP-sensitive potassium channels

(KATP channels) are widely distributed in various other tissues and cell types

like pancreas, smooth muscles and neurons (197). KATP channels activation/inhibition and cellular energy metabolism are coupled in most of the

excitable tissues where they are expressed. Inhibition of KATP channels by ATP

and stimulation by nucleotide diphosphates regulate insulin release in

pancreatic β-cells; vascular tone in smooth muscle cells; and action potential

duration (APD) shortening in cardiomyocytes and neurons (192). Pancreatic β-

cells offer the classical and simplistic paradigm to understand KATP channel

function. These channels are open under the basal conditions in the pancreas,

leading to significant K+ outflux and hyperpolarized state of the membrane.

However, when plasma glucose levels increase; an increase in pancreatic β-

cell [ATP/ADP]i ratio causes inhibition of these channels leading to opening of

calcium channels, consequent membrane depolarization, and insulin release.

This is considered to be the fundamental mechanism responsible for insulin

release in response to glucose. Interestingly, in the heart, these channels are

closed under the basal conditions, leading many people to question their

physiological relevance. Nevertheless, these channels shorten APD in

response to metabolic or ischemic stress and mediate a robust cardioprotective

response. As evidenced by recent reports and will be discussed later, these

channels do play an important role in normal physiological functions of the

heart as well (section 1.4).

2 1.1. Molecular basis

KATP channels are essential hetero-octamers, consisting of two subunits:

four pore forming subunits (Kir6.x), and four regulatory subunits (SURx) (2)

(Fig. 1.1). While Kir6 is a member of inwardly rectifier potassium channel

family; the other subunit SUR (for sulphonylurea receptor), is a member of

ATP-binding cassette (ABC) protein superfamily, which includes CFTR (Cystic

fibrosis transmembrane conductance regulator channel, a )

and P-glycoprotein etc. Two isoforms of Kir6 have been identified: Kir6.1,

Kir6.2; and three major isoforms have been described for SUR namely, SUR1,

SUR2A and SUR2B. Further, many small isoforms and splice variants have

also been described in the various tissues in mice (260). The exact

physiological role of these isoforms is unclear. Different subunit composition in

different tissues gives rise to KATP channels with different biophysical and

pharmacological properties: pancreatic β-cells (SUR1/Kir6.2), cardiac myocytes

(SUR2A/Kir6.2), brain (SUR1/Kir6.2), and vascular smooth muscle cells

(SUR2B/Kir6.1) (summarized in Figure 1.2).

It is now widely accepted that sarcolemmal KATP channels in cardiomyocytes are composed of Kir6.2 and SUR2A subunits (77). Both

-/- -/- Kir6.2 and SUR2 mice have no KATP channel current in cardiomyocytes (44,

159, 269). Nevertheless SUR1, SUR2B and Kir6.1 immunoreactivity and

transcripts have also been detected in the heart in several studies (104, 187,

221, 265).

3

Figure 1.1. Topology and assembly of KATP channels. The topologies of

SUR (Upper panel, left), and Kir6.2 (Upper panel, right) are illustrated. (Lower panel) The channel is assembled from SUR1/2 and KIR6.2 subunits as an octameric complex in 4:4 stoichiometry, with four Kir6.2 subunits forming the inner pore and each Kir6.2 subunit attached to one SUR subunit from outside.

Abbreviations: N = N-terminus, C = C-terminus, TMD = transmembrane domain, M = membrane domain, P = pore region, A and B = walker domains A and B, NBF = nucleotide binding fold. Adapted from Aguilar-Bryan and Bryan,

1999 (2).

4

Type Subunit Unitary Ki for IC50 by KCO’s Composition Conductance ATP sensitivity (pS) inhibition (nM) (μM)

Cardiac & SUR2A/ 80 100-170 6-9 P=C Skeletal Kir6.2 muscle Pancreatic SUR1/ Kir6.2 76 10 5 D>P>C Vascular SUR2B/ 33 >1000 20-100 P=C>D Smooth Kir6.1 Muscle

Figure 1.2. Comparison of biophysical and pharmacological properties of

different isoforms of KATP channels. Abbreviations: KCO = Potassium

, P = Pinacidil, C = , D = .

A recent study by Flagg et el. 2008 (77) has shown that SUR1 is an essential component of sarcolemmal KATP channels in atrial myocytes but not in

ventricular myocytes in mice. This might be responsible for different

pharmacological profile (different sensitivities to KCOs) and different functional

responses of atrial and ventricular myocytes during ischemic conditions. Using

transgenic approaches, it was further deduced that SUR1 containing channels

open much more readily as compared to SUR2A containing KATP channels in

the heart. Further, mice overexpressing SUR1 in the heart were more prone to

arrhythmic events as compared to SUR2A overexpressing hearts. Why SUR2A

containing KATP channels are silent in the heart as compared to SUR1 is not

clear.

5 Based on different sensitivities to ATP depletion and differential APD

shortening in epicardial versus endocardial myocytes under ischemic

conditions, there are reports proposing that there may be transmural

heterogeneity of KATP channel within the heart (81, 182). Adding to the diversity

and complexity of KATP channels are the studies showing that there are KATP channels in the subcellular organelles also, like mitochondria (119, 227, 265) and nucleus (227). Though the molecular identity of surface KATP channels is

much clearer, the subunit structure of KATP channels at intracellular locations is

still enigmatic.

1.2. Biophysical properties

Like other potassium channels, KATP channels are selectively permeable

to potassium and negligibly permeable to sodium. They exhibit time- and

voltage-independent kinetics and are ligand gated (ATP). The single-channel

conductance in symmetrical K+ solution is ~70-90 pS which is much higher than

any other potassium channel except calcium-activated potassium channels

(BKCa, ~210 pS). The kinetics of KATP channel openings are characterized by a

unique ‘flickering behavior’ consisting of ‘burst’ of channel openings followed by

relatively long closed periods or silent ‘gaps’. The main effect of ATP is by

reducing the burst duration (affects both mean open time and number of

openings per burst) and prolonging the gap period.

These channels are weak inwardly rectifiers at very positive potentials

as conduction of K+ ions is better in inward direction as compared to outward 6 direction at those potential. Rectification describes the relationship between the

voltage and the conductance of an ion. In non-rectifying current, conductance

is voltage independent and follows Ohm’s law as current is directly dependent

on voltage (V=IR). A channel is an inward rectifier when it passes current

(positive charged ion) more readily in inward direction (into the cell) as

compared to outward direction at potentials positive to reversal potential (Erev).

The main cause of rectification is the high-affinity voltage-dependent block by

intraceullar cations, principally Mg2+ and Na+, and polyamines, namely

spermine, spermidine and putrescine that block the channel pore at positive potentials (2).

1.3. Regulation

1.3.1. Nucleotide regulation

The most characteristic property of KATP channels is its inhibition by

intracellular free ATP which does not require Mg2+. The intracellular free ATP

appears to bind to the pore-forming subunit Kir6.2, while most pharmacological

agents (K+ channel openers and sulphonylureas) and MgADP target SUR2A

subunit. The ATP inhibition on Kir6.2 can be modulated by the respective

binding of ATP and ADP to NBF1 (Nucleotide binding fold 1) and NBF2 of

SUR2A subunit. Since Kir6.2 and SUR2A coassemble in 4:4 stoichiometry, as

per the widely accepted model, there are four ATP binding sites on KATP

channel and two MgADP binding site on each SUR2A subunit. Though only

one ATP site needs to be occupied to inhibit the channel, ATP binding at other 7 sites stabilizes the channel inhibition indicating a cooperative interaction. Thus,

ATP sensitivity of KATP channel is not constant throughout (183). Further,

considerable patch to patch variation besides transmural heterogeneity in ATP

sensitivity of KATP channels has also been noted (182). The ATP sensitivity also

decreases during metabolic stress and ischemic conditions (284). The

adenosine ring is important for high affinity inhibition, as other nucleoside

triphosphates like GTP ( triphosphate), ITP ()

and UTP () are much less effective as compared to ATP.

Most of the nucleoside diphosphates (NDP) like ADP and UDP (uridine

diphsophate) antagonize ATP dependent inhibition in Mg2+-dependent manner.

MgADP interaction at SUR2A allosterically affects the interaction of ATP at

Kir6.2 subunit. The effect of MgADP is paradoxical; it activates the channel at

lower concentrations while concentrations higher than 500 μM have inhibitory

effect (183).

An interesting aspect of many ion channels, including KATP channels is

‘run-down’, which is time-dependent decrease in open probability (events of

channel opening) which is often observed in cell-free excised patches (a patch

membrane excised from the cell by recording pipette). This indicates the

requirement of a cytoplasmic factor to maintain channel activity. Divalent

cations like Ca2+ and Mg2+ promote run-down when applied on the intracellular side, while ATP, ADP and UDP decrease run-down in Mg2+-dependent manner.

Even though there is no direct evidence indicating that ATP hydrolysis at NBFs

of SUR is required for channel opening, non-hydrolysable analogs of ATP like 8 ATP[γS] or βγ-methylene ATP are ineffective in preventing run-down, suggesting the requirement for possible ATP hydrolysis at NBFs of SUR subunit (203). Channel phosphorylation by protein kinases was ruled out using specific inhibitors. However, wortmannin was found to inhibit MgATP dependent recovery of KATP channels, indicating that Phosphoinositide-kinases are likely involved. Since PIP2 (Phosphatidylinositol-4,5-bisphosphate) can

activate and recover run-down KATP channel, it was proposed that lipid

phosphorylation may be involved in MgATP-dependent recovery of run-down

channels (306).

1.3.2. Pharmacological regulation

KATP channels are the primary target for classical antidiabetic drugs,

sulphonylureas (glibenclamide, etc) and selective potassium

channel openers (KCOs) (pinacidil and cromakalim etc). The binding of

sulphonylurea follows two-site model with a very high-affinity binding site on

SUR (nanomolar range) and a low-affinity site on Kir6.x (millimolar range,

clinically irrelevant site). The binding affinity and potency of inhibition by

sulphonylureas varies depending on the specific isoform of the KATP channel

(pancreatic > cardiac > vascular) and type of sulphonylurea drug (tolbutamide,

, and bind and block pancreatic isoform, Kir6.2/SUR1 with

much higher affinity). Sulphonylureas bind to the cytoplasmic domains of SUR

subunit and close KATP channel, however, the exact transduction mechanism is

not known. Further, this interaction is significantly modulated by intracellular 9 as increased ATP increases sulphonylurea sensitivity. There is a

competitive antagonistic interaction between MgADP (at least at lower

concentrations) and sulphonylurea, which might be one of the reasons for

decreased sulphonylurea sensitivity during metabolic stress in cardiac

myocytes (224).

Potassium channel openers (KCOs) are a group of structurally diverse

compounds that open KATP channels (176). These drugs bind to and act on

SUR subunit and require ATP hydrolysis (Mg2+ dependent) at NBFs as non-

hydrolysable analogs of ATP are ineffective. The C-terminus of SUR was also found to significantly affect the binding affinity with relative affinity among different isoforms following the order: SUR2B ≈ SUR1 > SUR2A. Mg- nucleotides are essential for the action of KCOs (255). Pinacidil and cromakalim are more effective on vascular and cardiac KATP channels (vascular

> cardiac) as compared to pancreatic-isoform; while diazoxide is ineffective on cardiac sarcolemmal KATP channels. The effect of KCOs is increased by

intracellular decrease in pH, indicating that protons also modulate ATP-

sensitivity. Furthermore, lactate accumulation during ischemic conditions has been shown to activate KATP channels and shorten action potential duration;

however the effect was later found to be mainly due to acidosis caused by

lactate instead of lactate itself (24).

10 1.3.3. Regulation by signaling molecules

Under normal circumstances, KATP channels do not play any discernible

role in the heart as they are inhibited by high ATP concentrations inside the

cells. However, under the conditions of metabolic stress or ischemia, KATP

channel mediated outward potassium current increases rapidly, leading to APD

shortening (with decrease in calcium influx) and cardiac protection (93, 194).

This APD shortening and decrease in calcium influx helps in preserving ATP during energy deficient conditions. Moreover, using Kir6.2 knockout mice, it has been demonstrated that KATP channels play an important role of homeostatic

control in adaptation to stress (168, 320).

The normal physiological levels of ATP in a cell fluctuate between 6mM to 10mM (7, 193). Even under severe metabolic stress (when both oxidative phosphorylation and glycolysis are inhibited), intracellular ATP levels ([ATP]i)

fell down only by ~80% of control over the time (~5 mins) when significant APD

shortening occurs. At the same time, developed pressure is significantly

reduced and KATP channels are activated (7, 69, 194). On the other hand,

studies have consistently shown that for KATP channels, Ki(ATP) ([ATP]i causing

half maximal inhibition of channel activity) in cardiac myocytes is 100 µM in

excised inside-out patches (197) and ~500 µM in open-cell attached patch and whole cell configurations (76, 136). This wide discrepancy between the

observations obtained in intact cells and in any of the patch configurations

studied, points out that the regulation of KATP channels is probably much more

complex than can be explained by the simple hypothesis of opening by bulk 11 decrease in [ATP]i (192). An open question in the field is what exactly makes

these KATP channels open up at millimolar concentration of ATP inside an intact cell (or in-vivo conditions) when even micromolar amounts are enough to inhibit the channel?

Recent studies have shown that besides nucleotides KATP channels are

also regulated by a large number of other modulators. These modulators

include adenosine (114, 274), bradykinin (196), acetylcholine (274), and

adrenergic agents (116, 251) which act through G-protein coupled receptors;

protein kinase C (112, 163); protein kinase A (164); nitric oxide (NO) (36); and

membrane phospholipids, especially PIP2 (21, 263). These modulators apparently regulate ATP-dependent gating of the KATP channels by decreasing

ATP-sensitivity and thus increasing open probability. Since many of these

mediators/ receptors are released or activated during metabolic stress

conditions, they might affect the channel microenvironment and thus alter the

ATP sensitivity (Grover and Garlid, 2000; Zhuo et al., 2005). However, there

might be “fuzzy space” like barriers to diffusion (below the sarcolemma) which

would slow the access of modulators to KATP channels (156). Therefore, the

spatial and temporal localization of these modulators should be very close to

KATP channels to realistically explain the rapid activation of KATP channels

during metabolic stress. In other words, it may be possible that this signaling is compartmentalized. There are no reports of KATP channel compartmentalization

in the literature, even though the KATP channel microenvironment is considered to be different from “bulk” cytosol. One of the reasons for this different local 12 environment is that these channels exist in mutlimolecular complexes with

many metabolic : glyceraldehyde-3-phosphate dehydrogenase, triose-

phosphate isomerase, pyruvate kinase (64) and adenylate kinase (39). In the

following sections, I will discuss some of the important regulators of KATP

channels and evidence implicating their compartmentalization either alone or in

combination with KATP channels on membrane microdomains.

1.3.3.1. Phosphoinositides (PIP2)

Phosphoinositides (in particular phosphatidylinositol-4,5-bisphosphate,

PIP2) regulation of ion channels was discovered about a decade ago for the

+ 2+ first time in KATP channels and Na /Ca exchanger in cardiac myocytes (110).

Similar results were later found in skeletal muscle, pancreatic β-cells and

recombinant KATP channels (72). Later on, many other targets of

2+ PIP2 were discovered like other Kir channels, Kv channels, L-type Ca

channels, TRP channels and CFTR etc (111). PIP2 is produced by the

phosphorylation of phosphatidylinositol (PI) and phosphatidylinositol-4-

monophosphate (PI-4P) on the plasma membrane by the sequential action of

PIP-4-kinase (PIP4K) and PIP-5-kinase (PIP5K). This PIP2 is further

hydrolyzed by receptor activated Phospholipase C (PLC) or dephosphorylated by inositolpolyphosphate phosphatase.

PIP2 activates KATP channel current by increasing the open probability to

almost double of the original value in the absence of ATP. Moreover, using

recombinant channels, it was shown that 10 min exposure of 5 μg/ml of PIP2 to 13 the cytosolic side decreased the ATP sensitivity by 340 times, made the

channel activity stable and increased the run-down time constant. Similar

values (100-700 fold) for shift in ATP sensitivity were obtained in native cardiac

and pancreatic cells. This may in part explains the opening of KATP channels at

physiological ATP levels. It was also shown that the effect is mediated by direct

binding of PIP2 to the Kir6.2 subunit, which is influenced by the SUR subunit. A

negatively charged head group and a lipid tail were found to be essental for effect as inositol trisphosphate (no lipid tail) or phosphatidylcholine (no

negatively charged head group) were ineffective; while PIP2, with three

negative charges, is more effective than PIP (which has two), and PI (which has only one) is ineffective (12, 21, 263). However, the exact kind of interaction between PIP2 and ATP on Kir6.2 subunit is ambiguous, as it has been shown

that the phosphate groups of PIP2 and ATP may compete for the same binding

site (262), but others have shown there are also non-overlapping positive

residues which mediate binding (235, 253). However, using homology

modeling and ligand docking, a recent study proposed that a direct competition

for same binding site is highly unlikely. Instead an allosteric model was

proposed, where PIP2 binding site is just above the ATP binding site, which can explain the experimental observations at large (236).

1.3.3.2. Protein kinase C

Phosphorylation of ion channels (including KATP channels) by kinases

plays an important role in regulating channel function by G-protein coupled 14 receptors (GPCRs). Though a variety of protein kinases have been shown to

affect cardiac KATP channels, the best characterized is the phosphorylation by

PKC (163). PKC is a family of serine-threonine protein kinases that depend on

lipids for activity and consist of atleast 12 isozymes. They are divided into three

subfamilies, based on their structure and second messenger requirements:

conventional or cPKC (α, βI, βII and γ) require Ca2+ and diacylglycerol (DAG)

for their activation, novel or nPKCs (δ, ε, η and θ) require DAG but not Ca2+

whereas atypical or aPKCs (λ, ι, ζ and μ) require neither Ca2+ nor DAG for

activation. GPCR which couple with Gq, activate PLC-β which causes hydrolysis of phosphatidylinositol 4,5-biphosphate into IP3 (Inositol 1,4,5-

phosphate) and DAG (Diacylglycerol). IP3 provides the second messenger of

this pathway by releasing Ca2+ from intracellular stores while DAG remains

embedded in the membrane and either activates the already membrane-bound

PKC isoform or provides the docking site for activated PKC from cytosol (33).

There are many potential PKC phosphorylation sites in both Kir6.2 and

SUR2A subunits (257). Of these sites, phosphorylation at a specific conserved

site, Thr-180 of Kir6.2 is the major site of PKC phosphorylation and is mainly

responsible for stimulatory effect of PKC on KATP channel (163). However,

prolonged stimulation (more than 15 minutes) of PKC can cause internalization

of KATP channels (113). Though, which PKC isoform is responsible for cardiac

KATP channel phosphorylation is not clear, majority of evidence points toward a

crucial role of PKCε. Various studies using KO or transgenic mice using either

deletion or overexpression of PKCε respectively, have clearly established its 15 role in cardioprotective (33). PKCε is the major isoform present in the adut

heart (27), and a membrane-bound, phorbol ester-sensitive,

compartmentalized pool of PKC (distinct from membrane-translocated PKC)

which includes PKCε and PKCδ has been detected in cardiomyocytes even

under non-stimulated conditions (27, 28). On activation, PKCε has been shown

to translocate to Z-line/T-tubule region of myocytes (region rich in ion channels,

anchoring structural proteins and signaling molecules) and bind to it much

more effectively as compared to PKCδ (66, 117, 118, 238, 239). Moreover, a

subset of PKC isoforms including PKCε are present in cardiomyocyte caveolae

at very low levels under basal, and much higher levels under PMA stimulated

conditions (185, 242). In cardiomyocyte, both caveolin-3 and RACK (receptor

for activated C-kinase) proteins can anchor PKC in caveolae. However, the

identity of PKC substrates in caveolar domains is unknown at present. Two

recent studies have shown that for vascular KATP channels, PKCε is responsible for inhibition of KATP current via caveolae dependent mechanism

(128, 244).

Likewise PKCε has also been linked to cardiac KATP channels in signal

transduction pathways mediating ischemic preconditioning (4, 33). Using whole cell patch clamp studies, it was found that PKCε primes the sarcKATP channel in

guinea pig myocytes to open by isoflurane, which is preconditioning-mimetic,

while PKCδ was not that effective (4). A majority of preconditioning studies

implicate mitoKATP channel as the primary candidate for the cardioprotective

phenomenon, and PKCε has been shown to be present endogenously in 16 mitochondria (15, 50) and also translocate to mitochondria (4, 201) on PC

stimulus. Accordingly, it was proposed that mitoKATP might be the primary target

of PKCε. Because of the complexity and redundancy of ischemic

preconditioning pathways, it is not clear whether PKCε is upstream (49, 125),

downstream (201, 293) or in positive feedback loop with mitoKATP. Some

studies have even proposed mitoKATP channel to act as a RACK for activated

or endogenous PKCε in mitochondria (133). Regardless, delineation of the

exact nature of relationship between PKCε and mitoKATP, has to await the

determination of molecular identity of mitoKATP channel (33).

1.3.3.3. Subsarcolemmal ATP

The concept of subsarcolemmal “fuzzy space” was first proposed to

conceptualize the observation that excitation–contraction (E–C) is possible

even in the absence of the L-type Ca2+ current, provided there was influx of Na+

via the Na+ channel. It was speculated that there is a diffusion-restricted

+ + subsarcolemmal space (“fuzzy space”), where [Na ]i is elevated (by Na influx via the Na+ channel) to such a level that NCX is activated in reverse mode

causing Ca2+ influx (19, 155, 156). These observations were later supported by

many studies which have been reviewed recently (19, 285). Similarly, the subsarcolemmal space in cardiac myocytes is an area which is envisaged to be quite different from bulk cytosol in terms of ATP levels depending on rate of production, rate of diffusion and rate of consumption (129, 135). In myocytes, there are many plasma membrane ATPases (e.g. ion channels, ion exchangers 17 and membrane pumps) which maintain the local ATP/ADP ratio quite distinct

from bulk cytosolic levels. Besides, molecular crowding in this space also restricts the free exchange of nucleotides.

In a highly compartmentalized cell like myocyte, elaborate mechanisms

exists which ensure high free energy of hydrolysis of ATP in all compartments

whether center or periphery of the cell. In cardiomyocytes, mitochondria

constitute ~30-35% of the cell volume and are regularly arranged in a longitudinal lattice between myofibrils along the length of the sarcomere (208,

283). Two populations of mitochondria have been described namely, subsarcolemmal (located beneath plasma membrane) and interfibrillar (located between myofibrils). Except for some minor biochemical differences, no major functional or morphological differences were found between these two

populations. It is speculated that interfibrillar mitochondria provide much of the

ATP required for contractile machinery, while subsarcolemmal mitochondria

cater to the membrane pumps and channels (209). To ensure smooth ATP

delivery at all times, muscle cells have also evolved an ATP-shuttling

mechanism via creatine kinase (CK) (ATP+Cr<=>Cr-P+ADP) and adenylate

kinase (2ADP<=>ATP+AMP) which rapidly equilibrate and maintain high ATP

levels in all parts of the myocyte (1, 140).

However, evidence is accumulating indicating that membrane ATPases

including KATP channels preferentially use glycolytic-ATP over mitochondrial-

ATP (65, 298). Studies using CK knockout mice (either mitochondrial or

cytosolic isoform, or both) have shown that these mice have normal cardiac 18 functions under at least moderate workload. This indicates that probably the

dependence on this ATP shuttling system is not absolute, especially for

membrane pumps which are far from mitochondria (249, 250). Secondly, when

cells were perfused with respective substrates for glycolysis, oxidative

phosphorylation or creatine kinase, they supported membrane and contractile

functions with different competencies (23, 174, 297). Finally, many studies

have shown that KATP channels (which act as metabolic sensors) are mainly

regulated by glycolytic ATP and metabolites. It has been shown that in

permeabilized myocytes; glycolytic substrates (with mitochondrial inhibitor) in

the presence of external ATP consuming system were far more competent in

inhibiting KATP channel activation as compared to substrates for mitochondrial

oxidative phosphorylation or creatine kinase system (296). Many glycolytic

enzymes like GAPDH (glyceraldehydes phosphate dehydrogenase),

triosephosphate isomerase, and pyruvate kinase have been shown to be

associated with KATP channel subunits using co-immunoprecipitation, yeast

two-hybrid screen or patch clamp studies (53, 64). Similarly, muscle isoform of

lactate dehydrogenase (M-LDH) (52), adenylate kinase (39) and creatine kinase (53) also associate with KATP channels. These proteins would either bind

directly to KATP channel subunits or indirectly via some scaffolding protein. In

this regard, it is interesting that some of the glycolytic enzymes, namely

phosphofructokinase-M (PFK) and aldolase, have been shown to be present in

caveolae in vascular smooth muscle. Also, in skeletal muscle, this

compartmentalization of PFK-M was shown to involve caveolin-3 interaction 19 and is glucose dependent (229, 252, 267, 282). Since the role of caveolae in

integrating signal transduction pathways is well-known, it is quite possible that

caveolae could be one of the hot spots for metabolic integration of diverse

signals as well (37, 73, 200).

There is a strong evidence implicating physical and functional

compartmentalization of glycolytic enzymes at sarcolemma and sarcoplasmic

reticulum in cardiac muscle and other tissues (65, 91, 145, 222, 307). This

would place the energy generating mechanism closer to the site of utilization,

especially membrane ATPases and metabolic sensors like KATP channels. It is worthwhile to note here that KATP channel microenvironment is considered to

be quite different from bulk cytosol and is secluded from brief oscillations in

cytosolic nucleotides levels (during normal APD). However, if there is any

major change in the cytosolic ATP, the ATP shutting mechanism complements

the metabolic signaling circuit by amplifying and conveying cellular energetic

state to KATP channels (1, 258). An interesting functional interaction between

KATP channel and Na/K-ATPase has been shown, where both compete for the same subsarcolemmal ATP (during metabolic stress) and regulates each other’s activity in the process (106, 135, 280). In fact, literature is replete with similar reports proposing that subsarcolemmal ATP is an important contributor to KATP channel regulation by various receptors like β–adrenergic (251), P2-

purinergic (14), endothelin-1 (295) and angiotensin-II receptor (281).

Interestingly, both pertussis toxin sensitive (Gi, increase ATP by inhibiting adenylate cyclase) and cholera toxin sensitive (Gs, decrease ATP by activating 20 adenylate cyclase) G-proteins were shown to be involved. It is important to realize here that such a regulation by subsarcolemmal ATP may not be tight enough during normal physiological functions but might have implications during metabolic impairment or ischemic conditions when mitochondrial supply of ATP is restricted.

1.3.3.4. Nitric oxide

The role of nitric oxide (NO) in cardioprotection during ischemia and in

ischemic preconditioning is well-known (59, 230). The main source of NO in cardiomyocytes is endothelial nitric oxide synthase (eNOS), and NO released can further upregulate iNOS (inducible). NO activation of KATP channels in cardiomyocytes is thought to involve cGMP-protein kinase G signaling pathway. NO acting via this pathway has been shown to activate sarcolemmal

KATP channels in gold-fish cardiomyocytes using patch clamp recordings (36)

and in intact hearts in various preconditioning studies (16, 186).

Pharmacological evidence also points towards a substantial role of mitoKATP

channels in NO-mediated cardioprotection (48, 247). Whether sarcKATP and mitoKATP are regulated differentially or in parallel by NO is not clear. The source

of NO might play a crucial role in determining which subtype plays an important

role. Although nitric oxide released from endothelium can diffuse to larger

distances and may act in a paracrine fashion, it has been proposed that a small

and local release of NO by eNOS present in cardiomyocyte caveolae (in

21 association with caveolin-3) might act in an autocrine fashion on nearby proteins (40).

1.3.3.5. G-protein coupled receptors (GPCRs)

KATP channels in cardiomyocytes are regulated by a variety of neurotransmitters and hormones via GPCRs namely adenosine (114, 274), bradykinin (196), acetylcholine (274), and adrenergic agents (116, 251). These

GPCRs are coupled to heterotrimeric G-proteins (Gαβγ). Upon activation by ligand, GPCRs promote exchange of Gα-bound GDP for GTP, which in turn favors the dissociation of this trimeric complex into Gα-GTP and Gβγ subunits.

Both of these active signaling molecules modulate various target effector systems and can modulate a majority of ion channels indirectly (via second messenger system pathways) or by direct interaction of ion channel subunits with G-protein subunits without the involvement of diffusible cytosolic factors

(membrane delimited activation by Gα or Gβγ). The Gα subunit has intrinsic

GTPase activity which terminates activation by hydrolyzing GTP to GDP and promoting the reassociation of Gα with βγ subunits.

Indirect G-protein regulation

The G-protein regulation of ion channels via second messengers is the most common and widely investigated pathway. GPCRs regulate multiple second messenger pathways namely: adenylyl cyclase (cAMP-PKA), guanylate

cyclase (cGMP-PKG), phospholipase C (PIP2→DAG/IP3/PKC), phospholipase

22 A (), receptor tyrosine kinase etc. The relevant pathways pertaining to regulation of KATP channels have been discussed before.

Membrane delimited direct G-protein regulation

The membrane delimited G-protein activaton of ion channels has been firmly established only for some members of voltage-activated calcium

channels (Cav) and G-protein activated inwardly rectifying potassium channels

(GIRKs). Interestingly, Gβγ inhibits Cav while it activates GIRKs (Kir3). Since

there are membrane-delimited G-protein dependent pathways like membrane

bound PKC activation, PLC activation and PIP2 generation, a set of criterion

has been proposed to distinguish direct G-protein regulation of an ion channel

(57). These criteria need to be satisfied for an ion channel to be characterized

as a direct G-protein regulated ion channel.

Criterion 1. Purified or over-expressed G-protein subunit replicates the effects

of GPCR-ligand; while deletion or withdrawal of the subunit using binding-

peptide or antibody abrogates the effect of the ligand.

Criterion 2. G-protein subunit act in a membrane-delimited fashion (using

‘inside-out’ or ‘cell-attached’ patch configuration).

Criterion 3. G-protein subunit binds directly to channels protein either in-vitro or preferably in-vivo (using co-precipitation or fluorescence resonance energy transfer (FRET)).

Criterion 4. Identification of a G-protein binding site on ion channel protein

(using mutational or competition assays).

23 Extensive evidence points out a role of pertussis toxin sensitive G-

proteins, Gαi (αi1, αi2, αi3) in the activation of KATP channels via GPCRs in native

cells and recombinant channels in transfected cells (246, 274). Studies using

excised inside-out patches in guinea pig myocytes have shown that GTP (in

the presence of external acetylcholine and adenosine), GTPγS (non-

- hydrolysable analog of GTP), AlF4 (mimics the action of terminal γ-phosphate

of GTP on the GDP-bound α-subunit) and purified G-protein subunits (αi1, αi2,

αo), all antagonize the ATP-dependent inhibition of KATP channels (274). Similar

results were obtained in other studies for Gαi (143, 246), while conflicting evidence was presented for Gβγ showing either inhibition (122) or no effect

(246).

A recent study (287), however, made a strong case proposing a role of

Gβγ for the direct activation of KATP channels. These authors expressed

recombinant channels (Kir6.2/SUR) in COS cells and showed that Gβγ directly

activates KATP channels by reducing ATP inhibition. Using in-vitro binding assay, these authors provided conclusive evidence by mapping the Gβγ

interaction site to both nucleotide binding loops (NBD1 and 2) of SUR subunit.

They also found that a single arginine residue in one of the binding segment

(656 in SUR2A and 665 in SUR1) is essential for the functional effect.

However, it does not mean that Gα-subunit does not play any role.

Since the affinity for G-protein subunits in the excised patches was

found to be very high, 5-10 pM for Gβγ and 30-50 pM for Gαi (much higher than

GIRK channel or adenylyl cyclase, respectively), it has been speculated that 24 “KATP channels might be already saturated by G-proteins at their ‘basal’ levels”

(57). Or it may be possible that KATP channels are compartmentalized with the respective GPCRs within the membrane microdomains ensuring efficient coupling with G-protein subunits. Nonetheless, compartmentalization of many

GPCRs with their heterotrimeric G-proteins and effectors in caveolar microdomains has been shown recently and is firmly established (17, 120,

153).

Adenosine

Adenosine is the main mediator released during ischemia due to alteration in degradation and regeneration of AMP (by adenosine kinase). The main effect is possibly due to decreased regeneration. Adenosine mediated cardioprotection is very well established, though it is debatable which mediator and end-effectors are involved. In addition to the heart, adenosine mediated protection against ischemia has been shown in liver, kidneys and brain also

(60, 188).

Three main subtypes of adenosine receptors (AR) have been identified so far: A1, A2 (A2a and A2b) and A3, and all have been shown to be present in cardiomyocytes (79). Among these subtypes, cardioprotective effect and KATP channel activation is thought to be mainly mediated by either A1 or A3 or both and is a subject of intense debate. Both A1 and A3-AR are coupled to G- proteins (Gαi, Gαo or Gαq) and mediate either inhibition of adenylate cyclase

(AC) or activation of phospholipases and PKC (via IP3/DAG pathway).

25 Most of the evidence for the involvement of adenosine in KATP channel

activation is derived from preconditioning studies where glibenclamide inhibited

the adenosine mediated cardioprotective effect. These studies implicate either

Gαi or PKC as the second messenger in KATP channel activation. However,

PKC hypothesis is the focus of a lot of interest and probably a major contributor

to the cardioprotection and activation of KATP channels, and thus gained much

more prevalence in literature (47, 114, 163, 170, 188). Using KO mice, it was

shown that deletion of A1-AR increases ischemic damage and abrogates the

protective effects of both pre- and post-conditioning while transgenic

overexpression of A1-AR gives additional cardioprotection compared to

respective wild-type strains (151, 179, 303). There is abundant in-vivo and in-

vitro pharmacological data which implicate the involvement of A1-AR in

cardioprotection and KATP channel activation (98, 162, 188, 240, 312). On the

other hand, A3-AR KO mice were more tolerant to ischemic damage and have

intact intrinsic preconditioning response (100, 105). However, pharmacological

data using a recently available selective for A3-AR proved otherwise

(88, 105, 289). There are only few studies which directly investigated the subtype of AR involved in activation of KATP channels using patch clamp

technique. A majority of these studies using selective antagonist showed that

A1 receptor antagonism inhibit the activation of KATP channels by adenosine

(114, 123, 124, 141, 143, 274). One recent study using selective A3-agonist

and A3-AR knockout mice implicates the involvement of A3-AR (289), though it

does not rule out any contribution from A1-AR. Possibly, both of these receptors 26 play a role in the activation of KATP channels, with A1-AR being the major contributor.

However, recent data paint a complex picture of adenosine mediated regulation of KATP channels. Though acute stimulation (~5 minutes) of AR activates KATP channels (114, 162), prolonged stimulation (15-30 minutes) causes internalization of KATP channels (113). Interestingly, A1-AR including the downstream signaling components (Gαi and PKC) has been shown to localize to caveolae in cardiomyocytes. Since, these A1-AR translocate out of caveolae upon 15 minutes of agonist stimulation (17, 153, 185, 242), it can be easily inferred that adenosine mediated regulation of KATP channels is probably more complex.

Noradrenaline

Noradrenaline released through sympathetic nerve stimulation in the heart causes strong ionotropic, chronotropic and dromotropic effects; increasing heart rate, strength of contraction and simultaneous reduction in action potential duration. The adrenergic regulation of many ion channel currents is well established like ICaL, INa, Ito and Ikr. However, there are very few studies investigating the adrenergic regulation of KATP channels, and the exact mechanism is unclear. Using isoproterenol as agonist and as antagonist, it was shown that β-AR (β-adrenergic recptor) can activate KATP channels in feline and guinea pig ventricular myocytes. Since this modulation was not mimicked by cAMP analog and was independent of PKA inhibition, it was proposed that isoproterenol activates KATP current via Gs, by consuming 27 subsarcolemmal ATP through the activation of adenylate cyclase activity (251,

279, 295). As mentioned before, this regulation might have more importance

during ischemic conditions as compared to normal physiological situations

(when ATP supply is abundant). However, a recent study has challenged this

view. Using isoproterenol and exercise challenge as physiological stressors,

these authors showed that though wild-type mice respond with reduced APD and more efficient calcium handling, this adaptive response was absent in

Kir6.2 KO mice. Under the same conditions, Kir6.2 KO mice were vulnerable to

electrical instability, dysregulated calcium handling and ultimately lethal

ventricular arrhythmias. Most importantly, no major shift in bulk

nucleotide content was noted between normal or stressed conditions, indicating

a direct or indirect β-adrenergic mediated activation of KATP channels during

physiological conditions. Though detailed mechanism of KATP activation was

not investigated, a compartmentalized regulation of KATP channels by β-AR was

emphasized (320). Nevertheless, β-AR and Gs (including adenylate cyclase) have been shown to be present in caveolae (107, 108, 305). The presence of

2+ KATP channels in the same microdomains with β-AR and L-type Ca channels

would allow more efficient signaling and direct control of calcium influx (17).

Acetylcholine

All the five subtypes of muscarinic receptors (M1, M2, M3, M4, M5) have

been localized to heart, with M2 being the predominant one which mediates the

negative ionotropic and chronotropic effects. Though acetylcholine (ACh)

mediated Gβγ-dependent activation of IKACh and inhibition of ICaL is well 28 established, the exact role and mechanism of muscarinic receptors in KATP channel regulation is unclear. Previous studies have implicated both PTX- sensitive Gα subunit and PKC-dependent pathways in KATP channels activation

(226, 274). On the other hand, there are reports purporting inhibition of KATP channels by muscarinic receptor stimulation (295) involving similar mechanisms as described before including PKC activation, PIP2 depletion and augmentation of subsarcolemmal ATP (123, 294). As such, the cholinergic regulation of KATP channels is still ambiguous.

Bradykinin

Bradykinin (BK), primarily released from endothelial cells during ischemia, mediates cardioprotection via BK2 receptors on cardiomyocytes (22).

The signaling mechanism of BK in ischemic preconditioning has been linked to activation of KATP channels in mitochondria (218) and sarcolemma (67).

Though direct electrophysiological evidence for KATP current activation by BK is lacking, putative mechanisms include generation of prostanoids and nitric oxide or activation of PKC (22). BK2 receptors have been shown to be present in caveolae in endothelial, smooth muscle and cardiomyocytes (61, 132, 228).

Endothelin

It is believed that during acute myocardial ischemia, the majority of endothelin-1 (ET-1) comes from the heart, as isolated-perfused hearts also release significant amounts of ET-1. The effect of ET-1 on cardioprotection is complex, as it mediates protection at lower concentrations while higher doses have no effect (109). Using patch clamp analysis, previous studies have 29 consistently shown that ET-1 inhibits KATP channels via ETA receptors coupled

to PTX sensitive G-proteins by inhibiting adenylate cyclase activity and thus

increasing subsarcolemmal ATP (248, 295). However, later studies showed

that ET-1 is also a preconditioning mimetic and may act as trigger for ischemic

preconditioning by opening KATP channels subsequent to PKC activation. Direct

involvement of Gαi was also speculated, though the receptor subtype was not

investigated (34, 109). Inline with the lingering controversy for ET-1 role in

cardioprotection, the role of ET-1 in KATP channel regulation also needs

clarification and dissection in terms of specific subtype of receptor involved.

Further, it would be interesting to know if this signaling is compartmentalized or

not, as ET receptors and associated signaling partners have been shown to be

present in caveolae (29, 43).

1.4. Functional role of KATP channels

It’s been known for a while that KATP channels in the pancreas mediate

insulin secretion. Pancreatic KATP channels remain open under the basal

conditions and close when plasma glucose levels increase leading to insulin

secretion. Surprisingly, cardiac KATP channels are closed under the basal conditions. However, they open-up during metabolic stressed conditions (like ischemia) and play significant role in outward potassium influx and resultant action potential duration (APD) shortening. This is expected to lead to decreased calcium influx and decreased calcium overload and resulting energy sparing will have cardioprotective effect (93, 194). The APD shortening effect of 30 KATP channels during metabolic stress has been confirmed in Kir6.2 knockout

mice also (270). Since KATP channel are sensitive to micromolar amount of

ATP, the exact mechanism leading to KATP channel opening under ischemic conditions (when intracellular levels of ATP are still in millimolar range) is not very clear. The putative mechanisms and associated signaling are discussed in

detail in section 1.3.3.

The protective effect of KATP channel opening during

ischemia/reperfusion has been demonstrated using KCOs and selective KATP

channel blockers (glibenclamide) also. Interestingly, ‘ischemic preconditioning’,

a cardioprotective phenomenon can be mimicked by KCOs and inhibited by

sulphonylureas (47). Ischemic preconditioning (IPC) is the protection of

myocardium from prolonged ischemic damage by a prior, brief and reversible

ischemic insult. The phenomenon originally discovered by Murry in canine

model in 1986 (189) has since been shown to occur in almost all other species,

including humans, and can be demonstrated in single cells and in isolated

perfused hearts in-vitro. Earlier, much emphasis was on two main types,

namely, early or classical preconditioning (effective within 2-3 hours) and

delayed or late preconditioning (effective between 24 to 96 hours). However,

recent intense effort is being directed towards a new-type called ‘post- conditioning’, which is thought to be more clinically relevant (207). Originally

shown as a reduction in infarct size in the heart, other end-points of IPC

protection include decrease in apoptosis and necrosis, decrease in post-

ischemic arrhythmic events, improvement in post-ischemic dysfunction and 31 decrease in stunning (a reversible and temporary loss of contractile ability that remains persistently evident despite restoration of coronary blood flow). There is wide variability and conflicting results in literature in terms of many correlates of IPC like duration and timing of preconditioning stimulus; duration of protection window; signaling pathways mediating IPC and; end-points of protection afforded by IPC. This variability could be due to interspecies variability and different model systems used in various studies (317). Because of redundancy and involvement of large number of signaling pathways in IPC

(207, 317), a comprehensive discussion is not possible here. So, I will mainly focus on relevance of KATP channels signaling in IPC. Since the discovery of the IPC, KATP channels have emerged as one of the prime candidates mediating the protective effects of IPC. It has been shown in many studies that

KCOs can mimic while glibenclamide can abolish the protective effects of preconditioning (94, 149, 150). Many mediators released during ischemia, like adenosine, bradykinin, prostacyclin and nitric oxide, have been shown to activate or facilitate opening of KATP channel (36, 114, 116, 196, 251, 274). The activation of PKC which is considered to be the key player for many signaling pathways (47) has also been linked to activation of KATP channel (discussed in section 1.3.3.2). Recently, Kir6.2 knockout mice have been shown to be resistant to the protective effects of preconditioning, indicating their importance in IPC. However, these knockout mice had more intense and rapidly developed contractures and decreased functional recovery in the absence of IPC stimulus than the wild-type controls. Though this indicates their prime role in protection 32 during ischemia-reperfusion, it is possible that ischemic injury is artefactually

enhanced and IPC effect is nullified (270). Nevertheless, there are studies

showing negligible role of KATP channels in IPC (169). As will be discussed

later, evidence also implicates the presence of another isoform of KATP

channels in mitochondria (mitoKATP) (119) which might mediate the protective effects of IPC either alone or alongwith sarcolemmal KATP (sarcKATP) channel .

Since most of the studies were done in healthy and young animals, it is further

debatable whether IPC can occur or is preserved in senescent or diseased

heart (207).

The role of KATP channels in arrhythmias is probably the most

controversial aspect currently. Since KATP channels shorten APD, they should be antiarrhythmic as they hyperpolarize the membrane, decrease calcium overload and decrease ischemic injury. Paradoxically, there is always a theoretical possibility for reentrant arrhythmias on the background of ischemic substrate, though experimental evidence is limited (41). It can be argued that the propensity of KATP channel to act as antiarrhythmic or arrhythmogenic

depend on the type of arrhythmias whether long QT arrhythmias or reentrant

arrhythmias respectively (190, 234). As mentioned earlier, SUR1 is recently

shown to be an essential component of sarcolemmal KATP channels in atrial

myocytes but not in ventricular myocytes in mice. Same group also showed that transgenic mice overexpressing SUR1/Kir6.2 containing KATP channels in

the heart were prone to arrhythmias including atrial fibrillation. Based on the

findings, it was speculated that SUR2 is cardioprotective while SUR1 is 33 arrhythmogenic (77). Nevertheless, several studies have demonstrated the protective effects of KCOs in prevention of post-ischemic arrhythmias (46, 134,

173).

Even though the role of cardiac KATP channel in ischemic conditions is well-characterized, their role under normal physiological conditions is just beginning to unravel. Recent studies (320) have shown that KATP channels in the heart may not be as silent as assumed previously. Though Kir6.2 knockout mice seem to develop normally, these mice have significantly less exertional capacity; less cardiac performance, disruptive calcium handling and increased vulnerability to lethal ventricular arrhythmias under sympathetic challenge. The metabolic demands of the sympathetic stimulation put a significant demand on the energy resources of the heart (increased cellular Ca2+ overload/ handling and associated energy depletion). To maintain proper cellular homeostasis, this requires some degree of fine tuning between the adequate escalation of Ca2+ influx and increase in compensatory outward hyperpolarizing current (via KATP current). This indicates that KATP channels play an important role of homeostatic control in adaptation to acute stress (168, 320).

Sarcolemmal vs. mitochondrial KATP channels

A detailed dissection of KCOs action in preconditioning (PC) led many to believe that channels on the sarcolemmal membrane (sarcKATP) may not be the only site of action of these drugs. Using a low-dose KCO , it was established that cardioprotection is independent of APD shortening in dogs

(311). Further, protection can be shown in isolated non-beating cardiomyocytes 34 in culture (10). Much of the attention shifted to mitochondria, once it was discovered that there are KATP channels (mitoKATP) in inner membrane of mitochondria (IMM) (119). Using proteoliposomes containing reconstituted KATP channels from bovine heart mitochondria, it was found that KCO diazoxide is

2000 times more selective for mitoKATP channel as compared to sarcKATP channels (87). A selective mitoKATP inhibitor, 5-hydroxy decanoate (5-HD), was also identified. HMR1098, which is selective for cardiac sarcKATP channels, was found to be minimally effective on mitoKATP. Based on these findings, two groups independently showed that mitoKATP may be the prime target of KCOs induced PC (86, 171).

Most of the evidence available for the existence of mitoKATP is based on: mitoKATP selective pharmacological agents; use of antibodies; and three main techniques to study its properties namely, patch clamp, flavoprotein fluorescence and spectrophotometeric measurement of mitochondrial swelling.

Many properties of the mitoKATP resemble those of sarcKATP channels, although some significant differences have been noted-especially the selectivity of respective pharmacological agents (198). MitoKATP channels exhibit similar inhibition by ATP (Ki ≈ 0.8 mM) when applied from the matrix side and exhibit sensitivity to 4-amino pyridine (a K+ channel inhibitor) and glibenclamide.

MitoKATP channel conductance was found to be well below the sarcKATP channel conductance (10 pS in 100 mM cytosolic and 33 mM matrix K+) (119).

However, a recent report using reconstituted human inner mitochondrial membrane (IMM) in lipid bilayer found that there could be multiple conductance 35 states depending on the concentration of ATP and other nucleotides (127).

Over the years, many investigations have been carried out to elucidate the

molecular composition of mitoKATP channels without any concrete results. Since

the subunit composition of sarcKATP channels is known for a while, it was

hypothesized that probably mitoKATP might share the same composition. Using

antibody approach, it has been shown by many that varied amounts of both

Kir6.2 and Kir6.1 are present in cardiac mitochondria, but of course, with

conflicting results (54, 147, 256, 265, 268). A recent study investigated specificity of the currently available commercial Kir6.1 antibodies using mass spectrometer proteomics approach. They concluded that most of them bound

only to non-specific mitochondrial proteins at best and could be one of the

reasons for discrepancy in literature (78). However, using DEAE-affinity

column, a 54 KDa protein was isolated from mitochondrial fraction and was

tentatively identified as mitoKATP subunit (216). Further, low affinity binding sites for sulphonylureas have also been identified in purified mitochondrial fractions.

It is not clear whether these binding sites represent the already identified SUR

clone or other minor splice variant (97, 272).

The physiological role of mitoKATP (198) is thought to be mitochondrial

matrix volume control via counterbalancing mitochondrial uniporter activity with

K+/H+ exchange. This volume control is thought to be important for controlling the redox-state of the mitochondria (K+ influx is supposed to decrease

mitochondrial potential and decrease Ca2+ overload) and also prevent bursting

of the mitochondrial during pathological swelling as happens 36 during ischemic-reperfusion injury. Another leading hypothesis is the

generation of reactive oxygen species (ROS) after mitoKATP opening which can

further activate many other protective signaling pathways including activation of

PKCε which can also be redox-activated (discussed in section 1.3.3.2). It is

hypothesized that mitoKATP opening causes increased protective ROS during

PC and inhibits pathological post-ischemic ROS generation. A recent study

investigating the mechanism of PC has given credence to the hypothesis that

mitoKATP channel opening and consequent mitochondrial swelling and ROS

generation are important components of the memory of PC (133).

However, it does not mean that sarcKATP channels do not play any role.

Concerns have been raised on the selectivity of so called ‘mitoKATP selective’

pharmacological agents (198, 317). These mitoKATP-selective agents have

been shown to act on many other mitochondrial targets significantly affecting

the function of mitochondrial respiratory enzymes (68, 103). Possibly, both sarcoplasmic and mitochondrial KATP channels play complementary roles in the protection of intact beating heart (96). Nevertheless, 17 years after its discovery (119), a uniform consensus for the molecular identity of mitoKATP is

not established and has to await cloning of the channel subunits.

37 2. Caveolae

2.1. Caveolae and caveolins

Discovered in 1950s (309) as the 50-100 nm flask-shaped invaginations

of the plasma membrane, caveolae, are now firmly established as a subcellular

organelle. Caveolae are a special type of lipid rafts (sterol- and sphingolipid-

enriched domains of the cellular membrane) tagged on the inner-leaflet of the membrane by caveolin proteins. Caveolae are found in almost all cell types

(except neurons and lymphocytes) and are most abundant in terminally- differentiated epithelial cells, endothelial cells, adipocytes and muscle cells

(232).

Though the techniques to specifically label raft structures for imaging are

controversial (121), caveolae can be labeled by their marker proteins,

caveolins. Caveolae are considered to be insoluble in non-ionic detergents at

low temperature because of the tight packing of cholesterol and spingolipids.

Since detergents can promote non-physiological interactions by raft-fusion, detergent-free extraction procedures are currently employed to isolate caveolae from cellular fractions (261, 266). They can be isolated by their buoyant density on sucrose-gradient.

There are four isoforms of caveolins: Cav-1 (α and β), 2 and 3, all of which have been shown to be present in cardiac myocytes. They are all small proteins of molecular weight ranging between 20-25 KDa. Cav-1 and 3 are most identical (65%) while Cav-2 is most divergent (38% identical to human

Cav-1). Cav-2 is dependent on Cav-1 for full expression and can not form 38 caveolae on its own. Cav-3 is the predominant isoform present exclusively in muscle tissue only (107, 273). Deletion of Cav-1 in most cell-types and specifically Cav-3 in muscle cells leads to loss of caveolae in respective cell- types. The molecular structure of caveolins is unusual with N and C terminus in the cytosol and a hairpin domain which traverses only the inner-leaflet of membrane (Fig. 1.3). They are synthesized as membrane proteins in endoplasmic reticulum (ER) and once in the golgi-complex go through first stage of oligomerization (~14-15 monomers). At some point during the exit from golgi, they associate with cholesterol and arrange themselves into higher order detergent-resistant complexes (complexes of ~10-11 oligomers) of ~350-400

KDa (210). Besides plasma membrane, studies have found Cav-1 in some intracellular structures like secretory vesicles of some exocrine and endocrine cells and also in mitochondria from airway epithelial cells (161). A detailed investigation for subcellular location of Cav-3 is lacking but it is thought to play an important role in developing T-tubules. Nevertheless, in adult cardiac myocytes, caveolin-3 is present in T-tubules and sarcolemma (107, 211, 266).

39

Figure 1.3. Topology and predicted membrane insertion of Cav-1. For simplicity a dimer of Cav-1 is shown. Both N- and C-terminal have membrane attachment domains and are exposed to cytosol. Most of the C-terminus is associated with membrane and is extensively palmitoylated as shown. The yellow color hairpin loop can be seen traversing the lipid membrane, but is not exposed to extracellular side. The position of the caveolin scaffolding domain

(CSD) is also shown. Adapted from Williams and Lisanti, 2004 (300).

40 2.2. Functional role of caveolae/caveolins

Vesicular trafficking

Caveolae play an important role in and of some

proteins. They are relatively immobile and stable structures on the plasma membrane. However, this immobility is significantly altered by cholesterol

depletion and various agents. Some viruses like SV40 and polyoma virus as

well as sterols and glycosphingolipids can initiate caveolar budding and are

endocytosed by caveolae. Interestingly, many components of endocytosis are

shared with other endocytic pathways like requirement of dynamin, actin

cytoskeleton and Rab5 etc (217, 259). After endocytosis, these lysosome-

resistant caveosomes (equivalent to endosome) can be directed towards a

variety of destinations, including recycling back to the plasma membrane.

Though a preferred route of endocytosis for the above mentioned agents or

some of the caveolar-resident proteins (CRPs), CRPs are not entirely

dependent on caveolae and can follow other endocytic pathways in their

absence (142).

Lipid homeostasis

Caveolins have high affinity for, and bind to cholesterol and fatty acids.

Cholesterol depletion/binding agents like methyl-β-cyclodextrin (MβCD), nystatin or filipin can flatten caveolae (175, 241). Cav-1 has been found to be associated with lipid droplets, and Cav-1 overexpression has been shown to increase fatty acid uptake and increase free cholesterol levels and export (80,

181). Since, caveolae are extremely abundant in adipocytes, the main storage 41 sites for fats, Cav-1-/- mice show decreased cholesterol in lipid droplets in

adipocytes. These mice show an interesting lean phenotype, have defect in

insulin-stimulated lipogenesis, and are resistant to atherosclerotic lesions or

diet-induced obesity (37, 231). These mice have highly impaired liver

regenerative response, with reduced lipid droplets in hepatocytes and low

survival rate, indicating the potential importance of Cav-1 in lipid homeostasis

(74). Further, cholesterol levels have been shown to control Cav-1 expression

(102). Thus, a unique relationship exists whereby both regulate each other’s

level of expression.

Tumor suppression

The role of caveolin in tumor suppression dates back to 1989 when Cav-

1 was first discovered as a tyrosine phosphorylated protein from v-Src-

transformed fibroblasts (89). It was later established that transformation of NIH

3T3 cells leads to decreased Cav-1 expression and loss of caveolae.

Overexpression of Cav-1 in H-Ras G12V-transformed cells can reverse transformation and prevent anchorage-independent growth (144). Accordingly,

Cav-1 is thought to recruit and organize proteins at focal adhesions for substrate attachment (62). The phenotype of Cav-1-/- mice support their role in

tumor suppression. Though these mice do not form spontaneous tumors, they

are highly vulnerable to oncogenic stimulus (38). This illustrates their role in

cell-cycle regulation and transformation.

42 Compartmentalized signal transduction: caveolae as signalosome

The ‘caveolae/raft signaling hypothesis’ was first proposed (165) based on the observation that a large number of signaling molecules are present in caveolae at much higher concentrations as compared to bulk plasma membrane. It states that caveolae serve as structural platform to specifically segregate and concentrate selective signaling molecules. It was hypothesized that this will lead to more efficient, selective and regulated signal transduction events instead of random chance interaction of signaling molecules. Over the years, most of the findings have corroborated this hypothesis to a large extent as an array of proteins like many G-protein coupled receptors (including their downstream components), kinases, adaptor proteins, structural proteins and ion channels have been preferentially localized to these membrane microdomains (120, 232, 316). Some of those molecules worth mentioning here are adrenergic receptors (β2>>β1 and α1), bradykinin receptor (B2),

adenosine receptor (A1), muscarinic receptor (M2), PKC (α, β and ε), PKA,

eNOS and Phosphoinositides and PI-kinases. Caveolin proteins act as

scaffolding proteins for many of these and regulate their functions. After it was

shown that they inhibit G-proteins in GTP hydrolytic assays (160), many other

proteins like eNOS, PKC isozymes, Src family tyrosine kinses and H-Ras were shown to be inhibited by interaction with caveolins. Though there are few

exceptions, the general inhibitor role goes well with the tumor suppressor role

of Cav-1 (214, 232).

43 An important question would be what mediates the localization of these

proteins to caveolae or lipid rafts. Though there is no absolute criterion, some

generalities have been found in many of these proteins. Many of these

caveolar-resident proteins (CRPs) have lipid modifications (lipid-protein

interactions); protein-protein interaction with caveolar resident scaffolding

protein; direct interaction with caveolins (120, 177). Till now experimental

evidence is limited to and has favored the third hypothesis. A scaffolding

domain in caveolin-1 (CSD) has been identified (160) which is capable of

binding with a broad array of signaling molecules. Analogous to Cav-1, a highly homologous CSD is present in Cav-3. Using CSD of Cav-1 as bait, consensus motifs were identified in some of the caveolar resident proteins termed as caveolin binding domains (CBD). These motifs are: ΦXΦXXXXΦ,

ΦXXXXΦXXΦ, ΦXΦXXXXΦXXΦ, where Φ means an aromatic residue (Phe,

Tyr, or Trp), and X means any amino acid. Most of the CRPs have at least one

copy of any of these motifs, indicating that this may be the site of direct

interaction (51). Nevertheless, presence of CBD does not mean that the protein

resides in caveolae, and absence does not negate caveolar localization.

Probably the most interesting insight in the caveolar field has come from

the investigations of dynamic interaction of GPCRs with caveolae before and

after ligand stimulation. It was shown that β2-adrenergic receptor and A1-

adenosine receptor (preferential CRPs) exit caveolae after agonist stimulation

(153, 205). On the other hand, the agonist stimulation of B1-, B2-bradykinin

receptors and M2-muscarinic receptor translocate them into caveolae (75, 243). 44 This was further shown to be important for their respective fates after

stimulation like internalization and termination or activation of signaling. This

kind of dynamic interaction was later found to have general applicability to

other signaling molecules too and further established the role of caveolae as

dynamic signalosomes (232).

2.3. Compartmentalization of ion channels

2.3.1. Cardiac ion channels

Though Cav-3 is the predominant isoform expressed in cardiac, skeletal

and vascular myocytes, current evidence has now established that Cav-1 and -

2 are also present in significant amounts in all three cell types (108, 237).

Though it is not clear whether Cav-1 and Cav-3 have over-lapping roles or not,

Cav-3-/- mice lack identifiable caveolae from muscle cells. These mice show severe skeletal muscle abnormalities with muscle degeneration and T-tubule abnormalities (83, 101). These mice also develop severe cardiomyopathy by 3-

4 months of age, with hypertrophy, dilation and reduced fractional shortening.

Analogous to the inhibitory role of Cav-1 in other cells, Cav-3-/- mice show

hyperactivation of the Ras-p42/44 MAPK cascade in cardiomyocytes (301).

In cardiac myocytes, compartmentalization of many ion channels or

exchangers (like voltage dependent Na channel (hNav1.5), voltage dependent

+ 2+ potassium channel (Kv1.5), the Na /Ca exchanger, the pacemaker channel

(HCN4), Na+/K+-ATPase, L-type Ca2+ channel) and signaling molecules (like G-

protein coupled receptors; G-proteins: Gαs, Gαi, Go, Gq; various types of 45 receptor and non-receptor kinases, GTPases, cellular proteins/adaptors,

structural proteins and other enzymes) have been demonstrated in lipid rafts

and caveolae (17, 18, 30, 177, 313). Cardiac muscle caveolae are very rich in

many of the already mentioned signaling molecules: adrenergic receptors (β2-

AR and α1-AR), bradykinin receptor (B2), adenosine receptor (A1), muscarinic

receptor (M2), PKC (α, β and ε), PKA, eNOS, nNOS, PIP and PIP2 (30, 35, 167,

175, 232). Caveolin-3 is present in T-tubules and sarcolemma of cardiac

myocytes (107, 211, 266) and modulates the function of many ion channels by

direct interaction.

The earlier work utilizing immunogold electron microscopy suggested the

lipid raft localization of sodium channels in electric eel and 1,4-dihydropyridine

receptor in skeletal muscle caveolae in rabbit (70, 130). Later on, a variety of

other ion channels from different ion channel families including potassium and

chloride were also found to localize to caveolae (175, 177). This raft/caveolae

localization of several ion channels was found to profoundly affect their

function. Since methyl-β-cyclodextrin (MβCD) depletes membrane cholesterol

and disrupts caveolae, it was extensively employed (along with other

cholesterol binding agents like nystatin, filipin) to study the functional effect

caveolar localization of ion channels. It was shown that MβCD treatment

decreased the frequency, amplitude and spatial size of Ca2+ sparks in

myocytes, indicating that caveolae might be the site for their formation (172).

MβCD treatment also decreased the amount of Na+/Ca2+ exchanger protein

(which has five potential CBDs) that can be co-immunoprecipitated with Cav-3 46 antibody (30). An interesting interaction between caveolae and cardiac sodium

channel (INa; Nav1.5) has been shown. Using Cav-3 antibodies it was shown

that Cav-3 inhibited the direct G-protein activation of sodium channels by

isoproterenol. It was concluded that caveolae increase the surface presentation

of sodium channels in response to β-AR receptor activation (313). Further, in

addition to MβCD treatment, siRNA against caveolin-3 also disrupted the β2-

2+ adrenergic activation of L-type Ca channel current while β1-mediated

activation remained intact (17).

Recent investigations have found that caveolins do play a direct role in

regulation of some ion channels, and this regulation is independent of caveolar

structure. Using Cav-1 scaffolding domain (CSD) in pipette solution, it was

found that Cav-1 exerts a strong inhibitory effect on Ca2+-activated K+ channels

(BKCa) in vascular endothelial cells (290). Following these observations,

another group showed that in human myometrial smooth muscle cells, actin

cytoskeleton is part of caveolar-complex regulating BKCa channels (31).

Overall, available data indicate the importance of caveolae and caveolin proteins in normal physiological functioning of ion channels and compartmentalization with respective modulators. Since dysregulated cell surface expression and functioning of ion channels is increasingly linked to ion channelopathies, understanding the lipid regulation of membrane excitability might provide new therapeutic avenues for those diseases.

47 2.3.2. KATP channels

Intriguingly, there are reports indicating that KATP channels in other tissues

do localize to caveolae. In vascular smooth muscle cells, KATP channels consist of Kir6.1 and SUR2B. These channels are tonically activated by cAMP- dependent PKA because of the constitutive activity of adenylyl cyclase. This

KATP current is essential in giving vasodilating drive to resting muscle tone

(245). It was found that caveolae integrity is essential for cAMP-dependent,

protein kinase A (PKA)-activated component of KATP current (Kir6.1/SUR2B).

Since diffusion of cAMP to distant sites is restricted by phosphodiesterases, it

was postulated that caveolae help in compartmentalizing the regulatory

pathways including adenylyl cyclase with vascular KATP channels. Further, it

was noticed that MβCD was unable to disrupt the interaction between Kir6.1

and Cav-3 indicating involvement of direct protein-protein interaction (245).

For pancreatic KATP channels (Kir6.2/SUR1), the evidence is less clear. In hamster pancreatic β-cells where Cav-1 is localized to both plasma membrane and insulin granules, it has been shown that both lipid rafts and caveolae are involved in insulin secretion (191). In a pancreatic β-cell line, it was shown earlier that both Kir6.2 and SUR1 are present in bulk-membrane fraction after detergent-free extraction and sucrose-gradient ultracentrifugation (304).

However, a recent study (225) demonstrated a direct protein-protein association between Cav-1 and Kir6.2 in a mouse pancreatic β-cell line.

Further, using siRNA mediated Cav-1 knockdown, same authors also showed that Cav-1 is essential for sulphonylurea or glucose induced insulin secretion 48 (225). Whether, the difference in two studies is related to species difference or

method of analysis is not clear. Other possibilities could be that KATP channels may associate with Cav-1 outside of caveolae; or the association and localization may depend on specific stimulation.

3. Objectives

ATP-sensitive potassium channels (KATP channels) play an important role in

cardioprotection under stress conditions. They link the cellular energetic state

to membrane excitability, and thus control Ca2+ influx. These channels are

evolutionary conserved across animal species and show protection under

various kinds of stressful conditions, especially under ischemic conditions.

Identification and elucidation of their subcellular localization and mechanism of

regulation in cardiac myocytes will be potentially useful in understanding the

endogenous cardioprotective mechanisms of the heart and their alteration in

some pathological conditions. Therapeutic strategies can be further designed

to enhance their cardioprotective regulatory mechanisms.

Accordingly, the objectives of the present investigation were three-fold,

each of which served as a basis for a submitted or published manuscript:

Hypothesis 1: Cardiac KATP channels (Kir6.2/SUR2A) reside in caveolin

enriched membrane microdomains and are regulated by caveolae-dependent adenosine receptor signaling.

49 Hypothesis 2: Caveolin-3 can functionally modulate the activity of KATP channels (Kir6.2/SUR2A).

Hypothesis 3: Kir6.2 containing KATP channels reside in mitochondrial

inner membrane and can be regulated by PKC activation.

50

CHAPTER 2

REGULATION OF ATP-SENSITIVE POTASSIUM CHANNELS BY

CAVEOLIN-ENRICHED MICRODOMAINS IN CARDIAC MYOCYTES

ABSTRACT

ATP-sensitive potassium (KATP) channels in the heart are critical regulators of

cellular excitability and action potentials during ischemia. However, little is

known about subcellular localization of these channels and their regulation. The

present study was designed to explore the potential role of caveolae in the

regulation of KATP channels in cardiac ventricular myocytes. Both adult and

neonatal rat cardiomyocytes were used. Subcellular fractionation by density

gradient centrifugation, Western blotting, co-immunoprecipitation and

immunofluorescence confocal microscopy were employed in combination with

whole-cell voltage clamp recordings and siRNA gene silencing. We detected

that the majority of KATP channels on the plasma membrane of cardiac myocytes were localized in caveolin-3-enriched microdomains by cell fractionation and ultracentrifugation followed by Western blotting.

Immunofluorescence confocal microscopy revealed extensive colocalization of 51 KATP channel pore-forming subunit Kir6.2 and caveolin-3 along the plasma

membrane. Co-immunoprecipitation of cardiac myocytes showed significant

association of Kir6.2, adenosine A1 receptors and caveolin-3. Furthermore,

whole-cell voltage clamp studies suggested that - mediated activation of KATP channels was largely eliminated by disrupting

caveolae with methyl-β-cyclodextrin or by small interfering RNA, whereas

pinacidil-induced KATP activation was not altered. We demonstrate that KATP

channels are localized to caveolin-enriched microdomains. This microdomain

association is essential for adenosine receptor-mediated regulation of KATP

channels in cardiac myocytes.

52 1. Introduction

The ATP-sensitive potassium (KATP) channel is a highly abundant

plasma membrane protein responsible for linking the cellular energy levels to

membrane potentials (2, 197). In the heart, KATP channels on the plasma

membrane are critical regulators of cellular excitability and action potentials

during ischemia. They are closed under the basal conditions due to inhibition by

intracellular ATP, but open under metabolic stress such as ischemia or hypoxia.

Cardiac KATP channels have been shown to protect against the metabolic insult of ischemia and contribute to adaptive responses to metabolic stress (139, 193,

212).

It’s been known that cardiac KATP channels open at millimolar range of intracellular ATP during ischemia while ATP required to inhibit this channel is in micromolar range (69, 197). This discrepancy is probably due to the fact that some signaling molecules, which are released during ischemia, regulate ATP- dependent gating of the KATP channels by decreasing ATP-sensitivity and thus

increasing open probability (36, 112, 263, 274). Since many signaling

molecules and their downstream mediators/effectors concentrate in the

microdomains of the plasma membrane, the localization of KATP channels to

these sites may have important implications for channel function and regulation.

For rapid and efficient activation of KATP channels during metabolic stress, the

spatial localization of KATP channels should be of immense importance. It is

likely that the KATP channel signaling is compartmentalized.

53 Caveolae are specialized plasma membrane microdomains involved in

numerous signaling transduction events. Caveolins are the main structural

components of caveolae that can bind cholesterol and coat the cytoplasmic

surface of these organelles. They comprise a family of three distinct 21- to 24-

kDa isoforms; caveolin-1 and -2 are almost ubiquitously expressed, whereas

caveolin-3 is a muscle-specific isoform (8, 232). In the present study, we tested

the hypothesis that subcellular localization of KATP channels to caveolin-

enriched microdomains on the plasma membrane of cardiac myocytes is

essential for regulation of KATP channels by adenosine receptors.

2. Materials and Methods

Cell Isolation

The investigation conforms to the Guide for the Care and Use of Laboratory

Animals published by the US National Institutes of Health (NIH Publication N0.

85-23, revised 1996). Adult rat ventricular myocytes were isolated from

Sprague Dawley (SD) rats (250 to 300 g) by enzymatic dissociation (112). In

brief, hearts were excised and retrogradely perfused via aorta with Tyrode's

solution containing (in mM) NaCl 136, KCl 5.4, CaCl2 1.0, MgCl2 1.0, NaH2PO4

0.33, HEPES 10, and glucose 10 at 37°C. The perfusate was then changed to a

Tyrode's solution that was nominally Ca2+ free but otherwise having the same

composition. The hearts were perfused with the same solution containing collagenase for 20 min. Softened ventricular tissues were removed, cut into small pieces, and mechanically dissociated by trituration. The digested cell 54 suspension was gently centrifuged, after which the supernatant was removed

and the remaining pellet was resuspended in a storage solution containing

(mM) KCl 20, KH2PO4 10, glucose 10, potassium glutamate 70, ß- hydroxybutyric acid 10, taurine 10, mannitol 5, and EGTA 5, along with 1%

albumin.

Neonatal myocytes were isolated from 1- or 2-day-old SD rats by

collagenase digestion as described previously (113). Hearts were removed and

collected in ice-cold Hank’s balanced salt solution (HBSS) and rinsed free of

excess blood. The tissue was minced and collected in 1 ml of fresh HBSS to

which 1 ml of 2 mg/ml collagenase type 2 was added. A total of seven rounds of

5-min digestions were performed at 30°C on a rocking platform. After each digestion, the tissue was gently aspirated up and down three times through a

10-ml pipette and allowed to settle for 1 min. The supernatants of the first three digestions were discarded. The supernatants collected from digestions 4-7 were collected, adjusted to 20% FCS, and centrifuged at 400 x g for 5 min. The cell pellets were immediately resuspended in 1 ml DMEM containing 10% FCS.

After preplating for 1 hr in a tissue culture grade petri dish at 37°C and 5% CO2

to allow for the attachment of fibroblasts, cell suspension was plated onto

collagen-coated coverslips. On second day, -β-D-arabinofuranoside (10

µg/ml) was added to the medium to inhibit fibroblast growth. Adherent viable

myocytes started to contract spontaneously 24 hr after plating. Nonmyocyte

cells were less than 10% of total cells.

55 Small Interfering RNA (siRNA) and transfection

The siRNA oligonucleotide targeting caveolin-3 was purchased from

Ambion Inc (Austin, TX, USA). A negative control siRNA (scrambled) was

included to monitor non-specific effects. Neonatal ventricular myocytes were transfected with siRNA and pEGFP using Amaxa kit (Amaxa, Gaithersburg,

MD) immediately after preplating step. Forty-eight to 72 hr after transfection,

Western blot was carried out to examine the knockdown of targeted proteins.

Fractionation of Caveolin-enriched Membrane

Caveolin-rich fractions from cardiomyocytes were prepared by using a

previously described detergent-free method with some modification (266).

Briefly, cell pellet was resuspended in 0.5 M sodium carbonate (pH 11.0; 2 ml) and homogenized sequentially by using a loose-fitting Dounce homogenizer, a

Polytron tissue grinder and a sonicator. The homogenate was adjusted to 45%

sucrose by addition of an equal volume of 90% sucrose in MBS (25 mM Mes,

pH 6.5/0.15 M NaCl) and placed at the bottom of an ultracentrifuge tube. A 5–

35% discontinuous sucrose gradient (in MBS containing 250 mM sodium

carbonate) was formed above, by overlaying with 4 ml of 35% sucrose

(prepared in MBS with 250 mM sodium carbonate) and then 4 ml of 5% sucrose

(again prepared in MBS with 250 mM sodium carbonate). Tubes were

centrifuged at 39,000 rpm for 18-20 hr in a SW41 rotor. Twelve 1-ml fractions

were collected from the top to the bottom of the gradient for subsequent

analysis by Western blot. 56 Western Blotting

Immunoblot analysis was carried out as described previously (84). The

density gradient fractions or cell lysates were denatured in a sample buffer,

electrophoresed on 10% SDS-polyacrylamide gels and transferred onto

nitrocellulose membranes. The transferred blots were blocked with 5% nonfat milk in Tris-buffered saline (TBS, 150 mM NaCl, 20 mM Tris-HCl, pH 7.4) and

incubated for 1 hr at room temperature with primary antibodies in TBS, 0.1%

Tween 20. After washing, the blots were reacted with peroxidase-conjugated secondary antibodies for 45 min and developed using the ECL detection system.

Co-immunoprecipitation

Immunoprecipitation experiments were performed as reported previously

(128). Adult rat ventricular myocytes or neonatal rat hearts were homogenized

in ice-cold lysis buffer (50 mM Tris-HCl, 150 mM NaCl, 1 mM EDTA, 2 mM

EGTA, pH 7.6) containing 1% Triton X-100 and 1 mM sodium orthovanadate

with protease inhibitor tablets (Roche). Insoluble materials were pelleted by

centrifugation at 10,000 x g for 15 min and the lysates were precleared by

incubation with r-protein-G agarose (Invitrogen; 1 hr, 4°C). 500 µl of cleared

lysate (1 µg/µl) was incubated with 5 µg of primary antibodies or 5 µg of

appropriate control IgG (or without primary in some cases) overnight at 4°C.

Antigen-antibody complexes were captured with r-protein-G agarose (4°C, 2

57 hr). Agarose beads were washed 4-times with solubilization buffer before

removal of bound proteins by boiling in SDS sample buffer. In some experiments, co-immunoprecipitation was performed in caveolin-rich fractions.

The fraction 4, 5 and 6 from sucrose-gradients centrifugation were combined,

diluted by adding MBS and centrifuged at 40,000g for 2 hr to pellet caveoale,

which was then suspended in lysis buffer and sonicated prior to immunoprecipitation. Samples were resolved by SDS-PAGE, transferred onto

nitrocellulose membrane, and analyzed by probing with various antibodies.

Immunofluorescence

As described previously (84), the cells were fixed with 4% formaldehyde

in PBS for 30 min, blocked, permeablized in 5% goat serum in PBS with 0.1%

Triton X-100 (30 min), and labeled with primary antibody for 2 hr. Cells were

then washed three times and labeled with fluorescence -conjugated secondary

antibody for 1 hr. Immunofluorescence was visualized with confocal laser

scanning microscopy. All images were analyzed using a background

subtraction method offline.

Electrophysiological Recordings

Membrane currents were recorded from adult and neonatal rat myocytes

using whole-cell perforated patch and conventional whole-cell configurations,

respectively. In the siRNA transfected cells, only GFP-positive neonatal

myocytes were used for recording. Cells were perfused with the bath solution

58 containing (mM) NaCl 135, KCl 5.4, MgCl2 1.0, CaCl2 1.0, NaH2PO4 0.33,

HEPES 10, and glucose 10 at pH 7.4. The pipette solution contained (mM) KCl

140, MgCl2 1.0, HEPES 10, EGTA 5, and GTP 0.1 at pH 7.3. For conventional

whole-cell recordings, 0.25 mM ATP was added to pipette solutions for

recordings. For perforated whole-cell configuration, amphotericin B stock

solution (30 mg/ml in DMSO) was diluted with the pipette solution to reach the

final concentration (300 µg/ml) with the aid of sonication. All experiments were

conducted at room temperature (22°C to 25°C).

Borosilicate glass electrodes (outer diameter, 1.5 mm) were used with

resistances in range of 3 to 5 MΩ when filled and connected to a patch-clamp

amplifier (Axopatch 200B, Axon Instruments, Union City, CA, USA). For the

time course of KATP currents, the was held at -40 mV. KATP

currents were calculated as glibenclamide-sensitive currents.

Data analysis

Group data were presented as means ± SE. Multiple group means were

compared by ANOVA followed by LSD post hoc test. Differences with a two-

tailed P<0.05 were considered statistically significant.

3. Results

Localization of KATP channels in caveolin-enriched microdomains

To determine whether KATP channels are present in caveolin-rich

membrane fractions in cardiac myocytes, we used a detergent-free sucrose 59 gradient extraction-procedure for adult rat ventricular myocytes. Figure 2.1A

shows Western blot analysis of ten 1-ml fractions collected from top to bottom

of the sucrose density gradient. The marker proteins for caveolae, caveolin-1 and caveolin-3, were found predominantly in the lower-density fractions 4 and 5

(Fig. 2.1A, upper panel). In the same caveolar-enriched fractions, we detected

immunoreactivity for a significant amount of KATP channel pore-forming subunit

Kir6.2 (~ 80%). As reported previously (107), caveolin-3 was also detected in heavier samples (fractions 9-12) though enrichment for caveolin-3 occurred only in samples 4 and 5. Clathrin heavy-chain was detected with antibody against clathrin heavy chain across a broad range of the gradient fractions

similar to that reported by Sampson et al (245). Although there was some

overlap between caveolins and clathrin in fraction 5, it should be noted that the

distribution of Kir6.2 did not follow that of clathrin but only in caveolin-enriched

fraction. Coomassie blue staining of the gel showed that little protein was found

in the earlier fractions (fractions 4 and 5) that contain substantial amounts of

caveolin-3 and caveolin-1 but excludes most of the cellular protein (Fig. 2.1A, lower panel). Measurement of cholesterol levels within each fraction showed cholesterol to be enriched in fraction 4 and 5 (Fig. 2.1B), consistent with the observation that these two fractions represent caveolae-containing membrane fraction of the density gradient. Therefore, KATP channel pore-forming subunit

Kir6.2 is present in the caveolin-rich membrane fractions of cardiac ventricular

myocytes.

60

Figure 2.1. Kir6.2 is enriched in caveolar plasma membrane in adult rat

cardiomyocytes. A, 1 ml fractions were collected from the top of the gradient and analyzed with antibody against Kir6.2, caveolin-1 (Cav-1), caveolin-3 (Cav-

3), clathrin heavy-chain (Clathrin-HC). Lower panel shows coomassie blue staining pattern of same sucrose gradient fractions. B, Relative cholesterol levels of each fraction. Three independent experiments were performed.

61

Figure 2.1. Kir6.2 is enriched in caveolar plasma membrane in adult rat cardiomyocytes.

62 Co-immunoprecipitation of KATP channels and caveolin-3 from adult rat cardiac myocytes

To determine whether muscle-specific caveolin isoform caveolin-3 and

Kir6.2 are associated with each other, we performed immunoprecipitation experiments from the cell lysates of adult rat cardiac myocytes. The cell homogenate was incubated with rabbit anti-Kir6.2 antibody to precipitate caveolin-3. The immune complexes were collected with r-protein G beads and analyzed by immunoblotting against mouse anti-caveolin-3 and anti-Kir6.2 antibodies. Figure 2.2 shows that anti-Kir6.2 antibody immunoprecipitated caveolin-3. Conversely, anti-caveolin-3 antibody immunoprecipitated Kir6.2.

Neither protein was immunoprecipitated with control IgGs (mouse IgG and rabbit IgG). These results suggest that the Kir6.2 associates with caveolin-3 in adult rat ventricular myocytes.

63

Figure 2.2. KATP channels are associated with caveolin-3 (Cav-3) in adult rat ventricular myocytes. Cav-3 was detected in the immunoprecipitates with anti-Kir6.2. Similarly, Kir6.2 was detected in the Cav-3 immunoprecipitates.

Control IgGs (mouse and rabbit IgG) did not immunoprecipitate the proteins.

64

Figure 2.2. KATP channels are associated with caveolin-3 (Cav-3) in adult

rat ventricular myocytes.

65 Colocalization of KATP channel pore-forming subunits and caveolin-3

To investigate whether KATP channel pore-forming subunit Kir6.2

colocalizes with caveolin-3, we employed immunofluorescence confocal

microscopy. Figure 2.3 shows fluorescence images from adult rat ventricular

myocytes. The antibody against Kir6.2 (red) or caveolin-3 (green) demonstrated a prominent surface sarcolemmal punctate staining area. Merged images showed significant punctuate areas of colocalization for Kir6.2 and caveolin-3 along the plasma membrane. In agreement with previous observations (187),

Kir6.2 was clearly localized in peripheral sarcolemma and T tubules. Caveolae

vesicles have been noted before by electron microscopic examination of the myocardium on both peripheral plasma membrane and T tubules. Our immunostaining also showed the presence of caveolin-3 in peripheral sarcolemma and T tubules (82, 157). These data suggests that cardiac KATP

channels are largely localized in caveolin-3-associated membrane domains.

67

Figure 2.3. Colocalization of Kir6.2 and caveolin-3 (Cav-3). Double labeling confocal images of adult rat ventricular myocytes with anti-Kir6.2 and anti-Cav-

3 antibodies. Punctate areas of colocalization (represented by yellow) are apparent along the plasma membrane. The white boxes represent enlarged areas shown in the lower panels.

68 Effect of disrupting lipid rafts with methy-β-cyclodextrin (MβCD) on adenosine

A1 receptor-mediated activation of KATP channels

KATP channels are activated by adenosine via adenosine A1-receptors

and an inhibitory G-protein (Gi) in the heart (113, 170). This adenosine-

mediated KATP activation has been linked to cardioprotection against ischemia.

Both A1-receptors and Gi are localized in caveolar-enriched fractions. To

determine whether cholesterol-enriched microdomains are essential for

adenosine receptor-mediated activation of KATP channels, we tested the effect

of disrupting caveolae with a cholesterol depleting agent MβCD on KATP channel

activity in adult rat ventricular myocytes. Non-selective adenosine receptor

antagonist 8-phenytheophlline (8-PT) was used to ensure that adenosine effect

involves activation of adenosine receptors. Due to strong inhibition of KATP

channels by intracellular ATP, most of modulators including adenosine

receptors are shown to activate KATP channels only at reduced intracellular ATP level or in the presence of a . Since we employed the whole-cell perforated patch-clamp technique where physiological ATP level was preserved, we studied the effect of adenosine on KATP channels in the

presence of pinacidil (50 μM). The effect of MβCD on KATP channels evoked by

pinacidil alone (100 μM) was used as a negative control since pinacidil

activates KATP channel by interacting with SUR subunit and its effect should be

independent of lipid rafts/caveolae microdomains. As previously reported,

adenosine alone did not activate KATP channels due to stronger inhibition by

intracellular ATP (data not shown). Pinacidil alone at 50 μM elicited a small K+ 69 current that was completely blocked by selective KATP

glibenclamide (10 μM, Fig. 2.4A). However, when pinacidil (50 μM) and

adenosine (100 μM) were superfused together in the bath solution, they

activated a large K+ conductance channel that was completely inhibited by

glibenclamide (Fig. 2.4B). When cells were superfused with 10 mM MβCD for

10 min, which was previously shown to disrupt the function of caveolar associated proteins (17), prior to application of adenosine plus pinacidil, we found that MβCD pre-treatment almost completely eliminated the response of

KATP channels to adenosine receptors (Fig. 2.4C, 13.19±2.28 vs. 2.37±1.21

pA/pF, adenosine + pinacidil vs. adenosine + pinacidil + MβCD, p< 0.05, n=3-

4). In contrast, MβCD did not reduce KATP currents activated by 100 μM

pinacidil (15.37±3.41 vs. 19.67±4.15 pA/pF, pinacidil vs. pinacidil + MβCD,

p=NS), indicating that KATP channels evoked by pinacidil is independent of

cholesterol-rich plasma membrane microdomains. These observations indicate

that disrupting cholesterol-rich microdomains by MβCD significantly reduced

effect of adenosine receptors on KATP channels.

70

Figure 2.4. Caveolar disruption with MβCD eliminates adenosine

receptor-mediated stimulation of KATP channels. Time course of KATP

currents in perforated whole-cell configuration was recorded in the presence of

pinacidil 50 μM alone (Pina, A), pinacidil plus Adenosine 100 μM (Pina+Ade, B)

at a holding potential of -40 mV in control cells or cells pretreated with MβCD for 10 min (C). Glibenclamide (Glib) was added subsequently to ensure KATP

channel activation. D, Averaged current density in the presence of various agents as indicated, which was calculated by dividing glibenclamide-sensitive currents by cell capacitance. *p<0.05 vs. Pina. n=3-4.

71

Figure 2.4. Caveolar disruption with MβCD eliminates adenosine receptor-mediated stimulation of KATP channels.

72 Effect of disrupting caveolae with siRNA targeting caveolin-3 on adenosine A1

receptor-mediated activation of KATP channels

Although acute MβCD treatment of adult rat myocytes caused caveolar

disruption and resulted in the loss of regulation of KATP channels by adenosine

receptors, interpretation of results could be complicated by MβCD-induced

cholesterol depletion that may influence the regulation of KATP channels outside

of caveolae. To determine whether specifically destroying caveolae by siRNA knockdown of caveolin-3 expression alters the regulation of KATP channels by adenosine A1 receptors, we investigated the effect of caveolin-3 siRNA on KATP

channel regulation. We first examined whether a predesigned siRNA

oligonucletide for rat caveolin-3 can reduce endogenous caveolin-3 protein

level in neonatal rat cardiomyocytes. We found that caveolin-3 siRNA

significantly suppressed caveolin-3 but not caveolin-1 protein expression 48 hr

after transfection with siRNA. In contrast, control siRNA had no effect (Fig.

2.5B). Cells exhibiting GFP that was co-transfected with siRNA were used for

whole-cell voltage clamp studies. As shown in Fig. 2.5A, adenosine alone can

activate KATP channels after dialysis of pipette solution with a low ATP (0.25 mM). Transfection of neonatal cardiomyocytes with caveolin-3 siRNA largely

eliminated the stimulatory effect of adenosine A1 receptors on KATP channels

whereas control siRNA showed no significant changes (1.5±0.6 vs. 24.1±4.0

pA/pF, caveolin-3 siRNA vs. control siRNA, p< 0.01, n=4-5). In contrast, KATP

channels activated by pinacidil (100 μM) were not altered by siRNA targeting caveolin-3 (38.21±13.02 vs. 45.85±10.40 pA/pF, control vs. siRNA, p=NS). 73 Pinacidil followed by glibenclamide was applied to confirm there were functional

KATP channels in the cells transfected with caveolin-3 siRNA. Further, the effect of adenosine on KATP channels was almost completely blocked by a selective adenosine A1 receptor antagonist 8-cyclopentyl-1,3-dipropylxanthine (DPCPX,

1 μM, Fig. 2.5C). Taken together, these findings suggest that adenosine receptor-mediated activation of KATP channels in the cardiac myocytes requires intact caveolae where KATP channels and adenosine receptors associate.

74

Figure 2.5. Knockdown of caveolin-3 expression with siRNA prevents adenosine receptor-mediated activation of KATP channels. A, Conventional whole-cell voltage clamp recordings of KATP currents were performed at a holding potential of -40 mV in neonatal rat cardiomyocytes transfected with nothing (Control), control siRNA or siRNA targeting caveolin-3 (Cav-3 siRNA).

Adenosine (Ade), glibenclamide (Glib), or pinacidil (Pina, 100 μM) were applied to the bath solutions at various times as indicated by horizontal lines. B,

Western blot shows significant reduction of endogenous caveolin-3 (Cav-3) by siRNA against caveolin-3 (Cav-3 siRNA) and no change in caveolin-1 (Cav-1).

C, Averaged current density induced by adenosine (100 μM) was shown in cells transfected with nothing, control siRNA or Cav-3 siRNA. DPCPX was used to ensure that KATP channel activation by adenosine was through the stimulation of adenosine A1 receptors. **p<0.01 vs. control for DPCPX group, vs. control siRNA for Cav-3 siRNA group. n=4-5.

75

Figure 2.5. Knockdown of caveolin-3 expression with siRNA prevents adenosine receptor-mediated activation of KATP channels.

76 Co-immunoprecipitation of KATP channels and caveolin-3 from neonatal rat hearts

To determine that caveolin-3 and KATP channels reside in the same membrane compartment in neonatal rat hearts as those in adult rat hearts, we subjected the caveolin-rich fractions (fractions 4-6) to immunoprecipitation with anti-caveolin-3 or anti-Kir6.2 antibodies. The immunoprecipitates were analyzed by immunoblotting with antibodies against Kir6.2 or caveolin-3. As shown in Fig.

2.6, Kir6.2 co-precipitated caveolin-3 and caveolin-3 co-precipitated Kir6.2. The control samples without primary antibody did not immunoprecipitate the proteins. We also examined the presence of adenosine A1 receptors in the samples obtained by immunoprecipitation with anti-caveolin-3 antibodies.

Adenosine A1 receptors were detected in these samples. These data suggest that KATP channels and adenosine A1 receptors associate with caveolin-3-rich membrane compartments in the neonatal rat heart.

77

Figure 2.6. Immunoblot analysis of samples from neonatal rat hearts immunoprecipitated with anti-caveolin-3 (Cav-3) or anti-Kir6.2 antibodies.

Caveolin-rich fractions were immunoprecipitated with specified primary antibodies or no primary antibody (No ab). Kir6.2 and adenosine A1 receptors were detected in anti-caveolin-3 or anti-Kir6.2 immunoprecipitates, but not detected in the samples without primary antibody.

78

Figure 2.6. Immunoblot analysis of samples from neonatal rat hearts immunoprecipitated with anti-caveolin-3 (Cav-3) or anti-Kir6.2 antibodies.

79 Association of adenosine A1 receptors and KATP channels in caveolin-3-rich fractions of cardiomyocytes

We have shown that Kir6.2 associates with caveolin-3 in the total cell lysates of adult rat cardiomyocytes. In order to determine whether Kir6.2 and adenosine A1 receptors reside in the same caveolin-enrich microdomains and further elucidate the mechanism involved in adenosine-mediated KATP activation, we performed co-immunoprecipitation experiments in the caveolin-3- enriched fractions of adult rat ventricular myocytes. Cells were pretreated with

MβCD (10 mM) for 30 min prior to sucrose gradient fractionation. Caveolin-rich fractions (fractions 4-6) were immunoprecipitated with antibodies against Kir6.2, adenosine A1 receptors and caveolin-3. As shown in Fig. 2.7, Kir6.2, adenosine

A1 receptors and caveolin-3 coimmunoprecipitated with anti-Kir6.2, anti- adenosine A1 receptors or anti-caveolin-3 immunoprecipitates. The 50 kDa IgG heavy chain band was detected in the blot when probing with the antibody against adenosine A1 receptors. There was no detection of Kir6.2, adenosine A1 receptors and caveolin-3 in the homogenates without the treatment of primary antibodies. MβCD diminished caveolin-3 that co-precipitated with Kir6.2 and adenosine A1 receptors, while the levels of Kir6.2 and adenosine A1 receptors were also reduced. These data indicate that KATP channels and adenosine A1 receptors may associate with each other in caveolin-rich membrane microdomains and this association depends on the level of cholesterol.

80

Figure 2.7. Effect of MβCD on Kir6.2, adenosine A1 receptors (Ade A1) and caveolin-3 (Cav-3) coprecipitation. Adult rat cardiomyocytes were pretreated with or without MβCD (10 mM) for 30 min prior to homogenation and sucrose gradient centrifugation. The light fractions (fractions 4-6) were immunoprecipitated with antibodies against Kir6.2, Ade A1 or Cav-3, or without

primary antibody (No ab). Immunoblots were probed for the protein indicated.

81 4. Discussion

In the present study, we demonstrate that localization of KATP channels to

caveolin-enriched microdomains in the plasma membrane of cardiac myocytes

is essential for adenosine receptor-mediated activation of KATP channels.

Specifically, we show that KATP channel pore-forming subunit Kir6.2 is localized

in caveolin-enriched membrane fractions, and is associated and colocalized with caveolin-3. Patch-clamp studies reveal that this microdomain association is necessary for KATP channel regulation by activation of adenosine receptors.

KATP channels are regulated by a number of endogenous activators,

most of which are released or increased during ischemia such as noradrenaline

(116, 251), adenosine (114, 274), protein kinase C (112, 163), nitric oxide (36)

and membrane lipids especially phosphatidylinositol-4,5-bisphosphate (263). All

of these regulators activate KATP channels by decreasing ATP-sensitivity and

thus increasing open probability. It has been proposed that activation of KATP

channels by these endogenous regulators may be one of reasons why KATP

channels can be activated at millimolar range of intracellular ATP during

ischemia while KATP channels are inhibited by micromolar range of ATP in

patch-clamp studies (192). It is likely that those molecules decrease the ATP- sensitivity of KATP channels within caveolae and thus activates the channels

even though intracellular ATP level is relatively high. A recent study shows that

KATP channel subunit Kir6.2 is mandatory for optimal adaptation capacity under

stress, further indicating that cardiac KATP channels may not be as silent as

originally thought and the mechanism may involve regulation of the channel 82 microenvironment (320). The present study revealed a novel mechanism by

which KATP channels in the plasma membrane of cardiac myocytes are

regulated by caveolin-rich microdomains. This finding indicates that the

regulation of KATP channels is more complex than that can be explained by the simple hypothesis of opening by bulk decrease in intracellular ATP.

In the heart, several ion channels and exchangers are localized in

caveolae and have shown functional modulation by caveolae-associated mechanism including the voltage-dependent Na+ channel, voltage-dependent

Ca2+ channel, a voltage-dependent K+ channel (Kv1.5), the Na+/Ca2+

exchangers and Na+/K+-ATPase (17, 18, 30, 167, 178, 313). In vascular smooth

muscles KATP channels (Kir6.1 and SUR2B) are also present in caveolae (244,

245). It is not known that whether cardiac KATP channels, which are composed

of Kir6.2 and SUR2A, are localized in caveolin-rich microdomains. While the

adenosine A1 receptor is well known to couple to cardiac KATP channels via

inhibitory G proteins, and inhibitory G proteins often couple to their effectors by

protein kinase C (114, 141, 143), it is also not clear whether cholesterol-rich microdomains on the plasma membrane of cardiac myocytes impacts the

regulation of KATP channels by adenosine A1-receptor signaling. The present

studies provide the first evidence that intact caveolin-rich microdomains are

essential for adenosine A1-receptor-mediated KATP channel signaling. This is

consistent with the previous finding that adenosine A1-receptors and their

downstream signaling molecules including Gi protein and protein kinase C are

localized in caveolae (17, 153, 206, 305). 83 In the present study, disruption of caveolae may affect other modulators

of KATP channels within caveolae. If adenosine-induced activation of KATP

channels is obscured by disrupting other caveolar-associated signaling

molecules that are active under basal conditions, pinacidil-evoked KATP currents

should also be reduced. Our data showed that disrupting caveolae by either

MβCD or caveolin-3 siRNA did not alter the amplitude of pinacidil-evoked KATP

currents, indicating that basal KATP activity may not be significantly affected by

the net effect of disrupting caveolae.

It is known that caveolin-3 is essential for the formation of caveolae in

cardiac myocytes, whereas caveolin-1 is expressed and important in cardiac

myocytes. We noticed that we detected significant amount of caveolin-1 in

addition to caveolin-3 in our cardiomyocyte preparations. This could raise a

question whether caveolin-1 is involved in adenosine-mediated regulation of

KATP channels. Our data from neonatal rat cardiomyocytes show that suppression of caveolin-3 by specific siRNA did not alter the expression of caveolin-1 but significantly affected the regulation of KATP channels by

adenosine receptors. This observation indicates that caveolin-3 not caveolin-1

play an important role in KATP channel regulation by adenosine receptors in

cardiomyocytes.

We understand that there are significant developmental changes in ion

channel expression and/or distribution. The subsarcolemmal localization of KATP

channels in adult rat heart may not be the same as that in neonatal rat heart.

However, our immunoprecipitation experiments from caveolin-3-enriched 84 fractions of neonatal rat heart demonstrate that Kir6.2 indeed associates with

caveolin-3, which is consistent with the observation that KATP channels are

functionally regulated by intact caveolae or caveolin-3.

We have previously shown that acute stimulation (<5 min) of adenosine

receptors activates KATP channels (114, 141) and prolonged stimulation (15-30

min) causes internalization of KATP channels in cardiac myocytes (113). In the

present study, the KATP currents started to increase significantly around 5 min

after superfusion with adenosine, and reached the steady state within 5-10 min.

The time course for adenosine to induce a detectable increase in KATP currents

and then reach a steady state depends on the rate of superfusion, the level of

intracellular ATP, metabolic state of the cell and the concentration of pinacidil.

Although adenosine A1 receptor has been shown to translocate out of caveolae

after agonist stimulation, this translocation occurs after prolonged stimulation

which lasts for 15 min (153). It is thus likely that the acute effect of adenosine

A1 receptors on KATP channels in cardiac myocytes is caveolae-dependent.

In recent years, it has become apparent that the regulation of ion

channel function is not the only means of controlling cell excitability. The

trafficking and localization of ion channels with signaling molecules also play an

important role. The precise subcellular localization of ion channels is often

necessary to ensure efficient integration of both intracellular and extracellular events. The present study has identified a novel mechanism for regulation of cardiac KATP channel. Giving that caveolae are emerging as important surface- associated organelles, our elucidation of caveolar microdomain-dependent 85 regulation of KATP channels in cardiac myocytes will significantly enhance our knowledge of the complexity of regulation of this macromolecular channel and provide mechanistic insight into how KATP channels control cellular functions.

Acknowldgement

This work was supported by the American Heart Association National

Center [0835117N to K.H.].

86

CHAPTER 3

CAVEOLIN-3 NEGATIVELY REGULATES RECOMBINANT CARDIAC ATP-

SENSITIVE POTASSIUM CHANNELS IN HEK293T CELLS

ABSTRACT

We have recently shown that ATP-sensitive potassium (KATP) channels in the heart are localized in the caveolae of cardiac myocytes and regulated by caveolae-related signaling. However, little is known about the role of caveolins, signature proteins of caveolae, in cardiac KATP channel function. The present study was designed to explore the potential functional interaction between caveolin-3 and KATP channels. The cardiac KATP channel subunits Kir6.2 and

SUR2A were transiently transfected in HEK293T cells with or without co- transfection of caveolin-3. Our data demonstrated that the recombinant KATP channel activity in HEK293T cells was inhibited by expression of caveolin-3 but not caveolin-1. Application of caveolin-3 scaffolding domain peptide, corresponding to amino acid residues 55-74 of caveolin-3, blocked the inhibitory effect of caveolin-3 on KATP channels. However, the same peptide did not have any significant effect on KATP channels in HEK293T cells without 87 caveolin-3 expression. We further confirmed that expressed caveolin-3 co- localized and co-immunoprecipitated with the KATP channel pore-forming subunit Kir6.2. Thus, our results indicate that caveolin-3 negatively regulates

Kir6.2/SUR2A channel function.

88 1. INTRODUCTION

A variety of proteins are located within caveolae. Caveolae bring

signaling molecules together into a confined microdomain, whereas caveolins,

structural proteins of caveolae, interact with a range of key proteins localized

within caveolae. Many signaling molecules are shown to interact with caveolins

such as G proteins, G protein–coupled receptors, Src family protein kinases,

Ha-Ras, and nitric oxide synthases (51). Caveolin is a scaffolding protein that

can bind and assemble multiple signaling molecules in a large complex and are

known to modulate the activity of most of its interacting partners. Direct

interaction between caveolins and interacting proteins may serve to either

anchor proteins to caveolae or modulate protein function.

Regulation of ion channel function is not the only means of controlling

cell excitability. The localization and trafficking of ion channels play an

important role in cellular function. The precise subcellular localization of ion

channels is important for rapid and efficient integration of cellular signaling events. Lipid raft association has been identified as a novel mechanism for the subcellular sorting of specific ion channels to the plasma membrane microdomains rich in signaling molecules (175, 210).

The ATP-sensitive potassium channel (KATP) is a highly abundant

plasma membrane protein responsible for linking the cellular energy levels to

membrane potential and cellular excitability (2, 197). KATP channels are

heterooctamers consisting of two structurally unrelated proteins: four pore-lining

subunits that belong to the Kir6 subfamily of inwardly rectifying potassium 89 channels, and four regulatory subunits of the ATP-binding cassette superfamily.

Cardiac KATP channels have been shown to protect against the metabolic insult of ischemia and contribute to adaptive responses to metabolic stress (139,

212). We have recently shown that KATP channels in the heart are primarily localized in the caveolae of cardiac myocytes and regulated by caveolae-

related adenosine receptor signaling (85). Giving that caveolin has been shown

to regulate the function of certain molecules accumulated in caveolae; we

sought to examine the potential functional role of caveolins on cardiac KATP

channels. In the present study, HEK293T cells were transfected with cardiac

type KATP channel subunits Kir6.2/SUR2A with or without caveolins. The effect

of caveolin-3 or caveolin-1 on recombinant KATP channel function was

determined by expression of caveolin-3, caveolin-1 or use of caveolin-3

scaffolding domain peptide (CSD). We demonstrate that caveolin-3 not

caveolin-1 negatively regulates recombinant cardiac KATP channel function.

2. Materials and methods

Chemicals and reagents

Caveolin-3 scaffolding domain peptide (CSD) containing the putative

scaffolding domain of caveolin-3 (amino acids 55–74,

DGVWRVSYTTFTVSKYWCYR) was purchased from AnaSpec Inc. (San Jose,

CA, USA). Pincidil and glibenclamide were purchased from Sigma-Aldrich (St.

Louis, MO, USA).

90

Cell culture and transfections

Human embryonic kidney (HEK)293T were cultured in DMEM,

supplemented with 10% fetal bovine serum, L-glutamine (2 mM), penicillin (100

IU/ml), and streptomycin (100 µg/ml). HEK293T cells were plated onto 35-mm culture dishes and transfected with cardiac KATP channel subunits, HA-tagged

Kir6.2 and SUR2A, with or without cavoelin-3-ECFP (180) or caveolin-1-EGFP

(204) by using FuGENE6 (Roche). Cells were also transfected with GFP to

ensure there is transfection. All mammalian expression constructs are either in

pcDNA3, pCMV, pECFP or pEGFP. Experiments were conducted 2-3 days after

transfection.

Electrophysiology Studies

The whole-cell KATP currents in HEK293T cells were recorded using

patch-clamp technique (112, 128). Only the cells exhibiting GFP were used for

recordings. Cells were superfused with the bath solution containing (mM) NaCl

135, KCl 5.4, MgCl2 1.0, CaCl2 1.0, NaH2PO4 0.33, HEPES 10, and glucose 10 at pH 7.4 (pH adjusted with NaOH). The pipette solution contained (mM) KCl

140, MgCl2 1.0, HEPES 10, EGTA 5, GTP 0.1, ADP 0.1 at pH 7.3 (pH adjusted

with KOH). 3 mM ATP was added to the pipette solution for spontaneous KATP

currents and 5 mM ATP was for pinacidil-evoked KATP currents. Data were

sampled at 13.3 kHz with an A/D converter (Digidata 1322A, Axon Instruments)

and stored on the hard disk of a computer for subsequent analysis. The 91 recordings were filtered with a low-pass corner frequency of 2 kHz. Borosilicate glass electrodes (outer diameter, 1.5 mm) were used with resistances in range of 3 to 5 MΩ when filled and connected to a patch-clamp amplifier (Axopatch

200B, Axon Instruments). The time course of KATP currents were recorded at a holding potential of 0 mV. The activation of KATP currents was verified by application of KATP channel selective blocker glibenclamide. All the Recordings were carried out at room temperature (22-25oC).

Co-immunoprecipitation and Western blotting

HEK293T cells were rinsed with ice-cold PBS (pH 7.4), homogenized in ice-cold lysis buffer (25 mM Tris-HCl; 250 mM NaCl; 10 mM EDTA; pH 7.6) containing 1% Triton X-100 and 2 mM phenylmethylsulfonyl fluoride with protease inhibitor tablets (Roche) (85). Insoluble material was pelleted by centrifugation at 10,000Xg for 15 min and the lysates were precleared by incubation with r-protein-G agarose (Invitrogen; 1 hour, 4°C). Approx. 800 µL of cleared lysate (1µg/µl) was incubated with anti-HA (1:100), anti-caveolin-3 antibody (1:100) for 2 hours at 4°C. Control sample was incubated without antibody. Antigen-antibody complexes were captured with r-protein-G agarose

(4°C, 2 hours). Agarose beads were washed 4-times with lysis buffer before removal of bound proteins by boiling in SDS sample buffer. Samples were resolved by SDS-PAGE (10–12% acrylamide gel) and transferred onto nitrocellulose membrane. The transferred blots were blocked with 5% nonfat- milk in Tris-buffered saline (TBST: 150 mM NaCl, 20 mM Tris-HCl, 0.05% 92 tween-20, pH 7.4) and incubated for 1 hr at room temperature with antibodies

against HA-tag or caveolin-3 in TBST. After washing, the blots were incubated

with peroxidase-conjugated secondary antibodies for 45 min and developed

using the ECL detection system (Amersham).

Immunofluorescence microscopy

Immunolabeling was performed in HEK293T cells transfected with

Kir6.2-HA/SUR2A/Cav-3-CFP (85, 113). Briefly, cells were fixed with 4% paraformaldehyde in PBS for 30 min, blocked, permeablized in 5% goat serum in PBS with 0.1% Triton X-100 (30 min), and labeled with a rat anti-HA antibody

for 2 hours (Roche, Indianapolis, IN, USA ). Cells were washed three times and

labeled with fluorescent secondary antibody for 45 minutes.

Immunofluorescence was visualized with a Nikon epifluorescence. Cells were

randomly selected for imaging and analysis.

Statistics

Group data were presented as means ± SE. Unpaired t-test was used to

compare between groups. Multiple group means were compared by ANOVA

followed by LSD post hoc test. Differences with a two-tailed P<0.05 were

considered statistically significant.

93 3. RESULTS

Effects of caveolin-3 on spontaneous Kir6.2/SUR2A channels

To determine whether caveolin-3 directly modulate KATP channel

function, we recorded the time course of spontaneous whole-cell KATP currents

at a holding potential of 0 mV in HEK293T cells transfected with Kir6.2/SUR2A or Kir6.2/SUR2A/caveolin-3 using a conventional whole-cell patch-clamp technique. Only cells show green fluorescence were used for recordings. The whole-cell spontaneous KATP currents were developed relative slowly when cells

were dialyzed with the pipette solution containing 3 mM ATP and 100 μM ADP.

We found that all 8 cells without transfection of caveolin-3 showed spontaneous

KATP currents whereas 6 out of 8 cells co-transfected with caveolin-3 did not

generate significant spontaneous KATP currents (Fig. 3.1A and 3.1B). As shown

in the figure 3.1, spontaneous KATP currents were blocked by application of the

KATP channel inhibitor glibenclamide (10 μM). The cells that did not show

significant spontaneous KATP currents were further tested by subsequent

application of the KATP channel opener pinacidil (100 μM) to ensure sufficient

expression of KATP channels (data not shown). Figure 3.1C summarizes the

current density from cells transfected with Kir6.2/SUR2A or

Kir6.2/SUR2A/caveolin-3 (66.79 ± 18.66 pA/pF vs. 8.58 ± 5.88 pA/pF, control

vs. caveolin-3, p<0.05, n = 8).

94

Figure 3.1. Modulation of spontaneous KATP currents by caveolin-3. (A)

The spontaneous Whole-cell KATP current was detected at 0 mV in HEK293T cells transfected with Kir6.2/SUR2A and blocked by 10 μM glibenclamide (Glib).

(B) The spontaneous Whole-cell KATP currents were detected at 0 mV in

HEK293T cells transfected with Kir6.2/SUR2A/caveolin-3. (C) Mean current

density (pA/pF) in HEK293T cells transfected with Kir6.2/SUR2A (Control) or

Kir6.2/SUR2A/caveolin-3 (Cav-3). Drugs were applied as indicated by the

horizontal lines. The gaps between the traces were interruptions for recording

step currents with various voltages. p<0.05, vs. control. n=8 cells.

95

Figure 3.1. Modulation of spontaneous KATP currents by caveolin-3.

96 Effect of caveolin-3 on pinacidil-induced Kir6.2/SUR2A channels

Since expression of caveolin-3 suppressed spontaneous KATP currents,

we decided to study further the effect of caveolin-3 on KATP currents evoked by

pinacidil. With pipette solution containing 5 mM ATP and 100 μM ADP,

spontaneous KATP currents in HEK293T cells transfected with Kir6.2/SUR2A

were usually not detected even after dialysis for 20-30 minutes. The whole-cell

KATP currents evoked by pinacidil were recorded 10 minutes after formation of whole-cell configuration. As shown in Figure 3.2, an outward current at a holding potential of 0 mV was activated in the presence of 50 μM pinacidil and completely blocked by 10 μM glibenclamide. Expression of caveolin-3 in

HEK293T cells significantly reduced pinacidl-induced currents (134.71 ± 30.19 pA/pF vs. 53.78 ± 15.15 pA/pF, control vs. caveolin-3, p<0.05, n=6-7).

However, when caveolin-1 was transfected into HEK293T cells, the amplitude of pinacidil-induced KATP currents was not different from that in control cells

(134.71 ± 30.19 pA/pF vs. 115.84 ± 36.85 pA/pF, control vs. caveolin-1, p=NS,

n = 6-7). Expression of caveolin-1 and caveolin-3 were verified by their tagged

fluorescent proteins. These observations indicate that the function of

recombinant cardiac KATP channels is specifically regulated by caveolin-3 not

caveolin-1.

97

Figure 3.2. Modulation of pinacidil-activated KATP currents by caveolin-3.

The time course of Whole-cell KATP currents were recorded at 0 mV in the presence of pinacidil (50 μM) or pinacidil plus glibenclamide (10 μM) in

HEK293T cells transfected with Kir6.2/SUR2A (A, Control),

Kir6.2/SUR2A/caveolin-3 (B, Cav-3) or Kir6.2/SUR2A/caveolin-1 (C, Cav-1). (D)

Mean current density (pA/pF) in HEK293T cells transfected with Kir6.2/SUR2A

(Control), Kir6.2/SUR2A/caveolin-3 (Cav-3) or Kir6.2/SUR2A/caveolin-1 (Cav-

1). The pipette solution contained 3 mM ATP and 100 μM ADP. Drugs were

applied as indicated by the horizontal lines. The gaps between the traces were

interruptions for recording step currents with various voltages. P< 0.05, vs.

control. n=6-7 cells.

98

Figure 3.2. Modulation of pinacidil-activated KATP currents by caveolin-3.

99 Effect of caveolin-3 scaffolding domain peptide on pinacidil-induced

Kir6.2/SUR2A channels

Next, we sought to examine whether functional disruption of caveolin-3 affects KATP channel activity in HEK293T cells transfected with

Kir6.2/SUR2A/caveolin-3. Caveolin-3 scaffolding domain peptide (CSD, 10 μM)

corresponding to putative scaffolding domain of caveolin-3 (amino acids 55–74,

DGVWRVSYTTFTVSKYWCYR) was introduced into the cell interior from the

patch pipette. Pinacidil (50 μM) was applied to induce KATP currents after 10-

min dialysis. This caveolin-3 CSD has been previously shown to block caveolin-

3 function (138). Figure 3.3 shows one representative experiment. The

amplitude of KATP current evoked by pinacidil in the cells transfected with

Kir6.2/SUR2A/caveolin-3 was significantly larger in the presence of CSD than

that in the absence of CSD.

Caveolin-3 CSD has been shown to directly modulate the function of

some proteins. To determine whether there is a direct functional interaction

between caveolin-3 CSD and KATP channels, we take advantage of HEK293T

cells that lack endogenous caveolin-3 to record pinacidil-induced KATP currents

in HEK293T cells without caveolin-3 transfection. Figure 3.3C summarizes the current density for control, caveolin-3, control plus caveolin-3 CSD and

caveolin-3 plus caveolin-3 CSD. While the expression of caveolin-3 inhibited

KATP channel currents, Caveolin-3 CSD significantly blocked the effects of caveolin-3 on KATP channels (53.78 ± 15.15 pA/pF vs. 131.53 ± 26.13 pA/pF,

caveolin-3 vs. caveolin-3 + caveolin-3 CSD, p<0.05, n=6-7). However, caveolin- 100 3 SDP did not have the direct functional effect on KATP channels (109.16 ±

12.05 pA/pF vs. 134.71 ± 30.19 pA/pF, control + caveolin-3 SDP vs. control, p=NS, n=5-7). The above data suggest that caveolin-3 negatively regulates the function of recombinant cardiac KATP channels.

101

Figure 3.3. Effect of caveolin-3 scaffolding domain peptide (CSD) on KATP

currents. The time course of pinacidil-activated whole-cell KATP currents in the

presence of caveolin-3 CSD (Cav-3 CSD, 10 μM ) was recorded at 0 mV in

HEK293T cells transfected with Kir6.2/SUR2A (A) or Kir6.2/SUR2A/caveolin-3

(B). Cav-3 CSD was applied in the pipette solution. The KATP currents evoked

by 50 μM pinacidl were verified by 10 μM glibenclamide (Glib). (C) Mean current density (pA/pF) in HEK293T cells transfected with or without caveolin-3

in the presence or absence caveolin-3 CSD. The pipette solution contained 5

mM ATP and 100 μM ADP. Drugs were applied as indicated by the horizontal

lines. The gaps between the traces were interruptions for recording step

currents with various voltages. p<0.05, vs. Cav-3 + Cav-3 CSD. n=5-7 cells.

102

Figure 3.3. Effect of caveolin-3 scaffolding domain peptide (CSD) on KATP

currents.

103 Colocalization and co-immunoprecipitation of KATP channel pore-forming

subunit Kir6.2 and caveolin-3

To determine whether caveolin-3 physically associates with KATP

subunits, co-immunoprecipitation experiments were performed from the cell

lysates of HEK293T cells transfected with Kir6.2/SUR2A and caveolin-3. The

cell homogenate was incubated with anti-HA antibody or caveolin-3 to

precipitate caveolin-3 or Kir6.2, respectively. The immune complex were

collected with protein G beads and analyzed by immunoblotting against anti-HA

and anti-caveolin-3 antibodies. Fig. 3.4A shows that anti-HA antibody

immunoprecipitated caveolin-3 and anti-caveolin-3 antibody immunoprecipitated

Kir6.2. Neither protein was immunoprecipitated in the samples without antibody

treatment.

To investigate whether KATP channels colocalizes with caveolin-3, we

employed immunofluorescence confocal microscopy. Fig. 3.4B shows the representative fluorescence images of cells transfected with

Kir6.2/SUR2A/caveolin-3. The plasma membrane of cells was strongly labeled with both Kir6.2 and caveolin-3. Substantial colocalization of Kir6.2 (red) with caveolin-3-ECFP (green) was observed in these cells, as indicated by yellow functate staining alone the plasma membrane in the merged image. These data suggests that recombinant cardiac KATP channels specifically associates with

expressed caveolin-3.

104

Figure 3.4. Co-immunoprecipitation and colocalization of Kir6.2/SUR2A channels with caveolin-3. HEK293T cells were transfected with Kir6.2-

HA/SUR2A/caveolin-3-ECFP (Cav-3-ECFP). (A) Immunoblots from immunoprecipitates with anti-caveolin-3 (Cav-3) or anti-HA antibodies (anti-HA).

Control sample was incubated without antibody (No ab). (B) Representative fluorescent images from transfected cells labeled with anti-HA antibody for

Kir6.2 (red) and ECFP for Cav-3-ECFP (green). The merged images showed significant colocalization between Cav-3-ECFP and Kir6.2.

105 4. Discussion

In the present study, we provide the first evidence that cardiac type KATP

channels expressed in HEK293T cells were negatively regulated by caveolin-3.

Specifically, we show that KATP channel activity was significantly inhibited by expression of caveolin-3 but not caveolin-1. Further, we demonstrate that

caveolin-3 CSD blocked caveolin-3-mediated suppression of KATP channels but

did not exhibit any direct functional effect on KATP channels. The physical

association of expressed KATP channels with caveolin-3 in HEK293T cells was confirmed by immunofluorecent co-localization and co-immunoprecipitation.

Caveolin is a major protein component of caveolae, flask-shaped plasma membrane invagination. Caveolae has been shown to participate in transmembrane signaling, such as G protein-coupled signaling (232). In addition, caveolin may directly interact with molecules that are accumulated in caveolae and regulate their function (202). KATP channels are regulated by a

number of endogenous activators (36, 112, 114, 116, 163, 251, 263, 274).

Interestingly, most of these modulators and receptors reside primarily in

caveolae (152, 153, 242, 305), suggesting a possible role for KATP channels in

caveolin-dependent cellular function. We have recently shown that KATP

channels are mostly localized in the caveolae of cardiac myocytes and

adenosine-induced KATP channel function is regulated by intact caveolae (85).

In the present study, we extend this finding by elucidating the functional role of

caveolin-3 on KATP channels. We expressed caveolin-3 or caveolin-1 in

HEK293T cells transfected with Kir6.2/SUR2A and studied the effect of 106 expressed caveolin on the basal KATP channel activity. Our data indicate that

caveolin-3 not caveolin-1 is a major regulator of cardiac KATP channel function.

It appears that caveolae is essential for regulation of cardiac KATP

channels by adenosine receptors while caveolin-3 negatively modulates KATP

channel function. Our previous data show that in cardiac myocytes disruption of

caveolae by suppressing caveolin-3 with siRNA affects adenosine-mediated

activation of KATP channels but does not alter the basal KATP channel function

(pinacidil-evoked KATP) (85). The observation that knockdown of caveolin-3 expression by siRNA did not significantly alter pinacidil-induced KATP currents

seems different from what we observed in the present study where caveolin-3

inhibits pinacidil-induced recombinant KATP channel activity. This discrepancy

could be attributed to the incomplete knockdown of endogenous caveolin-3 by

siRNA in native cardiac myocytes. This insufficient suppression of caveolin-3 by

siRNA appeared to be efficient to destroy caveolae structure and thus affect

adenosine-mediated KATP channel regulation (85) but did not provide enough

evidence for the role of caveolin-3 on KATP channel function.

Caveolin-3 CSD encode membrane proximal regions of caveolin-3 and

are known to functionally interact with some caveolin-associated proteins (138,

290). To assess the functional role of caveolin-3 on KATP channels, we tested

the effect of caveolin-3 CSD on pinacidil-evoked KATP channels in HEK293T

cells. Our data show that the caveolin-3 CSD blocked the inhibitory effect of

expressed caveolin-3 on KATP channels, indicating that the caveolin-3 has an

important functional role on KATP channel regulation. It is likely that the caveolin- 107 3 scaffolding domain peptide breaks the binding of caveolin-3 to other protein(s)

that regulates KATP channels. This notion was further confirmed by our observation that caveolin SDP did not have significant functional effect on KATP

channels in HEK293T cells without expression of caveolin-3. Although the exact

mechanism by which caveolin-3 suppresses KATP channel function remains

elusive, it is likely that caveolin-3 scaffolding domain does not exert a direct

functional effect on cardiac KATP channels expressed in HEK293T cells but

interacts with other binding partners to regulate KATP channel function indirectly

(138). The advantage of using HEK293T cells when compared with other cells

is that they lack endogenous caveolin-3, contain minimum amount of caveolin-1

and do not express detectable KATP channels. Thus we can dissect not only the

specific role of individual caveolin isoforms but also the functional effect of

putative scaffolding domain peptide on KATP channels without interference of

endogenous proteins.

While activation of KATP channels during metabolic stress is protective

against injury, excessive KATP channel opening may have deleterious

consequences. In the heart, active KATP channels cause the action potential

duration to shorten and contribute to cellular K+ loss, both are risk factors for

arrhythmia (126, 299). KATP channel blockers have been shown to prevent

ventricular arrhythmia in the canine model of sudden cardiac death (25). A tight

regulation of Caveolin-3 expression may play an important role in controlling

KATP channel function under metabolic stress to confer protection without

incurring serious damage. 108

Acknowledgments

We would like to thank Dr. Jeffery R. Martens (University of ,

MI) for kindly providing caveolin-3-ECFP and Dr. Mark A. McNiven (Mayo

Clinic, MN) for caveolin-1-EGFP DNA clones. This work was supported by the

American Heart Association National Center [0835117N to K.H.].

109

CHAPTER 4

PROTEIN KINASE C ISOFORM-DEPENDENT MODULATION OF ATP-

SENSITIVE POTASSIUM CHANNELS IN MITOCHONDRIAL INNER

MEMBRANE

ABSTRACT

+ The ATP-sensitive K (KATP) channels in both sarcolemmal (sarcKATP) and mitochondrial inner membrane (mitoKATP) are the critical mediators in

cellular protection of ischemic preconditioning (IPC). Whereas cardiac sarcKATP

contains Kir6.2 and sulfonylurea receptor SUR2A, the molecular identity of

mitoKATP remains elusive. In the present study, we tested the hypothesis that protein kinase C (PKC) may promote import of Kir6.2-containing KATP into

mitochondria. Fluorescence imaging of isolated mitochondria from both rat adult

cardiomyocytes and COS-7 cells expressing recombinant Kir6.2/SUR2A

showed that Kir6.2-containing KATP channels were localized in mitochondria and

this mitochondrial localization was significantly increased by PKC activation

with phorbol 12-myristate 13-acetate (PMA). Fluorescence resonance energy

transfer microscopy further revealed that a significant number of Kir6.2- 110 containing KATP channels were localized in mitochondrial inner membrane after

PKC activation. These results were supported by Western blotting showing that

the Kir6.2 protein level in mitochondria from COS-7 cells transfected with

Kir6.2/SUR2A was enhanced after PMA treatment and this increase was

inhibited by the selective PKC inhibitor chelerythrine. Furthermore, functional

analysis indicated that the number of functional KATP channels in mitochondria

was significantly increased by PMA, as shown by KATP-dependent decrease in

mitochondrial membrane potential in COS-7 cells transfected with

Kir6.2/SUR2A but not empty vector. Importantly, PKC-mediated increase in

mitochondrial Kir6.2-containing KATP channels was blocked by a selective PKCε

inhibitor peptide in both COS-7 cells and cardiomyocytes. We conclude that the

KATP channel pore-forming subunit Kir6.2 is indeed localized in mitochondria

and that the Kir6.2 content in mitochondria is increased by activation of PKCε.

PKC isoform-regulated mitochondrial import of KATP channels may have significant implication in cardioprotection of IPC.

111 1. INTRODUCTION

Activation of ATP-sensitive potassium (KATP) channels could protect the

heart under metabolic stress (96, 189, 193, 195). Importantly, the modulation of

KATP channels accounts for the ability of brief ischemia and reperfusion to

protect the heart against infarction induced by subsequent prolonged ischemia,

a phenomenon known as ischemic preconditioning (IPC) (95, 189, 199). Not

only do molecularly defined cell surface sarcKATP channels in the heart activate

on metabolic stress and cause action potential shortening and less energy

consumption (113, 195, 275), KATP channels in the mitochondrial inner

membrane (mitoKATP channels) are also implicated in cardioprotection during ischemia (86, 95, 199). However, despite immense interest and serious

attempts to clone mitochondrial KATP channels, their molecular nature remains

elusive. The KATP channels in the sarcolemma membrane, first described in the

heart by Noma (197) in 1983, are widely distributed in many tissues and cell types including pancreatic β-cell, neuron, skeletal muscle, and smooth muscle

(11, 13). The major function of the KATP channel is to couple the cell metabolic

state to its membrane potential by sensing changes in intracellular adenine

nucleotide concentration (6). The sarcKATP channels are heterooctamers

consisting of two structurally unrelated proteins (2, 3, 20, 254, 257): four pore-

lining subunits that belong to the Kir6 subfamily of inwardly rectifying potassium

channels and four regulatory subunit sulfonylurea receptors (SUR) of the ATP-

binding cassette (ABC) superfamily. It is currently known that the cardiac

112 sarcKATP channel is composed of an octomeric complex of two types of

subunits, the Kir6.2 and the SUR2A subunit.

Whereas the mitoKATP channel has been characterized

pharmacologically (50, 87), its molecular nature remains unknown. Whether

Kir6.2 contributes to mitoKATP channels is under debate; the subunit has been

found in heart mitochondria (54, 147, 265, 319). We believe that much of the

controversy or difficulty regarding the existence of Kir6.2-containing KATP

channels in mitochondria may be due to the low level of basal KATP channels

present in mitochondria. All these studies concerning the existence of Kir6.2-

containing KATP in mitochondria were conducted under basal conditions that do not involve IPC or its signaling molecule(s), a critical step that may upregulate

protein transport to mitochondria and thus confer cardioprotection (26).

Although KATP channel regulation and implication in cardioprotection

have been studied by many groups for years, no study has ever explored the

trafficking aspect of KATP function in mitochondria. The present study utilizing

KATP-deficient COS-7 cells represents a novel approach to elucidation of the molecular basis of the KATP channels in mitochondria and their regulatory

mechanism. We report the novel finding that a protein kinase C (PKC) isozyme,

PKCε, promotes mitochondrial import of Kir6.2-containing KATP channels from

cytosol, assessed with a combination of immunofluorescence microscopy,

Western blot, fluorescence resonance energy transfer (FRET) analysis, and

mitochondrial functional studies.

113 2. MATERIALS AND METHODS

DNA constructs

The hemagglutinin (HA) epitope was introduced into mouse Kir6.2 cDNA

by a sequential overlap extension polymerase chain reaction (PCR) as

described previously (318). The epitope was inserted at position 102 of Kir6.2.

The 11 amino acids (98GDLYAYMEKGIT99) were inserted into Kir6.2 before the

HA epitope. Mammalian expression constructs were either in pcDNA3

(Invitrogen) or pEGFP/pECFP/pEYFP (Clontech). Mito-cyan fluorescent protein

(CFP) or Mito-yellow fluorescent protein (YFP) was constructed by subcloning the mitochondrial presequence (the NH2- terminal 12 amino acids of the presequence) from subunit IV of cytochrome-c oxidase into the pEGFP-N1, pECFP-N1, or pEYFP-N1 vector (Clontech). To generate green fluorescent protein (GFP), YFP, or CFP fusion proteins, full-length Kir6.2 was amplified by

PCR and ligated into pEGFP-N1, pECFP-N1, or pEYFP-N1. All constructs were verified by DNA sequencing.

Cell culture and transfection

COS-7 cells were cultured in Dulbecco’s modified Eagle’s medium

(DMEM)-F-12 supplemented with 10% fetal bovine serum, 2 mM glutamine,

and penicillin- streptomycin (113, 115). Cells were grown on coverslips or petri

dishes depending on experimental purpose. Cells were transfected with HA-

tagged (Kir6.2-HA) or GFP-fused (Kir6.2-GFP) Kir6.2 and SUR2A (pCMV) KATP

channel subunits or vectors alone by using FuGENE6 (148). Two to three days 114 later, the cells were treated with phorbol 12- myristate 13-acetate (PMA, 100

nM), PMA plus the PKC inhibitor chelerythrine (10 µM), or PMA plus the

selective PKCε inhibitor peptide myristoylated PKCε V1-2 (PKCε V1-2, 10–20

µM) for 30–60 min before the experiments. 4α-PMA (100 nM) was used as negative control.

Isolation of cardiomyocytes

Adult ventricular myocytes were isolated from Wistar rats (250–300 g) by

enzymatic dissociation (112, 115). The animal procedure was approved by the

Ohio State University Institutional Laboratory Animal Care and Use Committee.

In brief, hearts were excised and retrogradely perfused via the aorta with

oxygenated (100% O2) Tyrode solution containing (mM) 126 NaCl, 5.4 KCl, 1.0

CaCl2, 1.0 MgCl2, 0.33 NaH2PO4, 10 HEPES, and 10 glucose at 37°C. The

perfusate was then changed to a Tyrode solution that was nominally Ca2+ free but otherwise had the same composition. The hearts were perfused with the same solution containing collagenase for 20 min. The digested cell suspension was gently centrifuged, and the pellet was resuspended in culture medium 199 with 2% fetal bovine serum and 1% penicillin-streptomycin at 37°C.

Cardiomyocytes were placed in 100-mm petri dishes and incubated at 37°C in a humidified 5% CO2-95% air mix. Cardiomyocytes were subjected to various

pretreatments before mitochondrial isolation.

115 Preparation of mitochondria and mitoplasts

Mitochondrial fractions and total cell homogenate were prepared from

cultured COS-7 cells transiently transfected with Kir6.2 and SUR2A, isolated by

differential centrifugation, and further purified by 30% Percoll ultracentrifugation.

Cells transfected with vector alone were always used as a negative control. For

intact mitochondria or mitoplast colocalization experiments, COS-7 cells were

transfected with Kir6.2-GFP and SUR2A. Both cardiomyocytes and COS-7 cells

were stained with the mitochondrial marker MitoTracker (250 nM) for 15 min before fractionation. Cells after treatment were collected in an ice-cold homogenizing buffer containing (mM) 250 sucrose, 5 HEPES, and 5 EDTA, with proteinase inhibitor cocktail. Two 15-s homogenization cycles were performed on ice. The homogenate was centrifuged at 1,000 g for 10 min to remove nuclei and debris. The supernatant was centrifuged at 8,500 g for 20 min. The pellet containing the mitochondrial fraction was resuspended in the homogenizing buffer and further centrifuged at 8,500 g for 20 min. The washed mitochondria were then resuspended. For Percoll purification, the crude mitochondrial suspension (0.5 ml) was laid on the top of 10 ml of a solution containing 30% Percoll, 0.25 M sucrose, 1 mM EDTA, and 10 mM HEPES (pH

7.4). A self-generating Percoll gradient was developed by centrifugation at

95,000 g for 30 min at 4°C. The mitochondrial band was collected with a

Pasteur pipette and washed in the homogenizing buffer. Mitoplasts were

prepared from intact mitochondria by osmotic shock (264). Briefly, mitochondria

were subjected to osmotic shock for 5 min in hypotonic solution containing 116 (mM) 5 sucrose, 5 HEPES and 1 EGTA. Mitoplasts were sedimented from this

solution by centrifugation at 3,920 g for 5 min and resuspended in hypertonic

solution containing (mM) 750 KCl, 100 HEPES, and 1 EGTA (pH 7.2 with

KOH).

Western blotting

Immunoblot analysis was carried out for the mitochondrial fraction and

total cell homogenate from COS-7 cells expressing Kir6.2 and SUR2A after

treatment. The purity of mitochondria was evaluated with antibody against the

mitochondrial marker protein prohibitin, the plasma membrane marker Na+-K+-

ATPase, and the endoplasmic reticulum marker calreticulin to ensure that there was no significant contamination of other membrane fractions in mitochondria.

Mitochondria and total cell homogenate were denatured in a sample buffer, electrophoresed on 8–10% SDS-polyacrylamide gels (148), and transferred onto nitrocellulose membranes (276). The transferred blots were blocked with

5% nonfat milk in Tris-buffered saline (TBST; 150 mM NaCl, 20 mM Tris _HCl,

0.05% Tween 20, pH 7.4) and incubated for 1 h at room temperature with the antibodies against HA epitope (Roche) or Kir6.2 or SUR2A (Santa Cruz

Biotechnology) in TBST. After being washed, the blots were reacted with peroxidase-conjugated secondary antibodies (Jackson Immunuolabs) for 45 min and developed with an enhanced chemiluminescence detection system.

The specificity of antibodies for Kir6.2 and SUR2A was tested by preincubation of antibodies with their antigen peptides and/or confirmed in non-transfected 117 COS-7 cells or transfected cells with different KATP subunit. Equal loading was

confirmed by staining with Ponceau S or using antibodies against α-tubulin.

Immunofluorescence microscopy

Mitochondria or mitoplasts from cells after treatments were fixed with 4%

formaldehyde in PBS for 30 min, blocked, permeabilized in 5% goat serum in

PBS with 0.1% Triton X-100 (30 min), and labeled with primary antibody against

HA tag or Kir6.2 for 2 h. Mitoplasts isolated from COS-7 cells transfected with

Kir6.2-GFP/SUR2A were fixed only. Cells were washed three times and

labeled with fluorescence-conjugated secondary antibody for 1 h (113). We

used anti-Kir6.2 or anti-HA primary antibody for labeling KATP channel subunit

Kir6.2 and the mitochondrion-specific stain MitoTracker (250 nM) for labeling mitochondria. The cells were incubated with MitoTracker for 15 min before isolation of mitochondria/mitoplasts. Immunofluorescence was visualized with a fluorescent microscope (Nikon) or a confocal scanning laser microscope

(Zeiss).

FRET measurements

COS-7 cells were transfected with Kir6.2-CFP/SUR2A or Mito-YFP or

cotransfected with both Kir6.2-CFP/SUR2A and Mito-YFP. Images were

acquired sequentially through CFP, YFP, and FRET filter channels. Filter sets

used were the donor CFP, the acceptor YFP, and FRET. A background value

was determined from a region in each image without any cells. The background 118 value was subtracted from the raw images before FRET calculations were

carried out. Corrected FRET (FRETC) was calculated for entire images or selected regions of images, such as individual mitochondria, by using the

equation FRETC = FRET - (0.5 x CFP) - (0.3 x YFP), where FRET, CFP, and

YFP correspond to background-subtracted images of cells coexpressing CFP

and YFP acquired through the FRET, CFP, and YFP channels, respectively.

The 0.5 and 0.3 values are the fractions of bleed through of CFP and YFP fluorescence, respectively, estimated from cells expressing either CFP or YFP fusion proteins. Mean FRETC values were calculated from mean fluorescence

intensities for each selected subregion. FRETC images are presented as a

quantitative pseudocolor image. Data were analyzed with ImagePro and

MetaMorph.

Mitochondrial membrane potential

The changes in mitochondrial membrane potential (∆ψm) were monitored with the dye 5,5’,6,6’-tetrachloro-1,1’,3,3’- tetraethylbenzimidazolcarbocyanine iodide (JC-1) (233). Cells were treated with

4α-PMA, PMA, PMA plus chelerythrine, or PMA plus PKCε V1-2 and then stained with JC-1 (5 µM) at 37°C for 15 min and rinsed three times with Tyrode solution. The mitoKATP opener diazoxide or diazoxide plus the mitoKATP inhibitor

5-hydroxydecanoic acid (5-HD) were added 20 min before the measurement.

The observations were made with a fluorescence microscope. Solution

changes in this protocol were made by aspirating and replacing the contents of 119 the recording chamber. One hundred areas were selected from each image,

and the average intensity for each region was quantified by MetaMorph. The

ratio of JC-1 aggregate (red fluorescence) to monomer (green fluorescence)

intensity for each region was calculated after background subtraction. A

decrease in this ratio was interpreted as decrease of ∆ψm, whereas an

increase in the ratio was interpreted as gain in ∆ψm (277). Cells lacking red fluorescence were considered severely damaged and were excluded from analysis.

Statistics

Group data are presented as means ± SE. Unpaired t-test was used to

compare between groups. Multiple group means were compared by ANOVA

followed by post hoc least significant difference (LSD) test. Differences with a

two-tailed P<0.05 were considered statistically significant.

3. RESULTS

Localization of Kir6.2 in isolated mitoplasts

To determine whether PKC induces import of cardiac KATP to mitochondria, we took advantage of the COS-7 cell line, which lacks native KATP

channels (131), to deliver genes encoding cardiac KATP channel subunits Kir6.2

and SUR2A. We first examined whether the PKC activator PMA promotes the

mitochondrial import of recombinant cardiac KATP channels. 4α-PMA was used

120 as a negative control. To avoid colocalization of Kir6.2 with the mitochondrial

marker MitoTracker due to close apposition between Kir6.2-containing vesicles

and mitochondria, we performed colocalization experiments in mitoplasts

(devoid of outer membrane) freshly isolated from COS-7 cells transfected with

SUR2A and Kir6.2 with GFP fused to its COOH terminus. The cells were

pretreated with 4α-PMA, PMA, or PMA plus PKCε V1-2 for 60 min and were

stained with MitoTracker before isolation of mitochondria. Mitoplasts were

obtained by subjecting intact mitochondria to hypotonic shock (see

MATERIALS AND METHODS for details, section 2). About 20–30 images in the same treatment group (2 or 3 coverslips) from each experiment were analyzed.

Three independent experiments were conducted. Without PMA (100 nM, data not shown) or with 4α-PMA (100 nM) treatment, only a small portion of fluorescent Kir6.2 was localized in mitoplasts that were stained with

MitoTracker. PMA treatment significantly increased mitoplast localization of

Kir6.2 (Fig. 4.1). Quantification analysis revealed that the activation of PKC by

PMA increased the number of Kir6.2-positive mitoplasts by ~100% (Fig. 4.1C;

205.78 ± 8.84% vs. 4α-PMA, P<0.01). To identify which PKC isozyme is involved in PKC-induced increase in mitochondrial localization of Kir6.2, we studied the effect of the selective PKCε peptide inhibitor PKCε V1-2. PKCε has been shown to modulate KATP channels (125) and is linked to IPC (223). As

expected, the effect of PMA on mitochondrial localization of Kir6.2 was completely prevented by PKCε V1-2 (10 µM; 110.39 ± 10.33%). These results

121 indicate that Kir6.2-containing KATP channels are localized in mitochondria at low levels under normal conditions and can be enhanced by the activation of

PKC isozyme PKCε.

122

Figure 4.1. Immunofluorescence microscopy of mitochondrial localization of Kir6.2 in mitoplasts isolated from COS-7 cells. A: transmitted and fluorescent images of a mitoplast (mitochondria devoid of outer membrane) labeled with the mitochondrial marker MitoTracker. Scale bar: 3 µm. B: COS-7 cells were transfected with sulfonylurea receptor (SUR)2A/Kir6.2-green fluorescent protein (GFP), and mitoplasts were prepared after COS-7 cells were treated with agents as indicated. Results are representative of 3 independent experiments. About 20–30 images in the same treatment group (2 or 3 coverslips) from each experiment were analyzed. Scale bar: 5 µm. PMA, phorbol 12-myristate 13-acetate; PKCε, protein kinase C isoform ε. C: % of mitoplasts (MitoTracker positive) with Kir6.2-positive fluorescence are shown

(data were normalized to the control without treatment). **P < 0.01 vs. 4α-PMA.

123

Figure 4.1. Immunofluorescence microscopy of mitochondrial localization of Kir6.2 in mitoplasts isolated from COS-7 cells.

124 FRET analysis of Kir6.2 in mitochondrial inner membrane

Although our observation that Kir6.2 is localized in mitoplasts indicates

that Kir6.2 is present in mitochondrial inner membrane, there is a possibility that

part of the outer membranes of mitochondria may still be attached to mitoplasts.

To further define whether KATP channels are indeed localized in mitochondrial

inner membrane, we tested in COS-7 cells whether CFP fused to the COOH

terminus of Kir6.2 is sufficiently close to Mito-YFP that is targeted into the

mitochondrial matrix to yield FRET (90). If KATP channels are in the outer

membrane of mitochondria, the fluorescent protein fused to the cytoplasmic

domain of Kir6.2 would be exposed to the cytosol. If these channels actually

reside in the mitochondrial inner membrane, their cytoplasmic domains should

be exposed to the matrix enclosed by the mitochondrial inner membrane—the

equivalent of the cytosol for these endosymbiotic organelles. As shown in Fig.

4.2, FRETC was minimal in all regions of cells that were not stimulated with

PMA (Fig. 4.2A), indicating that there was no significant amount of Kir6.2 localized in mitochondria. In particular, FRET signals were not observed in the mitochondria enriched with Mito-YFP or in the Golgi area containing a large amount of Kir6.2-CFP. This validates the accuracy of our method of correction for the non-FRET component of the FRET images. Incubation of cells with PMA at 37°C for 30–60 min led to rapid import of Kir6.2 protein into mitochondria and exhibited FRET signals due to energy transfer from CFP at the COOH terminus of Kir6.2 to YFP in the mitochondrial matrix. To compare FRET efficiencies, the

FRETC was normalized to the intensity of Kir6.2-CFP after background 125 subtraction. The FRET signal between Mito-CFP and Mito-YFP was comparable to that between Kir6.2-CFP and Mito-YFP after PMA treatment for

60 min, indicating that by then the KATP channels targeted to the mitochondria

had fully exposed the CFP at the COOH terminus of Kir6.2 to the matrix. Similar

FRET signals were observed in cells expressing Mito-YFP, SUR2A, and Kir6.2

with CFP fused to its NH2 terminus (data not shown). For dipole-dipole coupling

to induce energy transfer from donor to acceptor, the fluorophore at the NH2 or

COOH terminus of Kir6.2 must be within 6–10 nm of the fluorophore within the mitochondrial matrix (90, 219). These findings strongly suggest that PKC activation causes insertion of Kir6.2-containing KATP channels into the

mitochondrial inner membrane.

126

Figure 4.2. A: Proximity of Kir6.2-cyan fluorescent protein (CFP) (left) to

Mito-yellow fluorescent protein (YFP) in mitochondrial matrix (center)

yielded fluorescence resonance energy transfer (FRET) signals (right, in

quantitative pseudocolor) in cells treated with PMA (100 nM) for 60 min

(bottom) but not in cells without PMA treatment (top). FRETC, corrected FRET.

B: FRET values were normalized by dividing FRETC by the mean intensity of

CFP after background subtraction. The FRET signal between Mito-YFP and

Mito-CFP provides calibration for a strong signal produced by 2 matrix

fluorescent proteins.

127

Figure 4.2. Proximity of Kir6.2-cyan fluorescent protein (CFP) (left) to Mito- yellow fluorescent protein (YFP) in mitochondrial matrix (center) yielded fluorescence resonance energy transfer (FRET) signals

128 Alterations of Kir6.2 protein in mitochondria

To assess the effect of PKC activation on Kir6.2 protein level, immunoblot analysis was carried out for Percoll-purified mitochondria fraction and total cell homogenate prepared from COS-7 cells treated with 4α-PMA,

PMA, PMA plus chelerythrine, or PMA plus PKCε V1-2 for 60 min. Enrichment of mitochondria was established by labeling with anti-prohibitin, a molecular marker of mitochondria. With anti-HA antibody, which is known as a very good antibody, a single band at 37 KDa was detected in the mitochondrial fraction as well as total cell homogenate (Fig. 4.3, A and B). Anti-HA antibody did not show any nonspecific band in our study. Interestingly, the band in the mitochondrial fraction from COS-7 cells treated with PMA for 60 min exhibited a more intense signal than that from COS-7 cells without PMA treatment or with 4α-PMA treatment. The PMA effect was significantly eliminated not only by the general

PKC inhibitor chelerythrine (10 µM) but also by the selective PKCε inhibitor peptide PKCε V1-2 (10 µM). However, there was no difference in the Kir6.2 band in total cell homogenate between cells treated with 4α-PMA and PMA. No

Kir6.2 band was detected from cells without transfection. Similar results were obtained with antibody against Kir6.2 (data not shown). Figure 4.3B shows average Kir6.2 band intensity normalized to control from three independent experiments. PMA induces almost 100% increase in Kir6.2 protein level in the mitochondrial fraction compared with that in the control group (216.11 ±

20.29%), which is consistent with the data in mitochondrial localization of

Kir6.2, but has no effect on total cellular Kir6.2 protein level. When we tried a 129 similar experiment with anti-SUR2A antibody, there was a single faint band of

SUR2A at a molecular mass of 150 KDa in the mitochondrial fraction under control conditions, which was not affected by PMA pretreatment (Fig. 4.3C).

This result indicates that only a minimum amount of SUR2A exists in

mitochondria, if there is any. However, we did not detect SUR2A with antibody

against SUR2, which is similar to the report by Lacza et al., (147) showing that

SUR2 was not present in mitochondria with the same antibody (purchased from

the same company). This discrepancy is likely due to the differential sensitivities of two different antibodies. The specificity of anti-SUR2A antibody

was confirmed in nontransfected COS-7 cells and COS-7 cells transfected with

SUR2B showing no cross-reactivity with SUR2B. We noted that the anti-SUR2A

antibody is very specific and more sensitive than the anti-SUR2 antibody.

These results suggest that Kir6.2 is present in mitochondria and PKCε

activation by PMA further increases Kir6.2 protein level in mitochondria. This

increase may not be due to increased protein synthesis for Kir6.2 since Kir6.2

in the total cellular fraction was not changed after PMA treatment.

130

Figure 4.3. Immunoblot analysis of Kir6.2 protein in mitochondria. COS-7 cells were transfected with Kir6.2-hemagglutinin (HA)/SUR2A. A: total cell homogenate (cellular fraction, CF) and mitochondrial fraction (MF) of COS-7 cells were prepared after cells were treated with 4α-PMA, PMA, PMA + chelerythrine (Che), or PMA + PKCε V1-2. These proteins were separated on an 8–10% SDS-polyacrylamide gel and immunoblotted with antibodies against

HA (top) and prohibitin (bottom). Thirty micrograms of protein for each lane was loaded for cellular fraction, whereas 20 µg of protein was loaded for mitochondrial fraction analysis. B: signal intensity of Kir6.2 in cellular and mitochondrial fractions was quantified as values normalized to control from 3 independent experiments with Image J software. C: similar Western blot analysis is shown for SUR2A in mitochondrial fraction with antibody against

SUR2A.

131

Figure 4.3. Immunoblot analysis of Kir6.2 protein in mitochondria 132 Kir6.2-dependent changes in mitochondrial membrane potential

To determine whether Kir6.2-containing KATP channels in mitochondria

are functional, we measured the changes in ∆ψm in intact COS-7 cells

transfected with Kir6.2 and SUR2A, using the potential-sensitive dye JC-1.

About 100 regions were selected from each image, which contains 10–20 cells,

and 5 images were taken from each well. Three independent experiments were

conducted, and each experiment had triplicate wells of the same treatment. The

average intensity for each image was quantified after background subtraction.

The green fluorescence of monomer indicates low potential (depolarized), and

the red fluorescence of J-aggregate indicates high potential (hyperpolarized).

Since KATP channels are usually not active because of inhibition by matrix ATP,

the mitochondrion-selective KATP opener diazoxide was used to activate KATP in mitochondria. Diazoxide has been shown to confer the cardioprotection of IPC by selectively activating mitoKATP channels (86, 199, 292). If Kir6.2/SUR2A-

containing KATP channels are localized in mitochondrial inner membrane and

involved in the protection of IPC, this channel is expected to be sensitive to

activation by diazoxide. The KATP-dependent changes were verified by the

mitochondrial KATP-selective inhibitor 5-HD. In the absence of diazoxide, the JC-

1 ratio (J-aggregate/monomer) in cells pretreated with PMA for 60 min was not

significantly different from the control cells pretreated with 4α-PMA (Fig. 4.4B).

However, diazoxide (100 µM), within 15 min of application, decreased the JC-1

ratio significantly in PMA-treated cells (potential decreased, green fluorescence

increased; Fig. 4.4A) compared with that in 4α-PMA-pretreated cells with 133 diazoxide (Fig. 4.4; 53.04 ± 10.22% vs. 97.87 ± 9.24%, PMA + diazoxide vs.

4α–PMA + diazoxide, P 0.01). When diazoxide was added to the solution 5 min

after the application of 5-HD, the diazoxide-induced decrease in ∆ψm in cells

treated with PMA was inhibited not only by 5-HD (500 µM; 83.79 ± 6%) but also by pretreatment with PKCε V1-2 (20 µM; 82.61 ± 4.56%). Either 5-HD application or PKCε V1-2 pretreatment alone did not have any effect on ∆ψm

(data not shown). These data show that ∆ψm in cells pretreated with PMA was

not different from that in cells with 4α-PMA but could be decreased by activation of KATP channels in mitochondria with the mitoKATP opener diazoxide after PMA

pretreatment, an effect that was inhibited by the mitoKATP inhibitor 5-HD.

Furthermore, the diazoxide effect in PMA-treated cells was not seen in nontransfected COS-7 cells, indicating that the effect of diazoxide on ∆ψm was

due to activation of Kir6.2-containing KATP channels in mitochondrial inner

membrane. The finding that diazoxide induces KATP-dependent depolarization

of ∆ψm after activation of PKCε by PMA pretreatment but not in the control

group with 4α-PMA treatment suggests that K+ influx caused by diazoxide

under basal conditions may not be sufficient to cause significant changes in

∆ψm under our experimental conditions. This result does not exclude the

possibility that other mitochondrial functions may be altered by diazoxide under

normal conditions or in isolated mitochondria where no translocation occurs.

Furthermore, the observation that PMA pretreatment alone did not change ∆ψm

suggests that KATP channels in mitochondrial inner membrane are normally

inhibited by matrix ATP even though PMA may increase the number of 134 functional Kir6.2-containing KATP channels in mitochondria. It should be noted that PMA was used only for pretreatment and was washed out before measurement of ∆ψm with diazoxide.

135

Figure 4.4. PMA-induced KATP-dependent changes in mitochondrial

membrane potential (∆ψm). A: COS-7 cells transfected with SUR2A/Kir6.2-HA

were incubated with mitochondrial potential-sensitive dye JC-1 (5 µM) for 15

min after various pretreatments as indicated. Diaz, diazoxide. Scale bar: 20 µm.

B: ∆ψm expressed as a ratio of J-aggregate to monomer fluorescence in different treatments. Three independent experiments were conducted, and each experiment had triplicates of the same treatment where ~15 images were collected. The average fluorescent intensity for each image was calculated after background subtraction. Data from different treatment groups were normalized to the control group without PMA treatment. **P <0.01 vs. 4α-PMA.

136

Figure 4.4. PMA-induced KATP-dependent changes in mitochondrial

membrane potential (∆ψm)

137 Localization of Kir6.2 in isolated mitochondria from cardiomyocytes

To exclude the possibility that overexpression of recombinant KATP channels contributes to our findings, we performed similar colocalization experiments in isolated mitochondria from rat adult cardiomyocytes. Because of very tightly packed myofibrils around mitochondria and other organelles in cardiomyocytes, it is very difficult to study subcellular localization of membrane proteins in isolated cardiomyocytes. We therefore performed colocalization experiments in isolated mitochondria. Cardiomyocytes were subjected to pretreatment with 4α-PMA, PMA, or PMA plus PKCε V1-2 for 60 min before mitochondrial isolation. To avoid bias in our analysis, 30 images were randomly collected from the triplicates of the same treatment in each independent experiment. Two to five independent experiments depending on treatments were conducted, and ~1,000 mitochondria were counted for statistical analysis.

Anti-Kir6.2 antibody was used to detect native Kir6.2 protein. As Fig. 4.5A shows, PMA increased mitochondrial localization of Kir6.2, as indicated by increased yellow staining in the merged image. The pretreatment of PMA alone with PKCε V1-2 blocked the PMA effect. Quantitative analysis by normalizing

Kir6.2-positive mitochondria to the control without PMA treatment showed that

PMA significantly increased the number of Kir6.2-positive mitochondria (Fig.

4.5B; 240.31 ± 13.6% vs. 4α-PMA, P<0.01). This PMA effect was almost completely prevented by pretreatment with the PKC inhibitor chelerythrine (10

µM; data not shown) as well as the selective PKCε inhibitor peptide PKCε V1-2

(10 µM; 120.02 ± 9.77% vs. PMA group, P<0.01). These results are consistent 138 with our observation in COS-7 cells transiently transfected with Kir6.2/SUR2A and support the scenario that Kir6.2-containing KATP channels are localized in mitochondria and this mitochondrial localization was further enhanced by PKC activation.

139

Figure 4.5. Immunofluorescence microscopy of mitochondrial localization

of Kir6.2 in mitochondria isolated from rat adult cardiomyocytes. A: fluorescent images of mitochondria with Kir6.2 labeled with anti-Kir6.2 antibody.

The intact mitochondria were prepared after treatment of rat cardiomyocytes with agents as indicated. Results are representative of 2–5 independent experiments depending on the treatments, and each experiment had triplicates

of the same treatment with 30 images taken. Scale bar: 10 µm. B: normalized

% of mitochondria with Kir6.2-positive fluorescence over control. **P < 0.01 vs.

4α-PMA.

140

Figure 4.5. Immunofluorescence microscopy of mitochondrial localization of Kir6.2 in mitochondria isolated from rat adult cardiomyocytes

141 4. DISCUSSION

The present study points to a novel trafficking mechanism for dynamic

regulation of KATP channel number: PKC-induced import of KATP channels into mitochondria. We show that the KATP channel pore-forming subunit Kir6.2 is localized in mitochondria, using both cardiomyocytes and KATP channel deficient

COS-7 cells transiently transfected with Kir6.2 and SUR2A. Most importantly,

we demonstrate that activation of PKC isozyme PKCε significantly increases

the number of functional Kir6.2-containing KATP channels in mitochondrial inner

membrane, possibly by promoting import of Kir6.2 to mitochondria from cytosol.

One likely scenario is that KATP channels are simultaneously targeted to not only cell surface but also mitochondrial inner membrane after their synthesis in the cytosol and PKC activation enhances the import of these channels to

mitochondria. The specific pharmacological and biophysical properties of KATP

channels at the plasma membrane and mitochondrial inner membrane may

differ depending on the environment in which they reside and the regulatory

factors with which they are associated (55).

The function of ion channels depends critically on not only channel

activity but also the number of functional channels. Our observation that

activation of PKCε by PMA increases mitochondrial localization of Kir6.2- containing KATP channels points out a novel trafficking mechanism in regulation of KATP channel function. However, this finding does not exclude the possibility

that PKC may directly interact with the channel. Actually, PMA has been shown

to activate cell surface Kir6.2/SUR2A-containing KATP channels (112, 163) as 142 well as mitoKATP channels (125). Under our experimental conditions, PMA was

washed out before we tested diazoxide-induced changes in ∆ψm. If there is any

residue of PMA, it may not be enough to activate KATP channels and cause

alterations in ∆ψm. The present study mainly attempted to address whether KATP

channels translocated by PKC activation are functional.

Our data and that of others (54, 147, 265, 319) show that there is a basal

level of Kir6.2-containing KATP channels in mitochondria. It is therefore likely

that K+ channel openers or stimulators such as PMA may directly activate these channels under normal conditions or in isolated mitochondria where no translocation occurs. In the present study, diazoxide alone did not change ∆ψm

in control cells without PMA or pretreated with 4α-PMA. One explanation is that

K+ influx induced by diazoxide in COS-7 cells may not be high enough to cause

significant changes in ∆ψm but may produce changes in other mitochondrial

functions that we did not measure, such as matrix volume. On the other hand,

our observation that diazoxide caused changes in ∆ψm after PMA treatment

indicates that the KATP channels translocated by PMA are functional by

increasing K+ influx. However, our data do not imply that the same functional

consequence in mitochondria (∆ψm) caused by diazoxide-induced opening of

KATP channels would occur in native cardiomyocytes. We employed COS-7

cells with expressed recombinant KATP channels, which may be different from

cardiomyocytes in terms of the extent of K+ influx. We therefore do not believe

that our data conflict with that of others showing no significant changes in ∆ψm

by diazoxide induced opening of mitoKATP (50, 146) from isolated heart 143 mitochondria. Nevertheless, the purpose of the JC-1 experiment was not to

define ultimate changes in mitochondrial function by diazoxide but rather to

study whether the translocated Kir6.2-containing KATP channels are functional.

Diazoxide was used as a tool to activate KATP channels since KATP channels are

normally inactive.

Although it is generally assumed that the Kir6.2/SUR2Acontaining KATP

channel is localized exclusively at the sarcolemma there is evidence that Kir6.2 or SUR2A is present in mitochondria. Several groups have reported that Kir6.2

has been found more or less in isolated heart mitochondria (54, 147, 265, 319).

One study has reported that Kir6.2 is not part of the components of mitoKATP

channels in the rabbit heart (256). The discrepancy may be due to the relatively

low level of Kir6.2-containing KATP channels in mitochondria under normal

conditions. The technical difficulties in dissecting a small amount of membrane-

targeted proteins localized in mitochondria with conventional

immunofluorescence microscopy may contribute to the inconsistency,

especially in adult cardiomyocytes with tightly packed myofibrils and

suborganelles. Differential sensitivity of antibodies against individual KATP

subunits may also likely be a reason for the discrepancy. With regard to the

SUR subunit, less is known about its presence in mitochondria because of the

difficulty in getting antibody specific to SUR2A or SUR2B. We tested anti-

SUR2A antibody, recently available commercially, and found a faint band of

SUR2A present in mitochondria of COS-7 cells by Western blot, which is

consistent with some reports (55, 265). Similar to another study (147), we did 144 not detect a SUR2 band with anti-SUR2 antibody in mitochondria. Because of

the lack of series of studies with different approaches, we cannot make a

conclusion at the present time that SUR2A is indeed localized in mitochondria.

The contribution of Kir6.1 to mitoKATP channels is also under debate

(184, 268). The Kir6.1/SUR2B-based channels do not exhibit sensitivity to the inhibition by intracellular ATP (310), whereas the mitoKATP channels recorded in

mitochondria inner membrane from liver and reconstituted lipid bilayer from

purified heart mitochondria are very sensitive to ATP, a characteristic feature of

the KATP channel pore-forming subunit Kir6.2 (56). Although Kir6.1 protein level

was increased by ischemia, this only occurred after prolonged ischemia (60 min

of ischemia followed by 24–72 h of reperfusion). Sublethal ischemia (15–30 min

followed by 24 h of reperfusion), which is equivalent to the time course of PMA

treatment in the present study, did not induce Kir6.1 expression (5). It is

therefore less likely that Kir6.1 would contribute to functional mitoKATP channels

even though it was found in mitochondria (268). Furthermore, knockout of

Kir6.1 gene in intact mice does not disrupt mitoKATP opening (184).

Nevertheless, the present study focuses on the regulatory mechanism of Kir6.2- containing KATP channels in mitochondria and does not exclude the possibility

that other protein(s) may form the part of KATP channel complex in mitochondria.

The dual distribution at both plasma membrane and mitochondria has

been reported for other ion channel proteins, such as Kv1.3 (271), the Ca2+- activated BK potassium channels (264), the voltage-dependent anion channels

(20, 63), and connexin43 (26, 158). Interestingly, connexin43 in mitochondria 145 has recently been shown to be increased by IPC, possibly through

phosphorylation (26). These channels lack the NH2-terminal mitochondrial

targeting sequence but are targeted to mitochondrial inner membrane by

unknown mechanisms. Many mitochondrial proteins of the inner membrane

such as mitochondrial carriers lack the NH2-terminal targeting sequence but

instead contain sorting and targeting information throughout the mature protein,

which is hardly recognizable (32, 58). The targeting information seems to

comprise several distant amino acids spread throughout the entire protein

(278). Similarly, Kir6.2 does not exhibit a classic NH2-terminal mitochondrial

targeting presequence according to the prediction from TargetP (71) and may

contain unidentified internal targeting sequences. Although specific

posttranslational protein modifications and differential splicing have been

hypothesized to account for organelle protein targeting, the mechanisms

responsible for the simultaneous targeting of a protein to different cellular compartments are not known and are beyond the focus of the present study.

Import of most nucleus-encoded proteins into mitochondria is highly

regulated and mediated by mitochondria targeting sequence with the aid of a handful of distinct complexes including molecular chaperones (137). Both heat shock protein (HSP)70 and HSP90, which are linked to cardioprotection, have been shown to facilitate mitochondrial protein import to mitochondria (154, 278,

286, 288, 314, 315). HSP90 and HSP70 specifically interact with the mitochondrial protein import receptor TOM70 at the outer membrane and are required for translocation of precursor proteins. Furthermore, PKC has been 146 shown to be involved in HSP-mediated cardioprotection (45, 302). The present

finding points to an important mechanism in Kir6.2-containing KATP channel-

mediated cardioprotection: PKCε- induced mitochondrial import. It should be

noted that the present study was not aimed at defining the molecular identity of putative mitoKATP but rather at studying the regulatory mechanism involved in

Kir6.2-containing KATP channel trafficking. This observation does not exclude

the possibility that a distinct mitoKATP or other KATP component(s) in

mitochondria may exist. It has been reported that succinate dehydrogenase

may be a part of a mitoKATP macromolecular supercomplex that contributes to

cardioprotection (9). A Ca2+-activated K+ channel has also been found in

mitochondrial inner membrane and linked to cardioprotection (308).

The evidence supporting a protective role for the Kir6.2- containing KATP

channels has been provided by studies using KATP-deficient COS-7 cells and

Kir6.2-knockout mice. By cotransfection of Kir6.2/SUR2A genes, Jovanovic et

al., (131) demonstrated that delivery of Kir6.2 and SUR2A genes into COS-7

cells resulted in protection against hypoxia-reoxygenation injury as a result of

inhibition of intracellular Ca2+ loading. Studies in Kir6.2-knockout mice (99, 269) showed a failure of IPC to reduce infarct size and preserve contractile recovery in Kir6.2-deficient mice. The activation of PKC has also been linked to cardioprotection (223) and KATP channels (163, 291). A number of possibilities

have been suggested as to how opening the mitoKATP results in cardioprotection, such as reduced Ca2+ overload in mitochondria by

147 depolarizing, mild uncoupling and reduced free radical production, and moderate swelling and increased ATP production (50).

It is known that movement of key proteins into mitochondria during apoptosis is essential for the regulation of apoptosis (42). Mitochondria are increasingly recognized as key players in cell survival (92). Most of the mitochondrial proteins are nuclear encoded, synthesized in the cytosol, and transported to mitochondria by translocases (137, 166, 220). Given that mitochondria are a critical site for cardioprotection; our observations indicate that the dynamic regulation of protein trafficking to mitochondria represents a novel mechanism in cardioprotection of IPC.

ACKNOWLEDGMENTS

We are grateful to S. Seino for providing Kir6.2 and SUR2A clones and to R. Y. Tsien for the kind gifts of Mito-CFP and Mito-YFP constructs. We thank

Douglas Pfeiffer for comments on the mitochondrial aspect of the study. We also thank Chen Gu and Ofer Wiser for technical help on the FRET assay and

Lily Jan for valuable support and suggestions. This study was supported in part by a research grant from the American Heart Association.

148

CHAPTER 5

CONCLUSIONS AND FUTURE DIRECTIONS

The novel findings of my dissertation work are:

1) KATP channels in the heart are predominantly present in cholesterol-rich

membrane microdomains called as caveolae

2) This caveolar localization is essential for adenosine activation of KATP

channels

3) Caveolin-3 has an inhibitory effect on activation of KATP channels

4) PKC activation induces import of Kir6.2 containing KATP channels to

mitochondria.

The cellular plasma membrane is not homogenous throughout. It is

extensively compartmentalized based on specific lipid and protein composition

(rafts/caveolae). These membrane microdomains can segregate and

concentrate specific signaling pathway components for rapid and more efficient

communication between extracellular environment and intracellular

homeostasis. For ion channels, their subcellular-localization, trafficking and

local environment can profoundly affect their function and properties. ATP-

sensitive potassium channels link the cellular metabolism with the membrane 149 excitability in various cell-types. In the heart, not only they shorten APD in

response to ischemia, their cytoprotective role has gained increasing

recognition in recent years. Traditionally classified just as a plasma membrane

protein, these KATP channels have been shown to be present in other subcellular organelles. Their localization can alter their response to various

stimuli. With this theme in mind, my dissertation work aimed to investigate

three related areas: (1) To characterize KATP channel subcellular localization on

plasma membrane caveolae and mitochondria; (2) Regulation of KATP channel

function by these membrane caveolae and associated protein caveolin-3; (3)

Trafficking of KATP channels to mitochondria in response to PKC stimulation.

The overall project was broadly divided into two parts based on membrane

microdomain or mitochondrial localization and these parts were separately

investigated.

We first studied the specific membrane microdomain localization of KATP channel using different techniques to rule out any confounding factors related to specific techniques utilized for the investigation. We report for the first time that cardiac KATP channels, consisting of Kir6.2/SUR2A are predominantly

concentrated in cholesterol-rich caveolar microdomains of the plasma

membrane. To understand the functional significance of this localization, we

employed adenosine which is one of the main KATP channel regulator under

ischemic conditions. Since adenosine receptors are also present in caveolae,

localization of both KATP channel and adenosine in the same compartment

signifies that probably this spatial closeness is important for regulation of KATP 150 channel function. And indeed, our patch clamp studies confirmed that intact

caveolar structure is important for adenosine activation of KATP channel. Since

many of the other KATP channel modulators (besides adenosine) are also

present in caveolae, it would be interesting to know if those regulators also

follow the same paradigm of caveolae dependent signaling similar to

adenosine.

Further, our discovery of caveolin-3 mediated inhibition of KATP current

using a simplified system revealed another interesting yet paradoxical aspect of

KATP channel regulation by caveolae. Activation of only 1% of total membrane

KATP channel can shorten the APD by 50%. Since excessive KATP channel

activation would be detrimental for the heart leading to silencing of electrical

activity and arrhythmias. Caveolae dependent control of both activation and

inhibition of KATP channel seems quite logical and brings forth the fact that KATP channels in the heart are very tightly regulated.

It is well known that KATP channels play an important role in cardioprotection and ischemic preconditioning (93). Importance of caveolin proteins and

caveolae mediated signal transduction in cardioprotection is just beginning to

unravel (213, 215). It would be interesting to determine the relative contribution

of caveolar and non-caveolar KATP channels to cardioprotection. Nevertheless,

there are several other related questions which we wish to investigate in our

future studies. Most important of these is, how are KATP channels targeted to

caveolae? Secondly, what is the kinetics and dynamics of KATP channel stay in

caveolae? Identifying the stimuli and mechanisms which increase the levels of 151 KATP channels in caveolae, and increase their residence time in caveolae will

have significant therapeutic implications. Lastly, how does the caveolin-3

protein mediate its inhibitory effect on KATP channel? We identified that caveolin-3 scaffolding domain is important in this interaction, but we don’t know which subunit of KATP channels it binds to.

In our effort to characterize the sub-cellular localization of cardiac KATP

channels, we detected a small amount of Kir6.2 in mitochondria of cardiac

myocytes and transfected COS-7 cells under basal conditions. But it does not

affect the mitochondrial function significantly. We showed that this localization

of channel protein in mitochondria can be significantly increased by specific

stimuli namely activation of PKCε here. Further, KATP channels translocated to

mitochondria were shown to be functional and affected the mitochondrial

potential when pretreated with PMA. This corroborates the statement made

above that KATP channel localization in a cardiac myocyte is very dynamic and a

controlled process which should be investigated further in detail under various

physiological conditions. In a latter collaborative study, I found that indeed

hypoxic preconditioning (one of the activator of endogenous cardioprotective mechanism similar to IPC) enhances Kir6.2 targeting to mitochondria in a heat-

shock protein 90 (HSP90)-dependent manner.

Collectively, our elucidation of caveolar and mitochondrial localization;

and regulation of KATP channels will significantly enhance our knowledge of the

complexity of regulation of this metabolic sensor. This will be potentially useful

in understanding the endogenous cardioprotective strategies employed by the 152 heart and their alteration in some pathological conditions. Therapeutic strategies can be further designed to enhance these cardioprotective regulatory mechanisms. The hypothetical model based on the findings of this thesis work is shown in fig. 5.1.

153

Figure 5.1. Hypothetical model for the targeting of KATP channels to

plasma membrane, caveolar microdomains and mitochondria. After

synthesis in the endoplasmic reticulum, KATP channels (Kir6.2/SUR2A) are

targeted to plasma membrane. Caveolae help in compartmentalizing KATP channels with its modulator adenosine A1 receptor. Caveolin-3 can bind to and

inhibit KATP channel independent of its localization to caveolae also. PKC

activation induces import of Kir6.2 into mitochondria. Abbreviations: Cav-3 =

caveolin-3, TIM/TOM = protein translocation machinery of mitochondria.

154

BIBLIOGRAPHY

1. Abraham MR, Selivanov VA, Hodgson DM, Pucar D, Zingman LV, Wieringa B, Dzeja PP, Alekseev AE, and Terzic A. Coupling of cell energetics with membrane metabolic sensing. Integrative signaling through creatine kinase phosphotransfer disrupted by M-CK gene knock-out. The Journal of biological chemistry 277: 24427-24434, 2002.

2. Aguilar-Bryan L, and Bryan J. Molecular biology of adenosine triphosphate-sensitive potassium channels. Endocrine reviews 20: 101-135, 1999.

3. Aguilar-Bryan L, Clement JPt, Gonzalez G, Kunjilwar K, Babenko A, and Bryan J. Toward understanding the assembly and structure of KATP channels. Physiological reviews 78: 227-245, 1998.

4. Aizawa K, Turner LA, Weihrauch D, Bosnjak ZJ, and Kwok WM. Protein kinase C-epsilon primes the cardiac sarcolemmal adenosine triphosphate-sensitive potassium channel to modulation by isoflurane. Anesthesiology 101: 381-389, 2004.

5. Akao M, Otani H, Horie M, Takano M, Kuniyasu A, Nakayama H, Kouchi I, Murakami T, and Sasayama S. Myocardial ischemia induces differential regulation of KATP channel gene expression in rat hearts. The Journal of clinical investigation 100: 3053-3059, 1997.

6. Alekseev AE, Hodgson DM, Karger AB, Park S, Zingman LV, and Terzic A. ATP-sensitive K+ channel channel/enzyme multimer: metabolic

155 gating in the heart. Journal of molecular and cellular cardiology 38: 895-905, 2005.

7. Allen DG, Morris PG, Orchard CH, and Pirolo JS. A nuclear magnetic resonance study of metabolism in the ferret heart during hypoxia and inhibition of glycolysis. J Physiol 361: 185-204, 1985.

8. Anderson RG. The caveolae membrane system. Annual review of biochemistry 67: 199-225, 1998.

9. Ardehali H, Chen Z, Ko Y, Mejia-Alvarez R, and Marban E. Multiprotein complex containing succinate dehydrogenase confers mitochondrial ATP-sensitive K+ channel activity. Proceedings of the National Academy of Sciences of the United States of America 101: 11880-11885, 2004.

10. Armstrong SC, Liu GS, Downey JM, and Ganote CE. Potassium channels and preconditioning of isolated rabbit cardiomyocytes: effects of glyburide and pinacidil. Journal of molecular and cellular cardiology 27: 1765- 1774, 1995.

11. Ashcroft FM. Adenosine 5'-triphosphate-sensitive potassium channels. Annual review of neuroscience 11: 97-118, 1988.

12. Ashcroft FM. Exciting times for PIP2. Science (, NY 282: 1059-1060, 1998.

13. Ashcroft SJ, and Ashcroft FM. Properties and functions of ATP- sensitive K-channels. Cellular signalling 2: 197-214, 1990.

14. Babenko A, and Vassort G. Enhancement of the ATP-sensitive K+ current by extracellular ATP in rat ventricular myocytes. Involvement of adenylyl cyclase-induced subsarcolemmal ATP depletion. Circulation research 80: 589- 600, 1997.

15. Baines CP, Zhang J, Wang GW, Zheng YT, Xiu JX, Cardwell EM, Bolli R, and Ping P. Mitochondrial PKCepsilon and MAPK form signaling modules in the murine heart: enhanced mitochondrial PKCepsilon-MAPK

156 interactions and differential MAPK activation in PKCepsilon-induced cardioprotection. Circulation research 90: 390-397, 2002.

16. Baker JE, Su J, Fu X, Hsu A, Gross GJ, Tweddell JS, and Hogg N. Nitrite confers protection against myocardial infarction: role of oxidoreductase, NADPH oxidase and K(ATP) channels. Journal of molecular and cellular cardiology 43: 437-444, 2007.

17. Balijepalli RC, Foell JD, Hall DD, Hell JW, and Kamp TJ. Localization of cardiac L-type Ca(2+) channels to a caveolar macromolecular signaling complex is required for beta(2)-adrenergic regulation. Proceedings of the National Academy of Sciences of the United States of America 103: 7500-7505, 2006.

18. Barbuti A, Gravante B, Riolfo M, Milanesi R, Terragni B, and DiFrancesco D. Localization of pacemaker channels in lipid rafts regulates channel kinetics. Circulation research 94: 1325-1331, 2004.

19. Barry WH. Na"Fuzzy space": does it exist, and is it important in ischemic injury? Journal of cardiovascular electrophysiology 17 Suppl 1: S43-S46, 2006.

20. Bathori G, Parolini I, Tombola F, Szabo I, Messina A, Oliva M, De Pinto V, Lisanti M, Sargiacomo M, and Zoratti M. Porin is present in the plasma membrane where it is concentrated in caveolae and caveolae-related domains. The Journal of biological chemistry 274: 29607-29612, 1999.

21. Baukrowitz T, Schulte U, Oliver D, Herlitze S, Krauter T, Tucker SJ, Ruppersberg JP, and Fakler B. PIP2 and PIP as determinants for ATP inhibition of KATP channels. Science (New York, NY 282: 1141-1144, 1998.

22. Baxter GF, and Ebrahim Z. Role of bradykinin in preconditioning and protection of the ischaemic myocardium. British journal of pharmacology 135: 843-854, 2002.

23. Bessman SP, and Geiger PJ. Transport of energy in muscle: the phosphorylcreatine shuttle. Science (New York, NY 211: 448-452, 1981.

157 24. Bethell HW, Vandenberg JI, Smith GA, and Grace AA. Changes in ventricular repolarization during acidosis and low-flow ischemia. The American journal of physiology 275: H551-561, 1998.

25. Billman GE. A comprehensive review and analysis of 25 years of data from an in vivo canine model of sudden cardiac death: implications for future anti-arrhythmic drug development. Pharmacology & therapeutics 111: 808-835, 2006.

26. Boengler K, Dodoni G, Rodriguez-Sinovas A, Cabestrero A, Ruiz- Meana M, Gres P, Konietzka I, Lopez-Iglesias C, Garcia-Dorado D, Di Lisa F, Heusch G, and Schulz R. Connexin 43 in cardiomyocyte mitochondria and its increase by ischemic preconditioning. Cardiovascular research 67: 234-244, 2005.

27. Bogoyevitch MA, Parker PJ, and Sugden PH. Characterization of protein kinase C isotype expression in adult rat heart. Protein kinase C-epsilon is a major isotype present, and it is activated by phorbol esters, epinephrine, and endothelin. Circulation research 72: 757-767, 1993.

28. Boivin B, and Allen BG. Regulation of membrane-bound PKC in adult cardiac ventricular myocytes. Cellular signalling 15: 217-224, 2003.

29. Boivin B, Villeneuve LR, Farhat N, Chevalier D, and Allen BG. Sub- cellular distribution of endothelin signaling pathway components in ventricular myocytes and heart: lack of preformed caveolar signalosomes. Journal of molecular and cellular cardiology 38: 665-676, 2005.

30. Bossuyt J, Taylor BE, James-Kracke M, and Hale CC. The cardiac sodium-calcium exchanger associates with caveolin-3. Ann N Y Acad Sci 976: 197-204, 2002.

31. Brainard AM, Miller AJ, Martens JR, and England SK. Maxi-K channels localize to caveolae in human myometrium: a role for an actin- channel-caveolin complex in the regulation of myometrial smooth muscle K+ current. American journal of physiology 289: C49-57, 2005.

32. Brix J, Rudiger S, Bukau B, Schneider-Mergener J, and Pfanner N. Distribution of binding sequences for the mitochondrial import receptors Tom20, 158 Tom22, and Tom70 in a presequence-carrying preprotein and a non-cleavable preprotein. The Journal of biological chemistry 274: 16522-16530, 1999.

33. Budas GR, Churchill EN, and Mochly-Rosen D. Cardioprotective mechanisms of PKC isozyme-selective activators and inhibitors in the treatment of ischemia-reperfusion injury. Pharmacol Res 55: 523-536, 2007.

34. Bugge E, and Ytrehus K. Endothelin-1 can reduce infarct size through protein kinase C and KATP channels in the isolated rat heart. Cardiovascular research 32: 920-929, 1996.

35. Calaghan SC, and Taggart MJ. Compartmentalized signaling in cardiomyocyte lipid domains--do structure and function match up? J Mol Cell Cardiol 41: 1-3, 2006.

36. Cameron JS, Hoffmann KE, Zia C, Hemmett HM, Kronsteiner A, and Lee CM. A role for nitric oxide in hypoxia-induced activation of cardiac KATP channels in goldfish (Carassius auratus). J Exp Biol 206: 4057-4065, 2003.

37. Capozza F, Combs TP, Cohen AW, Cho YR, Park SY, Schubert W, Williams TM, Brasaemle DL, Jelicks LA, Scherer PE, Kim JK, and Lisanti MP. Caveolin-3 knockout mice show increased adiposity and whole body insulin resistance, with ligand-induced insulin receptor instability in skeletal muscle. American journal of physiology 288: C1317-1331, 2005.

38. Capozza F, Williams TM, Schubert W, McClain S, Bouzahzah B, Sotgia F, and Lisanti MP. Absence of caveolin-1 sensitizes mouse skin to carcinogen-induced epidermal hyperplasia and tumor formation. The American journal of pathology 162: 2029-2039, 2003.

39. Carrasco AJ, Dzeja PP, Alekseev AE, Pucar D, Zingman LV, Abraham MR, Hodgson D, Bienengraeber M, Puceat M, Janssen E, Wieringa B, and Terzic A. Adenylate kinase phosphotransfer communicates cellular energetic signals to ATP-sensitive potassium channels. Proc Natl Acad Sci U S A 98: 7623-7628, 2001.

40. Champion HC, Georgakopoulos D, Takimoto E, Isoda T, Wang Y, and Kass DA. Modulation of in vivo cardiac function by myocyte-specific nitric oxide synthase-3. Circulation research 94: 657-663, 2004. 159 41. Chi L, Uprichard AC, and Lucchesi BR. Profibrillatory actions of pinacidil in a conscious canine model of sudden coronary death. Journal of cardiovascular pharmacology 15: 452-464, 1990.

42. Chua BT, Volbracht C, Tan KO, Li R, Yu VC, and Li P. Mitochondrial translocation of cofilin is an early step in apoptosis induction. Nature cell biology 5: 1083-1089, 2003.

43. Chun M, Liyanage UK, Lisanti MP, and Lodish HF. Signal transduction of a G protein-coupled receptor in caveolae: colocalization of endothelin and its receptor with caveolin. Proceedings of the National Academy of Sciences of the United States of America 91: 11728-11732, 1994.

44. Chutkow WA, Samuel V, Hansen PA, Pu J, Valdivia CR, Makielski JC, and Burant CF. Disruption of Sur2-containing K(ATP) channels enhances insulin-stimulated glucose uptake in skeletal muscle. Proceedings of the National Academy of Sciences of the United States of America 98: 11760- 11764, 2001.

45. Coaxum SD, Martin JL, and Mestril R. Overexpression of heat shock proteins differentially modulates protein kinase C expression in rat neonatal cardiomyocytes. Cell stress & chaperones 8: 297-302, 2003.

46. Coetzee WA, Wells T, and Avkiran M. Anti-arrhythmic effects of levcromakalim in the ischaemic rat heart: a dual mechanism of action? European journal of pharmacology 402: 263-274, 2000.

47. Cohen MV, Baines CP, and Downey JM. Ischemic preconditioning: from adenosine receptor to KATP channel. Annual review of physiology 62: 79- 109, 2000.

48. Cohen MV, Yang XM, and Downey JM. Nitric oxide is a preconditioning mimetic and cardioprotectant and is the basis of many available infarct-sparing strategies. Cardiovascular research 70: 231-239, 2006.

49. Costa AD, Garlid KD, West IC, Lincoln TM, Downey JM, Cohen MV, and Critz SD. Protein kinase G transmits the cardioprotective signal from cytosol to mitochondria. Circulation research 97: 329-336, 2005.

160 50. Costa AD, Jakob R, Costa CL, Andrukhiv K, West IC, and Garlid KD. The mechanism by which the mitochondrial ATP-sensitive K+ channel opening and H2O2 inhibit the mitochondrial permeability transition. The Journal of biological chemistry 281: 20801-20808, 2006.

51. Couet J, Li S, Okamoto T, Ikezu T, and Lisanti MP. Identification of peptide and protein ligands for the caveolin-scaffolding domain. Implications for the interaction of caveolin with caveolae-associated proteins. The Journal of biological chemistry 272: 6525-6533, 1997.

52. Crawford RM, Budas GR, Jovanovic S, Ranki HJ, Wilson TJ, Davies AM, and Jovanovic A. M-LDH serves as a sarcolemmal K(ATP) channel subunit essential for cell protection against ischemia. The EMBO journal 21: 3936-3948, 2002.

53. Crawford RM, Ranki HJ, Botting CH, Budas GR, and Jovanovic A. Creatine kinase is physically associated with the cardiac ATP-sensitive K+ channel in vivo. Faseb J 16: 102-104, 2002.

54. Cuong DV, Kim N, Joo H, Youm JB, Chung JY, Lee Y, Park WS, Kim E, Park YS, and Han J. Subunit composition of ATP-sensitive potassium channels in mitochondria of rat hearts. Mitochondrion 5: 121-133, 2005.

55. D'Hahan N, Moreau C, Prost AL, Jacquet H, Alekseev AE, Terzic A, and Vivaudou M. Pharmacological plasticity of cardiac ATP-sensitive potassium channels toward diazoxide revealed by ADP. Proceedings of the National Academy of Sciences of the United States of America 96: 12162- 12167, 1999.

56. Dabrowski M, Tarasov A, and Ashcroft FM. Mapping the architecture of the ATP-binding site of the KATP channel subunit Kir6.2. The Journal of physiology 557: 347-354, 2004.

57. Dascal N. Ion-channel regulation by G proteins. Trends in endocrinology and metabolism: TEM 12: 391-398, 2001.

58. Davis AJ, Ryan KR, and Jensen RE. Tim23p contains separate and distinct signals for targeting to mitochondria and insertion into the inner membrane. Molecular biology of the cell 9: 2577-2593, 1998. 161 59. Dawn B, and Bolli R. Role of nitric oxide in myocardial preconditioning. Annals of the New York Academy of Sciences 962: 18-41, 2002.

60. de Jong JW, de Jonge R, Keijzer E, and Bradamante S. The role of adenosine in preconditioning. Pharmacology & therapeutics 87: 141-149, 2000.

61. de Weerd WF, and Leeb-Lundberg LM. Bradykinin sequesters B2 bradykinin receptors and the receptor-coupled Galpha subunits Galphaq and Galphai in caveolae in DDT1 MF-2 smooth muscle cells. The Journal of biological chemistry 272: 17858-17866, 1997.

62. del Pozo MA, Balasubramanian N, Alderson NB, Kiosses WB, Grande-Garcia A, Anderson RG, and Schwartz MA. Phospho-caveolin-1 mediates integrin-regulated membrane domain internalization. Nature cell biology 7: 901-908, 2005.

63. Dermietzel R, Hwang TK, Buettner R, Hofer A, Dotzler E, Kremer M, Deutzmann R, Thinnes FP, Fishman GI, Spray DC, and et al. Cloning and in situ localization of a brain-derived porin that constitutes a large-conductance anion channel in astrocytic plasma membranes. Proceedings of the National Academy of Sciences of the United States of America 91: 499-503, 1994.

64. Dhar-Chowdhury P, Harrell MD, Han SY, Jankowska D, Parachuru L, Morrissey A, Srivastava S, Liu W, Malester B, Yoshida H, and Coetzee WA. The glycolytic enzymes, glyceraldehyde-3-phosphate dehydrogenase, triose- phosphate isomerase, and pyruvate kinase are components of the K(ATP) channel macromolecular complex and regulate its function. The Journal of biological chemistry 280: 38464-38470, 2005.

65. Dhar-Chowdhury P, Malester B, Rajacic P, and Coetzee WA. The regulation of ion channels and transporters by glycolytically derived ATP. Cell Mol Life Sci 64: 3069-3083, 2007.

66. Disatnik MH, Buraggi G, and Mochly-Rosen D. Localization of protein kinase C isozymes in cardiac myocytes. Experimental cell research 210: 287- 297, 1994.

67. Driamov S, Bellahcene M, Ziegler A, Barbosa V, Traub D, Butz S, Buser PT, and Zaugg CE. Antiarrhythmic effect of ischemic preconditioning 162 during low-flow ischemia. The role of bradykinin and sarcolemmal versus mitochondrial ATP-sensitive K(+) channels. Basic research in cardiology 99: 299-308, 2004.

68. Dzeja PP, Bast P, Ozcan C, Valverde A, Holmuhamedov EL, Van Wylen DG, and Terzic A. Targeting nucleotide-requiring enzymes: implications for diazoxide-induced cardioprotection. Am J Physiol Heart Circ Physiol 284: H1048-1056, 2003.

69. Elliott AC, Smith GL, and Allen DG. Simultaneous measurements of action potential duration and intracellular ATP in isolated ferret hearts exposed to cyanide. Circulation research 64: 583-591, 1989.

70. Ellisman MH, and Levinson SR. Immunocytochemical localization of sodium channel distributions in the excitable membranes of Electrophorus electricus. Proceedings of the National Academy of Sciences of the United States of America 79: 6707-6711, 1982.

71. Emanuelsson O, Nielsen H, Brunak S, and von Heijne G. Predicting subcellular localization of proteins based on their N-terminal amino acid sequence. Journal of molecular biology 300: 1005-1016, 2000.

72. Fan Z, and Makielski JC. Anionic phospholipids activate ATP-sensitive potassium channels. The Journal of biological chemistry 272: 5388-5395, 1997.

73. Fecchi K, Volonte D, Hezel MP, Schmeck K, and Galbiati F. Spatial and temporal regulation of GLUT4 translocation by flotillin-1 and caveolin-3 in skeletal muscle cells. Faseb J 20: 705-707, 2006.

74. Fernandez MA, Albor C, Ingelmo-Torres M, Nixon SJ, Ferguson C, Kurzchalia T, Tebar F, Enrich C, Parton RG, and Pol A. Caveolin-1 is essential for liver regeneration. Science (New York, NY 313: 1628-1632, 2006.

75. Feron O, Smith TW, Michel T, and Kelly RA. Dynamic targeting of the agonist-stimulated m2 muscarinic acetylcholine receptor to caveolae in cardiac myocytes. The Journal of biological chemistry 272: 17744-17748, 1997.

163 76. Findlay I. ATP4- and ATP.Mg inhibit the ATP-sensitive K+ channel of rat ventricular myocytes. Pflugers Arch 412: 37-41, 1988.

77. Flagg TP, Kurata HT, Masia R, Caputa G, Magnuson MA, Lefer DJ, Coetzee WA, and Nichols CG. Differential structure of atrial and ventricular KATP: atrial KATP channels require SUR1. Circulation research 103: 1458- 1465, 2008.

78. Foster DB, Rucker JJ, and Marban E. Is Kir6.1 a subunit of mitoK(ATP)? Biochemical and biophysical research communications 366: 649- 656, 2008.

79. Fredholm BB, AP IJ, Jacobson KA, Klotz KN, and Linden J. International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacological reviews 53: 527-552, 2001.

80. Fu Y, Hoang A, Escher G, Parton RG, Krozowski Z, and Sviridov D. Expression of caveolin-1 enhances cholesterol efflux in hepatic cells. The Journal of biological chemistry 279: 14140-14146, 2004.

81. Furukawa T, Kimura S, Furukawa N, Bassett AL, and Myerburg RJ. Role of cardiac ATP-regulated potassium channels in differential responses of endocardial and epicardial cells to ischemia. Circulation research 68: 1693- 1702, 1991.

82. Gabella G. Inpocketings of the cell membrane (caveolae) in the rat myocardium. J Ultrastruct Res 65: 135-147, 1978.

83. Galbiati F, Engelman JA, Volonte D, Zhang XL, Minetti C, Li M, Hou H, Jr., Kneitz B, Edelmann W, and Lisanti MP. Caveolin-3 null mice show a loss of caveolae, changes in the microdomain distribution of the dystrophin- glycoprotein complex, and t-tubule abnormalities. The Journal of biological chemistry 276: 21425-21433, 2001.

84. Garg V, and Hu K. Protein kinase C isoform-dependent modulation of ATP-sensitive K+ channels in mitochondrial inner membrane. Am J Physiol Heart Circ Physiol 293: H322-332, 2007.

164 85. Garg V, Jiao J, and Hu K. Regulation of ATP-sensitive K+ channels by caveolin-enriched microdomains in cardiac myocytes. Cardiovascular research 82: 51-58, 2009.

86. Garlid KD, Paucek P, Yarov-Yarovoy V, Murray HN, Darbenzio RB, D'Alonzo AJ, Lodge NJ, Smith MA, and Grover GJ. Cardioprotective effect of diazoxide and its interaction with mitochondrial ATP-sensitive K+ channels. Possible mechanism of cardioprotection. Circulation research 81: 1072-1082, 1997.

87. Garlid KD, Paucek P, Yarov-Yarovoy V, Sun X, and Schindler PA. The mitochondrial KATP channel as a receptor for potassium channel openers. The Journal of biological chemistry 271: 8796-8799, 1996.

88. Ge ZD, Peart JN, Kreckler LM, Wan TC, Jacobson MA, Gross GJ, and Auchampach JA. Cl-IB-MECA [2-chloro-N6-(3-iodobenzyl)adenosine-5'- N-methylcarboxamide] reduces ischemia/reperfusion injury in mice by activating the A3 adenosine receptor. The Journal of pharmacology and experimental therapeutics 319: 1200-1210, 2006.

89. Glenney JR, Jr. Tyrosine phosphorylation of a 22-kDa protein is correlated with transformation by Rous sarcoma virus. The Journal of biological chemistry 264: 20163-20166, 1989.

90. Gordon GW, Berry G, Liang XH, Levine B, and Herman B. Quantitative fluorescence resonance energy transfer measurements using fluorescence microscopy. Biophysical journal 74: 2702-2713, 1998.

91. Green DE, Murer E, Hultin HO, Richardson SH, Salmon B, Brierley GP, and Baum H. Association of integrated metabolic pathways with membranes. I. Glycolytic enzymes of the red blood corpuscle and yeast. Archives of biochemistry and biophysics 112: 635-647, 1965.

92. Green DR, and Reed JC. Mitochondria and apoptosis. Science (New York, NY 281: 1309-1312, 1998.

93. Gross GJ. ATP-sensitive potassium channels and myocardial preconditioning. Basic Res Cardiol 90: 85-88, 1995.

165 94. Gross GJ, and Auchampach JA. Blockade of ATP-sensitive potassium channels prevents myocardial preconditioning in dogs. Circulation research 70: 223-233, 1992.

95. Gross GJ, and Fryer RM. Sarcolemmal versus mitochondrial ATP- sensitive K+ channels and myocardial preconditioning. Circulation research 84: 973-979, 1999.

96. Gross GJ, and Peart JN. KATP channels and myocardial preconditioning: an update. Am J Physiol Heart Circ Physiol 285: H921-930, 2003.

97. Grover GJ, and Garlid KD. ATP-Sensitive potassium channels: a review of their cardioprotective pharmacology. Journal of molecular and cellular cardiology 32: 677-695, 2000.

98. Grover GJ, Sleph PG, and Dzwonczyk S. Role of myocardial ATP- sensitive potassium channels in mediating preconditioning in the dog heart and their possible interaction with adenosine A1-receptors. Circulation 86: 1310- 1316, 1992.

99. Gumina RJ, Pucar D, Bast P, Hodgson DM, Kurtz CE, Dzeja PP, Miki T, Seino S, and Terzic A. Knockout of Kir6.2 negates ischemic preconditioning-induced protection of myocardial energetics. Am J Physiol Heart Circ Physiol 284: H2106-2113, 2003.

100. Guo Y, Bolli R, Bao W, Wu WJ, Black RG, Jr., Murphree SS, Salvatore CA, Jacobson MA, and Auchampach JA. Targeted deletion of the A3 adenosine receptor confers resistance to myocardial ischemic injury and does not prevent early preconditioning. Journal of molecular and cellular cardiology 33: 825-830, 2001.

101. Hagiwara Y, Sasaoka T, Araishi K, Imamura M, Yorifuji H, Nonaka I, Ozawa E, and Kikuchi T. Caveolin-3 deficiency causes muscle degeneration in mice. Human molecular genetics 9: 3047-3054, 2000.

102. Hailstones D, Sleer LS, Parton RG, and Stanley KK. Regulation of caveolin and caveolae by cholesterol in MDCK cells. Journal of lipid research 39: 369-379, 1998. 166 103. Hanley PJ, Mickel M, Loffler M, Brandt U, and Daut J. K(ATP) channel-independent targets of diazoxide and 5-hydroxydecanoate in the heart. The Journal of physiology 542: 735-741, 2002.

104. Harrell MD, Harbi S, Hoffman JF, Zavadil J, and Coetzee WA. Large- scale analysis of ion channel gene expression in the mouse heart during perinatal development. Physiological genomics 28: 273-283, 2007.

105. Harrison GJ, Cerniway RJ, Peart J, Berr SS, Ashton K, Regan S, Paul Matherne G, and Headrick JP. Effects of A(3) adenosine receptor activation and gene knock-out in ischemic-reperfused mouse heart. Cardiovascular research 53: 147-155, 2002.

106. Haruna T, Horie M, Kouchi I, Nawada R, Tsuchiya K, Akao M, Otani H, Murakami T, and Sasayama S. Coordinate interaction between ATP- sensitive K+ channel and Na+, K+-ATPase modulates ischemic preconditioning. Circulation 98: 2905-2910, 1998.

107. Head BP, Patel HH, Roth DM, Lai NC, Niesman IR, Farquhar MG, and Insel PA. G-protein-coupled receptor signaling components localize in both sarcolemmal and intracellular caveolin-3-associated microdomains in adult cardiac myocytes. The Journal of biological chemistry 280: 31036-31044, 2005.

108. Head BP, Patel HH, Roth DM, Murray F, Swaney JS, Niesman IR, Farquhar MG, and Insel PA. Microtubules and actin microfilaments regulate lipid raft/caveolae localization of adenylyl cyclase signaling components. The Journal of biological chemistry 281: 26391-26399, 2006.

109. Hide EJ, Piper J, and Thiemermann C. Endothelin-1-induced reduction of myocardial infarct size by activation of ATP-sensitive potassium channels in a rabbit model of myocardial ischaemia and reperfusion. British journal of pharmacology 116: 2597-2602, 1995.

110. Hilgemann DW, and Ball R. Regulation of cardiac Na+,Ca2+ exchange and KATP potassium channels by PIP2. Science (New York, NY 273: 956-959, 1996.

111. Hilgemann DW, Feng S, and Nasuhoglu C. The complex and intriguing lives of PIP2 with ion channels and transporters. Sci STKE 2001: RE19, 2001. 167 112. Hu K, Duan D, Li GR, and Nattel S. Protein kinase C activates ATP- sensitive K+ current in human and rabbit ventricular myocytes. Circ Res 78: 492-498, 1996.

113. Hu K, Huang CS, Jan YN, and Jan LY. ATP-sensitive potassium channel traffic regulation by adenosine and protein kinase C. Neuron 38: 417- 432, 2003.

114. Hu K, Li GR, and Nattel S. Adenosine-induced activation of ATP- sensitive K+ channels in excised membrane patches is mediated by PKC. Am J Physiol 276: H488-495, 1999.

115. Hu K, Mochly-Rosen D, and Boutjdir M. Evidence for functional role of epsilonPKC isozyme in the regulation of cardiac Ca(2+) channels. Am J Physiol Heart Circ Physiol 279: H2658-2664, 2000.

116. Hu K, and Nattel S. Mechanisms of ischemic preconditioning in rat hearts. Involvement of alpha 1B-adrenoceptors, pertussis toxin-sensitive G proteins, and protein kinase C. Circulation 92: 2259-2265, 1995.

117. Huang X, and Walker JW. Myofilament anchoring of protein kinase C- epsilon in cardiac myocytes. Journal of cell science 117: 1971-1978, 2004.

118. Huang XP, Pi Y, Lokuta AJ, Greaser ML, and Walker JW. Arachidonic acid stimulates protein kinase C-epsilon redistribution in heart cells. Journal of cell science 110 ( Pt 14): 1625-1634, 1997.

119. Inoue I, Nagase H, Kishi K, and Higuti T. ATP-sensitive K+ channel in the mitochondrial inner membrane. Nature 352: 244-247, 1991.

120. Insel PA, Head BP, Ostrom RS, Patel HH, Swaney JS, Tang CM, and Roth DM. Caveolae and lipid rafts: G protein-coupled receptor signaling microdomains in cardiac myocytes. Annals of the New York Academy of Sciences 1047: 166-172, 2005.

121. Ishitsuka R, Sato SB, and Kobayashi T. Imaging lipid rafts. Journal of biochemistry 137: 249-254, 2005.

168 122. Ito H, Sugimoto T, Kobayashi I, Takahashi K, Katada T, Ui M, and Kurachi Y. On the mechanism of basal and agonist-induced activation of the G protein-gated muscarinic K+ channel in atrial myocytes of guinea pig heart. The Journal of general physiology 98: 517-533, 1991.

123. Ito H, Tung RT, Sugimoto T, Kobayashi I, Takahashi K, Katada T, Ui M, and Kurachi Y. On the mechanism of G protein beta gamma subunit activation of the muscarinic K+ channel in guinea pig atrial cell membrane. Comparison with the ATP-sensitive K+ channel. The Journal of general physiology 99: 961-983, 1992.

124. Ito H, Vereecke J, and Carmeliet E. Mode of regulation by G protein of the ATP-sensitive K+ channel in guinea-pig ventricular cell membrane. The Journal of physiology 478 ( Pt 1): 101-107, 1994.

125. Jaburek M, Costa AD, Burton JR, Costa CL, and Garlid KD. Mitochondrial PKC epsilon and mitochondrial ATP-sensitive K+ channel copurify and coreconstitute to form a functioning signaling module in proteoliposomes. Circulation research 99: 878-883, 2006.

126. Janse MJ, and Wit AL. Electrophysiological mechanisms of ventricular arrhythmias resulting from myocardial ischemia and infarction. Physiological reviews 69: 1049-1169, 1989.

127. Jiang MT, Ljubkovic M, Nakae Y, Shi Y, Kwok WM, Stowe DF, and Bosnjak ZJ. Characterization of human cardiac mitochondrial ATP-sensitive potassium channel and its regulation by phorbol ester in vitro. Am J Physiol Heart Circ Physiol 290: H1770-1776, 2006.

128. Jiao J, Garg V, Yang B, Elton TS, and Hu K. Protein kinase C-epsilon induces caveolin-dependent internalization of vascular adenosine 5'- triphosphate-sensitive K+ channels. Hypertension 52: 499-506, 2008.

129. Jones DP. Intracellular diffusion gradients of O2 and ATP. The American journal of physiology 250: C663-675, 1986.

130. Jorgensen AO, Shen AC, Arnold W, Leung AT, and Campbell KP. Subcellular distribution of the 1,4-dihydropyridine receptor in rabbit skeletal

169 muscle in situ: an immunofluorescence and immunocolloidal gold-labeling study. The Journal of cell biology 109: 135-147, 1989.

131. Jovanovic A, Jovanovic S, Lorenz E, and Terzic A. Recombinant cardiac ATP-sensitive K+ channel subunits confer resistance to chemical hypoxia-reoxygenation injury. Circulation 98: 1548-1555, 1998.

132. Ju H, Venema VJ, Liang H, Harris MB, Zou R, and Venema RC. Bradykinin activates the Janus-activated kinase/signal transducers and activators of transcription (JAK/STAT) pathway in vascular endothelial cells: localization of JAK/STAT signalling proteins in plasmalemmal caveolae. The Biochemical journal 351: 257-264, 2000.

133. Juhaszova M, Zorov DB, Kim SH, Pepe S, Fu Q, Fishbein KW, Ziman BD, Wang S, Ytrehus K, Antos CL, Olson EN, and Sollott SJ. Glycogen synthase kinase-3beta mediates convergence of protection signaling to inhibit the mitochondrial permeability transition pore. The Journal of clinical investigation 113: 1535-1549, 2004.

134. Kaab S, Zwermann L, Barth A, Hinterseer M, Englert HC, Gogelein H, and Nabauer M. Selective block of sarcolemmal IKATP in human cardiomyocytes using HMR 1098. Cardiovascular drugs and therapy / sponsored by the International Society of Cardiovascular Pharmacotherapy 17: 435-441, 2003.

135. Kabakov AY. Activation of KATP channels by Na/K pump in isolated cardiac myocytes and giant membrane patches. Biophysical journal 75: 2858- 2867, 1998.

136. Kakei M, Noma A, and Shibasaki T. Properties of adenosine- triphosphate-regulated potassium channels in guinea-pig ventricular cells. J Physiol 363: 441-462, 1985.

137. Kaldi K, and Neupert W. Protein translocation into mitochondria. BioFactors (Oxford, England) 8: 221-224, 1998.

138. Kamishima T, Burdyga T, Gallagher JA, and Quayle JM. Caveolin-1 and caveolin-3 regulate Ca2+ homeostasis of single smooth muscle cells from

170 rat cerebral resistance arteries. Am J Physiol Heart Circ Physiol 293: H204-214, 2007.

139. Kane GC, Liu XK, Yamada S, Olson TM, and Terzic A. Cardiac KATP channels in health and disease. Journal of molecular and cellular cardiology 38: 937-943, 2005.

140. Kendrick-Jones J, and Perry SV. The enzymes of adenine nucleotide metabolism in developing skeletal muscle. The Biochemical journal 103: 207- 214, 1967.

141. Kim E, Han J, Ho W, and Earm YE. Modulation of ATP-sensitive K+ channels in rabbit ventricular myocytes by adenosine A1 receptor activation. The American journal of physiology 272: H325-333, 1997.

142. Kirkham M, and Parton RG. Clathrin-independent endocytosis: new insights into caveolae and non-caveolar lipid raft carriers. Biochimica et biophysica acta 1745: 273-286, 2005.

143. Kirsch GE, Codina J, Birnbaumer L, and Brown AM. Coupling of ATP-sensitive K+ channels to A1 receptors by G proteins in rat ventricular myocytes. The American journal of physiology 259: H820-826, 1990.

144. Koleske AJ, Baltimore D, and Lisanti MP. Reduction of caveolin and caveolae in oncogenically transformed cells. Proceedings of the National Academy of Sciences of the United States of America 92: 1381-1385, 1995.

145. Korge P, and Campbell KB. The importance of ATPase microenvironment in muscle fatigue: a hypothesis. International journal of sports medicine 16: 172-179, 1995.

146. Kowaltowski AJ, Seetharaman S, Paucek P, and Garlid KD. Bioenergetic consequences of opening the ATP-sensitive K(+) channel of heart mitochondria. Am J Physiol Heart Circ Physiol 280: H649-657, 2001.

147. Lacza Z, Snipes JA, Miller AW, Szabo C, Grover G, and Busija DW. Heart mitochondria contain functional ATP-dependent K+ channels. Journal of molecular and cellular cardiology 35: 1339-1347, 2003.

171 148. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227: 680-685, 1970.

149. Lamping KA, Christensen CW, Pelc LR, Warltier DC, and Gross GJ. Effects of and on protection of ischemic myocardium. Journal of cardiovascular pharmacology 6: 536-542, 1984.

150. Lamping KA, and Gross GJ. Improved recovery of myocardial segment function following a short coronary occlusion in dogs by nicorandil, a potential new antianginal agent, and nifedipine. Journal of cardiovascular pharmacology 7: 158-166, 1985.

151. Lankford AR, Yang JN, Rose'Meyer R, French BA, Matherne GP, Fredholm BB, and Yang Z. Effect of modulating cardiac A1 adenosine receptor expression on protection with ischemic preconditioning. American journal of physiology 290: H1469-1473, 2006.

152. Lanzafame AA, Turnbull L, Amiramahdi F, Arthur JF, Huynh H, and Woodcock EA. Inositol phospholipids localized to caveolae in rat heart are regulated by alpha1-adrenergic receptors and by ischemia-reperfusion. Am J Physiol Heart Circ Physiol 290: H2059-2065, 2006.

153. Lasley RD, and Smart EJ. Cardiac myocyte adenosine receptors and caveolae. Trends in cardiovascular medicine 11: 259-263, 2001.

154. Latchman DS. Heat shock proteins and cardiac protection. Cardiovascular research 51: 637-646, 2001.

155. Leblanc N, and Hume JR. Sodium current-induced release of calcium from cardiac sarcoplasmic reticulum. Science (New York, NY 248: 372-376, 1990.

156. Lederer WJ, Niggli E, and Hadley RW. Sodium-calcium exchange in excitable cells: fuzzy space. Science 248: 283, 1990.

157. Levin KR, and Page E. Quantitative studies on plasmalemmal folds and caveolae of rabbit ventricular myocardial cells. Circulation research 46: 244- 255, 1980.

172 158. Li H, Brodsky S, Kumari S, Valiunas V, Brink P, Kaide J, Nasjletti A, and Goligorsky MS. Paradoxical overexpression and translocation of connexin43 in homocysteine-treated endothelial cells. Am J Physiol Heart Circ Physiol 282: H2124-2133, 2002.

159. Li RA, Leppo M, Miki T, Seino S, and Marban E. Molecular basis of electrocardiographic ST-segment elevation. Circulation research 87: 837-839, 2000.

160. Li S, Okamoto T, Chun M, Sargiacomo M, Casanova JE, Hansen SH, Nishimoto I, and Lisanti MP. Evidence for a regulated interaction between heterotrimeric G proteins and caveolin. The Journal of biological chemistry 270: 15693-15701, 1995.

161. Li WP, Liu P, Pilcher BK, and Anderson RG. Cell-specific targeting of caveolin-1 to caveolae, secretory vesicles, cytoplasm or mitochondria. Journal of cell science 114: 1397-1408, 2001.

162. Liang BT. Direct preconditioning of cardiac ventricular myocytes via adenosine A1 receptor and KATP channel. The American journal of physiology 271: H1769-1777, 1996.

163. Light PE, Bladen C, Winkfein RJ, Walsh MP, and French RJ. Molecular basis of protein kinase C-induced activation of ATP-sensitive potassium channels. Proceedings of the National Academy of Sciences of the United States of America 97: 9058-9063, 2000.

164. Lin YF, Jan YN, and Jan LY. Regulation of ATP-sensitive potassium channel function by protein kinase A-mediated phosphorylation in transfected HEK293 cells. Embo J 19: 942-955, 2000.

165. Lisanti MP, Scherer PE, Tang Z, and Sargiacomo M. Caveolae, caveolin and caveolin-rich membrane domains: a signalling hypothesis. Trends in cell biology 4: 231-235, 1994.

166. Lithgow T. Targeting of proteins to mitochondria. FEBS letters 476: 22- 26, 2000.

173 167. Liu L, and Askari A. Beta-subunit of cardiac Na+-K+-ATPase dictates the concentration of the functional enzyme in caveolae. Am J Physiol Cell Physiol 291: C569-578, 2006.

168. Liu XK, Yamada S, Kane GC, Alekseev AE, Hodgson DM, O'Cochlain F, Jahangir A, Miki T, Seino S, and Terzic A. Genetic disruption of Kir6.2, the pore-forming subunit of ATP-sensitive K+ channel, predisposes to catecholamine-induced ventricular dysrhythmia. Diabetes 53 Suppl 3: S165- 168, 2004.

169. Liu Y, and Downey JM. Ischemic preconditioning protects against infarction in rat heart. The American journal of physiology 263: H1107-1112, 1992.

170. Liu Y, Gao WD, O'Rourke B, and Marban E. Synergistic modulation of ATP-sensitive K+ currents by protein kinase C and adenosine. Implications for ischemic preconditioning. Circulation research 78: 443-454, 1996.

171. Liu Y, Sato T, O'Rourke B, and Marban E. Mitochondrial ATP- dependent potassium channels: novel effectors of cardioprotection? Circulation 97: 2463-2469, 1998.

172. Lohn M, Furstenau M, Sagach V, Elger M, Schulze W, Luft FC, Haller H, and Gollasch M. Ignition of calcium sparks in arterial and cardiac muscle through caveolae. Circ Res 87: 1034-1039, 2000.

173. Lu L, Reiter MJ, Xu Y, Chicco A, Greyson CR, and Schwartz GG. Thiazolidinedione drugs block cardiac KATP channels and may increase propensity for ischaemic ventricular fibrillation in pigs. Diabetologia 51: 675- 685, 2008.

174. Lynch RM, and Paul RJ. Compartmentation of glycolytic and glycogenolytic metabolism in vascular smooth muscle. Science (New York, NY 222: 1344-1346, 1983.

175. Maguy A, Hebert TE, and Nattel S. Involvement of lipid rafts and caveolae in cardiac ion channel function. Cardiovascular research 69: 798-807, 2006.

174 176. Mannhold R. KATP channel openers: structure-activity relationships and therapeutic potential. Medicinal research reviews 24: 213-266, 2004.

177. Martens JR, O'Connell K, and Tamkun M. Targeting of ion channels to membrane microdomains: localization of KV channels to lipid rafts. Trends in pharmacological sciences 25: 16-21, 2004.

178. Martens JR, Sakamoto N, Sullivan SA, Grobaski TD, and Tamkun MM. Isoform-specific localization of voltage-gated K+ channels to distinct lipid raft populations. Targeting of Kv1.5 to caveolae. The Journal of biological chemistry 276: 8409-8414, 2001.

179. Matherne GP, Linden J, Byford AM, Gauthier NS, and Headrick JP. Transgenic A1 adenosine receptor overexpression increases myocardial resistance to ischemia. Proceedings of the National Academy of Sciences of the United States of America 94: 6541-6546, 1997.

180. McEwen DP, Li Q, Jackson S, Jenkins PM, and Martens JR. Caveolin regulates kv1.5 trafficking to cholesterol-rich membrane microdomains. Molecular pharmacology 73: 678-685, 2008.

181. Meshulam T, Simard JR, Wharton J, Hamilton JA, and Pilch PF. Role of caveolin-1 and cholesterol in transmembrane fatty acid movement. Biochemistry 45: 2882-2893, 2006.

182. Michailova A, Lorentz W, and McCulloch A. Modeling transmural heterogeneity of K(ATP) current in rabbit ventricular myocytes. American journal of physiology 293: C542-557, 2007.

183. Michailova A, Saucerman J, Belik ME, and McCulloch AD. Modeling regulation of cardiac KATP and L-type Ca2+ currents by ATP, ADP, and Mg2+. Biophysical journal 88: 2234-2249, 2005.

184. Miki T, Suzuki M, Shibasaki T, Uemura H, Sato T, Yamaguchi K, Koseki H, Iwanaga T, Nakaya H, and Seino S. Mouse model of Prinzmetal angina by disruption of the inward rectifier Kir6.1. Nature medicine 8: 466-472, 2002.

175 185. Mineo C, Ying YS, Chapline C, Jaken S, and Anderson RG. Targeting of protein kinase Calpha to caveolae. The Journal of cell biology 141: 601-610, 1998.

186. Moncada GA, Kishi Y, Numano F, Hiraoka M, and Sawanobori T. Effects of acidosis and NO on nicorandil-activated K(ATP) channels in guinea- pig ventricular myocytes. British journal of pharmacology 131: 1097-1104, 2000.

187. Morrissey A, Rosner E, Lanning J, Parachuru L, Dhar Chowdhury P, Han S, Lopez G, Tong X, Yoshida H, Nakamura TY, Artman M, Giblin JP, Tinker A, and Coetzee WA. Immunolocalization of KATP channel subunits in mouse and rat cardiac myocytes and the coronary vasculature. BMC physiology 5: 1, 2005.

188. Mubagwa K, and Flameng W. Adenosine, adenosine receptors and myocardial protection: an updated overview. Cardiovascular research 52: 25- 39, 2001.

189. Murry CE, Jennings RB, and Reimer KA. Preconditioning with ischemia: a delay of lethal cell injury in ischemic myocardium. Circulation 74: 1124-1136, 1986.

190. Nattel S, Li D, and Yue L. Basic mechanisms of atrial fibrillation--very new insights into very old ideas. Annual review of physiology 62: 51-77, 2000.

191. Nevins AK, and Thurmond DC. Caveolin-1 functions as a novel Cdc42 nucleotide dissociation inhibitor in pancreatic beta-cells. The Journal of biological chemistry 281: 18961-18972, 2006.

192. Nichols CG. KATP channels as molecular sensors of cellular metabolism. Nature 440: 470-476, 2006.

193. Nichols CG, and Lederer WJ. Adenosine triphosphate-sensitive potassium channels in the cardiovascular system. Am J Physiol 261: H1675- 1686, 1991.

176 194. Nichols CG, and Lederer WJ. The regulation of ATP-sensitive K+ channel activity in intact and permeabilized rat ventricular myocytes. J Physiol 423: 91-110, 1990.

195. Nichols CG, Ripoll C, and Lederer WJ. ATP-sensitive potassium channel modulation of the guinea pig ventricular action potential and contraction. Circulation research 68: 280-287, 1991.

196. Ningaraj NS, Rao M, Hashizume K, Asotra K, and Black KL. Regulation of blood-brain tumor barrier permeability by calcium-activated potassium channels. J Pharmacol Exp Ther 301: 838-851, 2002.

197. Noma A. ATP-regulated K+ channels in cardiac muscle. Nature 305: 147-148, 1983.

198. O'Rourke B. Evidence for mitochondrial K+ channels and their role in cardioprotection. Circulation research 94: 420-432, 2004.

199. O'Rourke B. Myocardial K(ATP) channels in preconditioning. Circulation research 87: 845-855, 2000.

200. Oh YS, Cho KA, Ryu SJ, Khil LY, Jun HS, Yoon JW, and Park SC. Regulation of insulin response in skeletal muscle cell by caveolin status. Journal of cellular biochemistry 99: 747-758, 2006.

201. Ohnuma Y, Miura T, Miki T, Tanno M, Kuno A, Tsuchida A, and Shimamoto K. Opening of mitochondrial K(ATP) channel occurs downstream of PKC-epsilon activation in the mechanism of preconditioning. Am J Physiol Heart Circ Physiol 283: H440-447, 2002.

202. Okamoto T, Schlegel A, Scherer PE, and Lisanti MP. Caveolins, a family of scaffolding proteins for organizing "preassembled signaling complexes" at the plasma membrane. The Journal of biological chemistry 273: 5419-5422, 1998.

203. Okuyama Y, Yamada M, Kondo C, Satoh E, Isomoto S, Shindo T, Horio Y, Kitakaze M, Hori M, and Kurachi Y. The effects of nucleotides and potassium channel openers on the SUR2A/Kir6.2 complex K+ channel

177 expressed in a mammalian cell line, HEK293T cells. Pflugers Arch 435: 595- 603, 1998.

204. Orlichenko L, Huang B, Krueger E, and McNiven MA. Epithelial growth factor-induced phosphorylation of caveolin 1 at tyrosine 14 stimulates caveolae formation in epithelial cells. The Journal of biological chemistry 281: 4570-4579, 2006.

205. Ostrom RS, Gregorian C, Drenan RM, Xiang Y, Regan JW, and Insel PA. Receptor number and caveolar co-localization determine receptor coupling efficiency to adenylyl cyclase. The Journal of biological chemistry 276: 42063- 42069, 2001.

206. Ostrom RS, Violin JD, Coleman S, and Insel PA. Selective enhancement of beta-adrenergic receptor signaling by overexpression of adenylyl cyclase type 6: colocalization of receptor and adenylyl cyclase in caveolae of cardiac myocytes. Molecular pharmacology 57: 1075-1079, 2000.

207. Otani H. Ischemic preconditioning: from molecular mechanisms to therapeutic opportunities. Antioxidants & redox signaling 10: 207-247, 2008.

208. Page E, McCallister LP, and Power B. Sterological measurements of cardiac ultrastructures implicated in excitation-contraction coupling. Proceedings of the National Academy of Sciences of the United States of America 68: 1465-1466, 1971.

209. Palmer JW, Tandler B, and Hoppel CL. Biochemical differences between subsarcolemmal and interfibrillar mitochondria from rat cardiac muscle: effects of procedural manipulations. Archives of biochemistry and biophysics 236: 691-702, 1985.

210. Parton RG, and Simons K. The multiple faces of caveolae. Nature reviews 8: 185-194, 2007.

211. Parton RG, Way M, Zorzi N, and Stang E. Caveolin-3 associates with developing T-tubules during muscle differentiation. J Cell Biol 136: 137-154, 1997.

178 212. Patel HH, Gross ER, Peart JN, Hsu AK, and Gross GJ. Sarcolemmal KATP channel triggers delayed ischemic preconditioning in rats. Am J Physiol Heart Circ Physiol 288: H445-447, 2005.

213. Patel HH, Head BP, Petersen HN, Niesman IR, Huang D, Gross GJ, Insel PA, and Roth DM. Protection of adult rat cardiac myocytes from ischemic cell death: role of caveolar microdomains and delta-opioid receptors. Am J Physiol Heart Circ Physiol 291: H344-350, 2006.

214. Patel HH, Murray F, and Insel PA. Caveolae as organizers of pharmacologically relevant signal transduction molecules. Annual review of pharmacology and toxicology 48: 359-391, 2008.

215. Patel HH, Tsutsumi YM, Head BP, Niesman IR, Jennings M, Horikawa Y, Huang D, Moreno AL, Patel PM, Insel PA, and Roth DM. Mechanisms of cardiac protection from ischemia/reperfusion injury: a role for caveolae and caveolin-1. Faseb J 2007.

216. Paucek P, Mironova G, Mahdi F, Beavis AD, Woldegiorgis G, and Garlid KD. Reconstitution and partial purification of the glibenclamide-sensitive, ATP-dependent K+ channel from rat liver and beef heart mitochondria. The Journal of biological chemistry 267: 26062-26069, 1992.

217. Pelkmans L, Puntener D, and Helenius A. Local actin polymerization and dynamin recruitment in SV40-induced internalization of caveolae. Science (New York, NY 296: 535-539, 2002.

218. Penna C, Mancardi D, Rastaldo R, Losano G, and Pagliaro P. Intermittent activation of bradykinin B2 receptors and mitochondrial KATP channels trigger cardiac postconditioning through redox signaling. Cardiovascular research 75: 168-177, 2007.

219. Periasamy A, and Diaspro A. Multiphoton microscopy. Journal of biomedical optics 8: 327-328, 2003.

220. Pfanner N, and Meijer M. The Tom and Tim machine. Curr Biol 7: R100-103, 1997.

179 221. Philip-Couderc P, Tavares NI, Roatti A, Lerch R, Montessuit C, and Baertschi AJ. Forkhead transcription factors coordinate expression of myocardial KATP channel subunits and energy metabolism. Circulation research 102: e20-35, 2008.

222. Pierce GN, and Philipson KD. Binding of glycolytic enzymes to cardiac sarcolemmal and sarcoplasmic reticular membranes. The Journal of biological chemistry 260: 6862-6870, 1985.

223. Ping P, Zhang J, Pierce WM, Jr., and Bolli R. Functional proteomic analysis of protein kinase C epsilon signaling complexes in the normal heart and during cardioprotection. Circulation research 88: 59-62, 2001.

224. Proks P, Reimann F, Green N, Gribble F, and Ashcroft F. Sulfonylurea stimulation of insulin secretion. Diabetes 51 Suppl 3: S368-376, 2002.

225. Puddu A, Salani B, Cordera R, Viviani GL, and Maggi D. Caveolin-1 is essential for -induced insulin secretion in the pancreatic betaTC-6 cell line. Biochemical and biophysical research communications 375: 235-237, 2008.

226. Qian YZ, Levasseur JE, Yoshida K, and Kukreja RC. KATP channels in rat heart: blockade of ischemic and acetylcholine-mediated preconditioning by glibenclamide. The American journal of physiology 271: H23-28, 1996.

227. Quesada I, Rovira JM, Martin F, Roche E, Nadal A, and Soria B. Nuclear KATP channels trigger nuclear Ca(2+) transients that modulate nuclear function. Proceedings of the National Academy of Sciences of the United States of America 99: 9544-9549, 2002.

228. Quinlan CL, Costa AD, Costa CL, Pierre SV, Dos Santos P, and Garlid KD. Conditioning the heart induces formation of signalosomes that interact with mitochondria to open mitoKATP channels. Am J Physiol Heart Circ Physiol 295: H953-H961, 2008.

229. Raikar LS, Vallejo J, Lloyd PG, and Hardin CD. Overexpression of caveolin-1 results in increased plasma membrane targeting of glycolytic

180 enzymes: the structural basis for a membrane associated metabolic compartment. Journal of cellular biochemistry 98: 861-871, 2006.

230. Rakhit RD, and Marber MS. Nitric oxide: an emerging role in cardioprotection? Heart (British Cardiac Society) 86: 368-372, 2001.

231. Razani B, Combs TP, Wang XB, Frank PG, Park DS, Russell RG, Li M, Tang B, Jelicks LA, Scherer PE, and Lisanti MP. Caveolin-1-deficient mice are lean, resistant to diet-induced obesity, and show hypertriglyceridemia with adipocyte abnormalities. The Journal of biological chemistry 277: 8635- 8647, 2002.

232. Razani B, Woodman SE, and Lisanti MP. Caveolae: from cell biology to animal physiology. Pharmacological reviews 54: 431-467, 2002.

233. Reers M, Smith TW, and Chen LB. J-aggregate formation of a carbocyanine as a quantitative fluorescent indicator of membrane potential. Biochemistry 30: 4480-4486, 1991.

234. Remme CA, and Wilde AA. KATP channel openers, myocardial ischemia, and arrhythmias--should the electrophysiologist worry? Cardiovascular drugs and therapy / sponsored by the International Society of Cardiovascular Pharmacotherapy 14: 17-22, 2000.

235. Ribalet B, John SA, and Weiss JN. Molecular basis for Kir6.2 channel inhibition by adenine nucleotides. Biophysical journal 84: 266-276, 2003.

236. Ribalet B, John SA, Xie LH, and Weiss JN. Regulation of the ATP- sensitive K channel Kir6.2 by ATP and PIP(2). Journal of molecular and cellular cardiology 39: 71-77, 2005.

237. Robenek H, Weissen-Plenz G, and Severs NJ. Freeze-fracture replica immunolabelling reveals caveolin-1 in the human cardiomyocyte plasma membrane. Journal of cellular and molecular medicine 12: 2519-2521, 2008.

238. Robia SL, Ghanta J, Robu VG, and Walker JW. Localization and kinetics of protein kinase C-epsilon anchoring in cardiac myocytes. Biophysical journal 80: 2140-2151, 2001.

181 239. Robia SL, Kang M, and Walker JW. Novel determinant of PKC-epsilon anchoring at cardiac Z-lines. Am J Physiol Heart Circ Physiol 289: H1941-1950, 2005.

240. Roscoe AK, Christensen JD, and Lynch C, 3rd. Isoflurane, but not halothane, induces protection of human myocardium via adenosine A1 receptors and adenosine triphosphate-sensitive potassium channels. Anesthesiology 92: 1692-1701, 2000.

241. Rothberg KG, Heuser JE, Donzell WC, Ying YS, Glenney JR, and Anderson RG. Caveolin, a protein component of caveolae membrane coats. Cell 68: 673-682, 1992.

242. Rybin VO, Xu X, and Steinberg SF. Activated protein kinase C isoforms target to cardiomyocyte caveolae : stimulation of local protein phosphorylation. Circulation research 84: 980-988, 1999.

243. Sabourin T, Bastien L, Bachvarov DR, and Marceau F. Agonist- induced translocation of the kinin B(1) receptor to caveolae-related rafts. Molecular pharmacology 61: 546-553, 2002.

244. Sampson LJ, Davies LM, Barrett-Jolley R, Standen NB, and Dart C. Angiotensin II-activated protein kinase C targets caveolae to inhibit aortic ATP- sensitive potassium channels. Cardiovascular research 76: 61-70, 2007.

245. Sampson LJ, Hayabuchi Y, Standen NB, and Dart C. Caveolae localize protein kinase A signaling to arterial ATP-sensitive potassium channels. Circulation research 95: 1012-1018, 2004.

246. Sanchez JA, Gonoi T, Inagaki N, Katada T, and Seino S. Modulation of reconstituted ATP-sensitive K(+)-channels by GTP-binding proteins in a mammalian cell line. The Journal of physiology 507 ( Pt 2): 315-324, 1998.

247. Sasaki N, Sato T, Ohler A, O'Rourke B, and Marban E. Activation of mitochondrial ATP-dependent potassium channels by nitric oxide. Circulation 101: 439-445, 2000.

182 248. Satoh H. Endothelin-1 inhibition of the ATP-sensitive K+ channel in guinea-pig ventricular cardiomyocytes. General pharmacology 26: 1549-1552, 1995.

249. Saupe KW, Spindler M, Hopkins JC, Shen W, and Ingwall JS. Kinetic, thermodynamic, and developmental consequences of deleting creatine kinase isoenzymes from the heart. Reaction kinetics of the creatine kinase isoenzymes in the intact heart. The Journal of biological chemistry 275: 19742-19746, 2000.

250. Saupe KW, Spindler M, Tian R, and Ingwall JS. Impaired cardiac energetics in mice lacking muscle-specific isoenzymes of creatine kinase. Circulation research 82: 898-907, 1998.

251. Schackow TE, and Ten Eick RE. Enhancement of ATP-sensitive potassium current in cat ventricular myocytes by beta-adrenoreceptor stimulation. The Journal of physiology 474: 131-145, 1994.

252. Scherer PE, and Lisanti MP. Association of phosphofructokinase-M with caveolin-3 in differentiated skeletal myotubes. Dynamic regulation by extracellular glucose and intracellular metabolites. The Journal of biological chemistry 272: 20698-20705, 1997.

253. Schulze D, Krauter T, Fritzenschaft H, Soom M, and Baukrowitz T. Phosphatidylinositol 4,5-bisphosphate (PIP2) modulation of ATP and pH sensitivity in Kir channels. A tale of an active and a silent PIP2 site in the N terminus. The Journal of biological chemistry 278: 10500-10505, 2003.

254. Schwanstecher M, Schwanstecher C, Chudziak F, Panten U, Clement JPt, Gonzalez G, Aguilar-Bryan L, and Bryan J. ATP-sensitive potassium channels. Methods in enzymology 294: 445-458, 1999.

255. Schwanstecher M, Sieverding C, Dorschner H, Gross I, Aguilar- Bryan L, Schwanstecher C, and Bryan J. Potassium channel openers require ATP to bind to and act through sulfonylurea receptors. The EMBO journal 17: 5529-5535, 1998.

256. Seharaseyon J, Ohler A, Sasaki N, Fraser H, Sato T, Johns DC, O'Rourke B, and Marban E. Molecular composition of mitochondrial ATP-

183 sensitive potassium channels probed by viral Kir gene transfer. Journal of molecular and cellular cardiology 32: 1923-1930, 2000.

257. Seino S. ATP-sensitive potassium channels: a model of heteromultimeric potassium channel/receptor assemblies. Annual review of physiology 61: 337-362, 1999.

258. Selivanov VA, Alekseev AE, Hodgson DM, Dzeja PP, and Terzic A. Nucleotide-gated KATP channels integrated with creatine and adenylate kinases: amplification, tuning and sensing of energetic signals in the compartmentalized cellular environment. Molecular and cellular biochemistry 256-257: 243-256, 2004.

259. Sharma DK, Brown JC, Choudhury A, Peterson TE, Holicky E, Marks DL, Simari R, Parton RG, and Pagano RE. Selective stimulation of caveolar endocytosis by glycosphingolipids and cholesterol. Molecular biology of the cell 15: 3114-3122, 2004.

260. Shi NQ, Ye B, and Makielski JC. Function and distribution of the SUR isoforms and splice variants. Journal of molecular and cellular cardiology 39: 51-60, 2005.

261. Shogomori H, and Brown DA. Use of detergents to study membrane rafts: the good, the bad, and the ugly. Biological chemistry 384: 1259-1263, 2003.

262. Shyng SL, Cukras CA, Harwood J, and Nichols CG. Structural determinants of PIP(2) regulation of inward rectifier K(ATP) channels. The Journal of general physiology 116: 599-608, 2000.

263. Shyng SL, and Nichols CG. Membrane phospholipid control of nucleotide sensitivity of KATP channels. Science (New York, NY 282: 1138- 1141, 1998.

264. Siemen D, Loupatatzis C, Borecky J, Gulbins E, and Lang F. Ca2+- activated K channel of the BK-type in the inner mitochondrial membrane of a human glioma cell line. Biochemical and biophysical research communications 257: 549-554, 1999.

184 265. Singh H, Hudman D, Lawrence CL, Rainbow RD, Lodwick D, and Norman RI. Distribution of Kir6.0 and SUR2 ATP-sensitive potassium channel subunits in isolated ventricular myocytes. Journal of molecular and cellular cardiology 35: 445-459, 2003.

266. Song KS, Scherer PE, Tang Z, Okamoto T, Li S, Chafel M, Chu C, Kohtz DS, and Lisanti MP. Expression of caveolin-3 in skeletal, cardiac, and smooth muscle cells. Caveolin-3 is a component of the sarcolemma and co- fractionates with dystrophin and dystrophin-associated glycoproteins. The Journal of biological chemistry 271: 15160-15165, 1996.

267. Sotgia F, Bonuccelli G, Minetti C, Woodman SE, Capozza F, Kemp RG, Scherer PE, and Lisanti MP. Phosphofructokinase muscle-specific isoform requires caveolin-3 expression for plasma membrane recruitment and caveolar targeting: implications for the pathogenesis of caveolin-related muscle diseases. The American journal of pathology 163: 2619-2634, 2003.

268. Suzuki M, Kotake K, Fujikura K, Inagaki N, Suzuki T, Gonoi T, Seino S, and Takata K. Kir6.1: a possible subunit of ATP-sensitive K+ channels in mitochondria. Biochemical and biophysical research communications 241: 693- 697, 1997.

269. Suzuki M, Li RA, Miki T, Uemura H, Sakamoto N, Ohmoto-Sekine Y, Tamagawa M, Ogura T, Seino S, Marban E, and Nakaya H. Functional roles of cardiac and vascular ATP-sensitive potassium channels clarified by Kir6.2- knockout mice. Circulation research 88: 570-577, 2001.

270. Suzuki M, Sasaki N, Miki T, Sakamoto N, Ohmoto-Sekine Y, Tamagawa M, Seino S, Marban E, and Nakaya H. Role of sarcolemmal K(ATP) channels in cardioprotection against ischemia/reperfusion injury in mice. The Journal of clinical investigation 109: 509-516, 2002.

271. Szabo I, Bock J, Jekle A, Soddemann M, Adams C, Lang F, Zoratti M, and Gulbins E. A novel potassium channel in lymphocyte mitochondria. The Journal of biological chemistry 280: 12790-12798, 2005.

272. Szewczyk A, Wojcik G, Lobanov NA, and Nalecz MJ. The mitochondrial sulfonylurea receptor: identification and characterization. Biochemical and biophysical research communications 230: 611-615, 1997.

185 273. Tang Z, Scherer PE, Okamoto T, Song K, Chu C, Kohtz DS, Nishimoto I, Lodish HF, and Lisanti MP. Molecular cloning of caveolin-3, a novel member of the caveolin gene family expressed predominantly in muscle. J Biol Chem 271: 2255-2261, 1996.

274. Terzic A, Tung RT, Inanobe A, Katada T, and Kurachi Y. G proteins activate ATP-sensitive K+ channels by antagonizing ATP-dependent gating. Neuron 12: 885-893, 1994.

275. Tong X, Porter LM, Liu G, Dhar-Chowdhury P, Srivastava S, Pountney DJ, Yoshida H, Artman M, Fishman GI, Yu C, Iyer R, Morley GE, Gutstein DE, and Coetzee WA. Consequences of cardiac myocyte-specific ablation of KATP channels in transgenic mice expressing dominant negative Kir6 subunits. Am J Physiol Heart Circ Physiol 291: H543-551, 2006.

276. Towbin H, and Gordon J. Immunoblotting and dot immunobinding-- current status and outlook. Journal of immunological methods 72: 313-340, 1984.

277. Troyan MB, Gilman VR, and Gay CV. Mitochondrial membrane potential changes in osteoblasts treated with parathyroid hormone and . Experimental cell research 233: 274-280, 1997.

278. Truscott KN, Brandner K, and Pfanner N. Mechanisms of protein import into mitochondria. Curr Biol 13: R326-337, 2003.

279. Tseng GN, and Hoffman BF. Actions of pinacidil on membrane currents in canine ventricular myocytes and their modulation by intracellular ATP and cAMP. Pflugers Arch 415: 414-424, 1990.

280. Tsuchiya K, Horie M, Haruna T, Ai T, Nishimoto T, Fujiwara H, and Sasayama S. Functional communication between cardiac ATP-sensitive K+ channel and Na/K ATPase. Journal of cardiovascular electrophysiology 9: 415- 422, 1998.

281. Tsuchiya K, Horie M, Watanuki M, Albrecht CA, Obayashi K, Fujiwara H, and Sasayama S. Functional compartmentalization of ATP is involved in angiotensin II-mediated closure of cardiac ATP-sensitive K+ channels. Circulation 96: 3129-3135, 1997. 186 282. Vallejo J, and Hardin CD. Expression of caveolin-1 in lymphocytes induces caveolae formation and recruitment of phosphofructokinase to the plasma membrane. Faseb J 19: 586-587, 2005.

283. Vendelin M, Beraud N, Guerrero K, Andrienko T, Kuznetsov AV, Olivares J, Kay L, and Saks VA. Mitochondrial regular arrangement in muscle cells: a "crystal-like" pattern. American journal of physiology 288: C757-767, 2005.

284. Venkatesh N, Lamp ST, and Weiss JN. Sulfonylureas, ATP-sensitive K+ channels, and cellular K+ loss during hypoxia, ischemia, and metabolic inhibition in mammalian ventricle. Circulation research 69: 623-637, 1991.

285. Verdonck F, Mubagwa K, and Sipido KR. [Na(+)] in the subsarcolemmal 'fuzzy' space and modulation of [Ca(2+)](i) and contraction in cardiac myocytes. Cell calcium 35: 603-612, 2004.

286. Voos W. A new connection: chaperones meet a mitochondrial receptor. Molecular cell 11: 1-3, 2003.

287. Wada Y, Yamashita T, Imai K, Miura R, Takao K, Nishi M, Takeshima H, Asano T, Morishita R, Nishizawa K, Kokubun S, and Nukada T. A region of the sulfonylurea receptor critical for a modulation of ATP-sensitive K(+) channels by G-protein betagamma-subunits. The EMBO journal 19: 4915-4925, 2000.

288. Walker DM, Pasini E, Kucukoglu S, Marber MS, Iliodromitis E, Ferrari R, and Yellon DM. Heat stress limits infarct size in the isolated perfused rabbit heart. Cardiovascular research 27: 962-967, 1993.

289. Wan TC, Ge ZD, Tampo A, Mio Y, Bienengraeber MW, Tracey WR, Gross GJ, Kwok WM, and Auchampach JA. The A3 adenosine receptor agonist CP-532,903 [N6-(2,5-dichlorobenzyl)-3'-aminoadenosine-5'-N- methylcarboxamide] protects against myocardial ischemia/reperfusion injury via the sarcolemmal ATP-sensitive potassium channel. The Journal of pharmacology and experimental therapeutics 324: 234-243, 2008.

290. Wang XL, Ye D, Peterson TE, Cao S, Shah VH, Katusic ZS, Sieck GC, and Lee HC. Caveolae targeting and regulation of large conductance 187 Ca(2+)-activated K+ channels in vascular endothelial cells. The Journal of biological chemistry 280: 11656-11664, 2005.

291. Wang Y, and Ashraf M. Role of protein kinase C in mitochondrial KATP channel-mediated protection against Ca2+ overload injury in rat myocardium. Circulation research 84: 1156-1165, 1999.

292. Wang Y, Haider HK, Ahmad N, and Ashraf M. Mechanisms by which K(ATP) channel openers produce acute and delayed cardioprotection. Vascular pharmacology 42: 253-264, 2005.

293. Wang Y, Hirai K, and Ashraf M. Activation of mitochondrial ATP- sensitive K(+) channel for cardiac protection against ischemic injury is dependent on protein kinase C activity. Circulation research 85: 731-741, 1999.

294. Wang YG, and Lipsius SL. beta-Adrenergic stimulation induces acetylcholine to activate ATP-sensitive K+ current in cat atrial myocytes. Circulation research 77: 565-574, 1995.

295. Watanuki M, Horie M, Tsuchiya K, Obayashi K, and Sasayama S. Endothelin-1 inhibition of cardiac ATP-sensitive K+ channels via pertussis-toxin- sensitive G-proteins. Cardiovascular research 33: 123-130, 1997.

296. Weiss JN, and Lamp ST. Cardiac ATP-sensitive K+ channels. Evidence for preferential regulation by glycolysis. The Journal of general physiology 94: 911-935, 1989.

297. Weiss JN, and Lamp ST. Glycolysis preferentially inhibits ATP-sensitive K+ channels in isolated guinea pig cardiac myocytes. Science (New York, NY 238: 67-69, 1987.

298. Weiss JN, Yang L, and Qu Z. Systems biology approaches to metabolic and cardiovascular disorders: network perspectives of cardiovascular metabolism. Journal of lipid research 47: 2355-2366, 2006.

299. Wilde AA. Role of ATP-sensitive K+ channel current in ischemic arrhythmias. Cardiovascular drugs and therapy / sponsored by the International Society of Cardiovascular Pharmacotherapy 7 Suppl 3: 521-526, 1993.

188 300. Williams TM, and Lisanti MP. The caveolin proteins. Genome biology 5: 214, 2004.

301. Woodman SE, Park DS, Cohen AW, Cheung MW, Chandra M, Shirani J, Tang B, Jelicks LA, Kitsis RN, Christ GJ, Factor SM, Tanowitz HB, and Lisanti MP. Caveolin-3 knock-out mice develop a progressive cardiomyopathy and show hyperactivation of the p42/44 MAPK cascade. The Journal of biological chemistry 277: 38988-38997, 2002.

302. Wu JM, Xiao L, Cheng XK, Cui LX, Wu NH, and Shen YF. PKC epsilon is a unique regulator for hsp90 beta gene in heat shock response. The Journal of biological chemistry 278: 51143-51149, 2003.

303. Xi L, Das A, Zhao ZQ, Merino VF, Bader M, and Kukreja RC. Loss of myocardial ischemic postconditioning in adenosine A1 and bradykinin B2 receptors gene knockout mice. Circulation 118: S32-37, 2008.

304. Xia F, Gao X, Kwan E, Lam PP, Chan L, Sy K, Sheu L, Wheeler MB, Gaisano HY, and Tsushima RG. Disruption of pancreatic beta-cell lipid rafts modifies Kv2.1 channel gating and insulin . The Journal of biological chemistry 279: 24685-24691, 2004.

305. Xiang Y, Rybin VO, Steinberg SF, and Kobilka B. Caveolar localization dictates physiologic signaling of beta 2-adrenoceptors in neonatal cardiac myocytes. The Journal of biological chemistry 277: 34280-34286, 2002.

306. Xie LH, Takano M, Kakei M, Okamura M, and Noma A. Wortmannin, an inhibitor of phosphatidylinositol kinases, blocks the MgATP-dependent recovery of Kir6.2/SUR2A channels. The Journal of physiology 514 ( Pt 3): 655- 665, 1999.

307. Xu KY, and Becker LC. Ultrastructural localization of glycolytic enzymes on sarcoplasmic reticulum vesticles. J Histochem Cytochem 46: 419-427, 1998.

308. Xu W, Liu Y, Wang S, McDonald T, Van Eyk JE, Sidor A, and O'Rourke B. Cytoprotective role of Ca2+- activated K+ channels in the cardiac inner mitochondrial membrane. Science (New York, NY 298: 1029-1033, 2002.

189 309. Yamada E. The fine structure of the gall bladder epithelium of the mouse. The Journal of biophysical and biochemical cytology 1: 445-458, 1955.

310. Yamada M, Isomoto S, Matsumoto S, Kondo C, Shindo T, Horio Y, and Kurachi Y. Sulphonylurea receptor 2B and Kir6.1 form a sulphonylurea- sensitive but ATP-insensitive K+ channel. The Journal of physiology 499 ( Pt 3): 715-720, 1997.

311. Yao Z, and Gross GJ. Effects of the KATP channel opener bimakalim on coronary blood flow, monophasic action potential duration, and infarct size in dogs. Circulation 89: 1769-1775, 1994.

312. Yao Z, and Gross GJ. Glibenclamide antagonizes adenosine A1 receptor-mediated cardioprotection in stunned canine myocardium. Circulation 88: 235-244, 1993.

313. Yarbrough TL, Lu T, Lee HC, and Shibata EF. Localization of cardiac sodium channels in caveolin-rich membrane domains: regulation of sodium current amplitude. Circ Res 90: 443-449, 2002.

314. Yellon DM, and Marber MS. Hsp70 in myocardial ischaemia. Experientia 50: 1075-1084, 1994.

315. Young JC, Hoogenraad NJ, and Hartl FU. Molecular chaperones Hsp90 and Hsp70 deliver preproteins to the mitochondrial import receptor Tom70. Cell 112: 41-50, 2003.

316. Zajchowski LD, and Robbins SM. Lipid rafts and little caves. Compartmentalized signalling in membrane microdomains. European journal of biochemistry / FEBS 269: 737-752, 2002.

317. Zaugg M, and Schaub MC. Signaling and cellular mechanisms in cardiac protection by ischemic and pharmacological preconditioning. Journal of muscle research and cell motility 24: 219-249, 2003.

318. Zerangue N, Schwappach B, Jan YN, and Jan LY. A new ER trafficking signal regulates the subunit stoichiometry of plasma membrane K(ATP) channels. Neuron 22: 537-548, 1999.

190 319. Zhou M, Tanaka O, Sekiguchi M, He HJ, Yasuoka Y, Itoh H, Kawahara K, and Abe H. ATP-sensitive K+-channel subunits on the mitochondria and endoplasmic reticulum of rat cardiomyocytes. J Histochem Cytochem 53: 1491-1500, 2005.

320. Zingman LV, Hodgson DM, Bast PH, Kane GC, Perez-Terzic C, Gumina RJ, Pucar D, Bienengraeber M, Dzeja PP, Miki T, Seino S, Alekseev AE, and Terzic A. Kir6.2 is required for adaptation to stress. Proc Natl Acad Sci U S A 99: 13278-13283, 2002.

191