+ All Categories
Home > Documents > The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

Date post: 05-Nov-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
178
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2013-12-04 The Role of L- and T-type Ca2+ Channels in Rat Cerebral Arteries Abd El-Rahman, Rasha Abd El-Rahman, R. (2013). The Role of L- and T-type Ca2+ Channels in Rat Cerebral Arteries (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/28330 http://hdl.handle.net/11023/1172 doctoral thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca
Transcript
Page 1: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2013-12-04

The Role of L- and T-type Ca2+ Channels in Rat

Cerebral Arteries

Abd El-Rahman, Rasha

Abd El-Rahman, R. (2013). The Role of L- and T-type Ca2+ Channels in Rat Cerebral Arteries

(Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/28330

http://hdl.handle.net/11023/1172

doctoral thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

UNIVERSITY OF CALGARY

The Role of L- and T-type Ca2+ Channels in Rat Cerebral Arteries

by

Rasha R. Abd El-Rahman

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CARDIOVASCULAR AND RESPIRATORY SCIENCES

CALGARY, ALBERTA

November, 2013

Rasha R. Abd El-Rahman 2013

Page 3: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

ii

Abstract

The overall goal of this thesis was to identify which voltage-gated Ca2+ channels are

expressed in rat cerebral arterial smooth muscle and to determine their contributions to

myogenic tone regulation. We began by exploring which voltage-gated Ca2+ channels are

expressed in cerebral arterial smooth muscle. A combination of molecular,

electrophysiological and functional measurements revealed the presence of L- (CaV1.2) and

T-type (CaV3.1 and CaV3.2) Ca2+ channel subtypes in rat cerebral arteries. Both types

contribute to arterial tone development, although the contribution of the L-type channels to

tone development is greater. We then investigated the role of a specific T-type Ca2+ channel

subtype, CaV3.2, in cerebral arterial smooth muscle by functional assessment and a structural

approach using immunohistochemistry, proximity ligation assay, electron-tomography, and

immunogold labeling, combined with computational modeling and electrophysiological

measurements. Results indicate that Ca2+ influx through CaV3.2 channels elicits dilation by

activating ryanodine receptors and inducing Ca2+ sparks, localized events that activate BKCa

channels. In conclusion this work provided evidence for the presence of different types of

voltage-gated Ca2+ channels and provided evidence of their diverse functional roles in

regulating myogenic tone in rat cerebral arteries. Overall, the conclusions indicate the

importance of the different functional roles of voltage-gated Ca2+ channels, which have

substantial physiological relevance to the function of the cerebral

vasculature.………………………………………..

Page 4: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

ii

Acknowledgments

This thesis work would not have been possible without the help, support and

encouragement of many wonderful people. I would like to express my sincere gratitude to

my interim supervisor, Dr. Gary Kargacin, whose expertise, and patience, added

considerably to my graduate experience. I thank him for always being available when I

needed him the most. I am forever grateful to Dr. Meg Kargacin and Dr. Michael Walsh for

their invaluable guidance, help in editing my thesis, and encouragement. It is a pleasure to

thank those who made this thesis possible, and I owe my deepest gratitude to my supervisory

committee members, Dr. Michael Walsh, and Dr. William C. Cole, who have provided me

with valuable feedback and insight for my projects, my thesis and their tremendous input and

many stimulating discussions, whom without their help, I could not proceed with this

journey. All of my appreciation goes to Dr. Andy Braun for his great help and support to

make this thesis come together. Special thanks go to Dr. Ray Turner and his technician Mirna

Kruskic for their great help and support in teaching me and answering my questions. I would

like to extend my warm gratitude to my friend and colleague Dr. Rania Mufti for her great

support, advice and guidance. I also would like to thank my past and present colleagues Dr.

Kevin Luykenaar, Dr. Cam Ha Tran and Osama Harraz for providing me with their support

and suggestions during these years. Thank you for being there for me at all times and keeping

me motivated and inspired.

Lastly, I would not be where I am without the love and support of my family and words

cannot possibly capture my deep gratitude. First and foremost, I wish to thank my beloved

parents for their unconditional love, for raising me to be the person that I’m now, for giving

Page 5: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

iii

me the freedom to pursue my dreams and for being there whenever I need guidance and

support. Mom, I wish you were here but I know that you are watching over me. Last, but by

no means least, I would like to thank my husband Ehab Ali whose love, support and

encouragement allowed me to finish this journey, and my lovely daughters Malak Ali and

Laila Ali, who helped and taught me a lot during this journey.

My thanks to many of my friends Dr. Eman Afkar, Naglaa Ali, Rasha Saleh, Nermeen

Sharaf, and Abeer Ahmad, for cheering me up when I was down and for being my supporters

but also my critics and most of all for their friendship.

Page 6: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

vi

Dedication

I dedicate this thesis to the soul of my beloved mom and my dad, my husband Ehab,

and my two lovely daughters (Malak & Laila), my brothers Mohamed and Ahmad, without

whom this journey would have been impossible.

Page 7: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

vi

Table of Contents

Abstract ................................................................................................................................ ii

Acknowledgments ............................................................................................................... iii

Dedication ............................................................................................................................. v

Table of Contents ................................................................................................................. vi

List of Tables ....................................................................................................................... x

List of Figures and Illustrations .......................................................................................... xi

List of Symbols, Abbreviations and Nomenclature ........................................................... xiii

CHAPTER ONE: INTRODUCTION .................................................................................. 1

1.1 Overview .................................................................................................................... 1

1.2 Role of Ca2+ in smooth muscle (Ca2+ homeostasis). .................................................. 3

1.3 Ca2+ channels ............................................................................................................. 7

1.3.1 Non-voltage-gated Ca2+ channels ................................................................... 7

1.3.2 Voltage-gated Ca2+ channels (VGCCs) .......................................................... 9

1.3.3 General structure of VGCCs .......................................................................... 9

1.3.4 L-type channel structure ............................................................................... 10

1.3.5 T-type channel structure ............................................................................... 13

1.3.6 General functions of VGCCs ....................................................................... 14

1.3.7 L-type channel (CaV1.2) function................................................................. 15

1.3.8 T-type channel (CaV3.1 and CaV3.2) function ............................................ 15

1.4 Potassium channels in smooth muscle ...................................................................... 18

1.5 Sarcoplasmic reticulum (SR) .................................................................................... 19

1.5.1 Inositol 1,4,5-trisphosphate receptors (IP3R) .............................................. 19

1.5.2 Ryanodine receptors (RyR) .......................................................................... 20

1.5.3 Ca2+ sparks .................................................................................................. 22

1.6 Summary ................................................................................................................... 23

1.7 Hypothesis ................................................................................................................ 24

Page 8: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

vii

CHAPTER TWO: IDENTIFICATION OF L- AND T-TYPE CA2+ CHANNELS IN RAT

CEREBRAL ARTERIES: ROLE IN MYOGENIC TONE DEVELOPMENT…………...25

2.1 Introduction ............................................................................................................... 26

2.2 Materials and Methods .............................................................................................. 28

2.2.1 Animal procedures ........................................................................................... 28

2.2.2 Vessel myography ........................................................................................... .28

2.2.3 Isolation of arterial smooth muscle cells .......................................................... 29

2.2.4 PCR analysis .................................................................................................... 30

2.2.5 Western blotting .............................................................................................. 31

2.2.6 Electrophysiology ............................................................................................ 32

2.2.7 Computational blood flow modeling ............................................................... 33

2.2.8 Statistical analysis ........................................................................................... 34

2.2.9 Solutions and chemicals .................................................................................. 34

2.3 Results ....................................................................................................................... 34

2.3.1 Ca2+ channel α1-subunit expression in middle and posterior cerebral

arteries…………………………………………………………………………….....34

2.3.2 Whole-cell Ba2+ currents in cerebral arterial smooth muscle cells .................. 35

2.3.3 Ca2+ channel blockers and myogenic tone ....................................................... 38

2.4 Discussion ................................................................................................................. 40

2.4.1 Background ...................................................................................................... 41

2.4.2 Molecular and electrical characterization ......................................................... 42

2.4.3 T-type Ca2+ channels and myogenic tone ......................................................... 44

2.4.4 Implications to blood flow control ................................................................... 46

2.4.5 Summary .......................................................................................................... 46

2.4.6 Limitations and future directions………………………………………….…...47

(i) Molecular and biochemical approach…………………...……………………...47

(ii) Patch-clamp electrophysiological approach..…………………………….……51

Page 9: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

viii

(iii) Vessel myography approach……………………………………………….…..54

(iv) Computational approach………………………………………………….……55

2.5 Figures ....................................................................................................................... 57

CHAPTER THREE: CaV3.2 CHANNELS AND THE INDUCTION OF A NEGATIVE

FEEDBACK RESPONSE IN CEREBRAL ARTERIAL SMOOTH MUSCLE .............. .81

3.1 Introduction ............................................................................................................... 82

3.2 Materials and Methods .............................................................................................. 83

3.2.1 Animal procedures .......................................................................................... 83

3.2.2 Vessel myography and measurements of membrane potentials ....................... 84

3.2.3 Immunohistochemistry ..................................................................................... 85

3.2.4 Electron tomography ........................................................................................ 86

3.2.5 Immunogold-labeling ....................................................................................... 87

3.2.6 Isolation of arterial smooth muscle cells .......................................................... 88

3.2.7 Proximity ligation assay (PLA) ........................................................................ 88

3.2.8 Computational modeling .................................................................................. 89

3.2.9 Electrophysiology ............................................................................................. 90

3.2.10 Ca2+ sparks ..................................................................................................... 91

3.2.11 Statistical analysis .......................................................................................... 92

3.2.12 Solutions and chemicals ................................................................................. 92

3.3 Results ....................................................................................................................... 93

3.3.1 Vasomotor and electrical effects of CaV3.2 blockade ...................................... 93

3.3.2 Ca2+ channel localization .................................................................................. 93

3.3.3 Computational modeling ................................................................................. 97

3.3.4 BKCa currents and Ca2+ sparks ......................................................................... 97

3.3.5 Functional implications of CaV3.2 blockade .................................................... 98

3.4 Discussion ................................................................................................................. 99

3.4.1 Background ..................................................................................................... 99

3.4.2 Mechanism of CaV3.2-induced vasodilation .................................................. 100

Page 10: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

ix

3.4.3 Summary ........................................................................................................ 103

3.4.4 Limitations and future directions ................................................................... 104

(i) Vessel myography approach ............................................................................ 104

(ii) Structural approach .......................................................................................... 105

(iii) STOCs and Ca2+ sparks ................................................................................. 109

(iv) Computational approach ................................................................................. 111

3.5 Figures ..................................................................................................................... 112

CHAPTER FOUR: DISCUSSION ................................................................................... 128

4.1 Overview ................................................................................................................. 128

4.2 Objective #1: Identification of L- and T-type Ca2+ channel subtypes in rat

cerebral arterial smooth muscle cells and their roles in myogenic tone

development ........................................................................................................... .129

4.3 Objective #2: CaV3.2 and the induction of a negative feedback response in

cerebral arterial smooth muscle ......................................................... ……………..131

4.4 Future directions, experimental considerations, clinical relevance, and

conclusion…………………………………………………………………….........134

4.4.1 Future directions ............................................................................................. 134

4.4.2 Clinical relevance .......................................................................................... 144

4.4.3 Conclusion ...................................................................................................... 145

APPENDIX ....................................................................................................................... 146

REFERENCES ................................................................................................................. 153

Page 11: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

x

List of Tables

Table 1. IC50 values for inhibition of VGCCs by various blockers………………………....17

Table 2. Immunogold-labelling for RyR and CaV3.2……………………………………….96

Page 12: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

xi

List of Figures and Illustrations

Figure 1. Structure of arteries…………………………………………………………….…2

Figure 2. Regulation of smooth muscle contraction through the phosphorylation

and dephosphorylation of the myosin II regulatory light chain (LC20) by

MLCK and MLCP, respectively……………………………………………….....5

Figure 3. The topology of the high voltage-gated calcium channels.……………………...11

Figure 4. mRNA expression of the α1-subunits of voltage-gated Ca2+ channels….….......57

Figure 5. Protein expression of the α1-subunits of voltage-gated Ca2+ channels……….....59

Figure 6. Inward Ba2+ currents in cerebral arterial smooth muscle cells……………….....61

Figure 7. Nifedipine-sensitive and -insensitive Ba2+ currents in cerebral

arterial smooth muscle cells……………………………………………………..63

Figure 8. Effects of mibefradil on the nifedipine-insensitive component of the

inward Ba2+ current……………………………………………………………...65

Figure 9. Effects of kurtoxin on the nifedipine-insensitive component of the

inward Ba2+current….………………………………………………………......67

Figure 10. Effects of efonidipine on the nifedipine-insensitive component of the

inward Ba2+ current…………….…………………………………..…………….69

Figure 11. The effects of nifedipine and mibefradil on myogenic tone……………..….…..71

Figure 12. The effects of nifedipine and kurtoxin on myogenic tone…………….…..….…73

Figure 13. The effects of nifedipine and efonidipine on myogenic tone…………………...75

Figure 14. The effects of nifedipine and mibefradil on myogenic tone through a full

range of intravascular pressures……………………………………………...….77

Figure 15. Computational modeling predicts that T-type Ca2+ channel modulation

alters tissue blood flow……………………………………………………….....79

Figure 16. Effect of Ni2+ on pressure-induced tone and membrane potential (Vm)

in rat cerebral arteries.……………………………………………………........112

Figure 17. Ca2+ channel localization in rat cerebral arteries…..…………………………....114

Figure 18. Controls of smooth muscle actin, CaV3.2, RyR, and CaV1.2……...……………116

Page 13: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

xii

Figure 19 Proximity ligation assay and the colocalization of CaV3.2 and RyR in

rat cerebral arterial smooth muscle cells……………………………………….118

Figure 20. Electron microscopic imaging of rat cerebral arterial smooth muscle cells…..120

Figure 21. A microdomain model of smooth muscle Ca2+ dynamics……………………..122

Figure 22. Spontaneous transient outward currents (STOCs) and Ca2+ spark

production in cerebral arterial smooth muscle cells……………………………124

Figure 23. Effect of Ni2+ and paxilline on myogenic tone in rat cerebral arteries………..126

Figure 24. Diagram highlighting the proposed function of CaV3.2 channels in

the regulation of arterial tone………………………………………………....133

Figure 25. Diagram highlighting the different proposed possible functions

of CaV3.2 channels in the regulation of arterial tone………………………....142

Figure 26. Diagram highlighting the possible separate functions of CaV3.1 and

CaV3.2 channels in the regulation of arterial tone…………………………….143

Figure 27. CaV3.1 protein expression in vascular smooth muscle cells……………………147

Figure 28. BKCa protein expression in vascular smooth muscle cells……………………..149

Figure 29. IP3R protein expression in vascular smooth muscle cells………………….......151

Page 14: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

xiii

List of Symbols, Abbreviations and Nomenclature

ATP Adenosine 5´-triphosphate

BKCa Large conductance calcium-activated potassium channels

[Ca2+]i Cytosolic free Ca2+ concentration

CaM Calmodulin

cAMP Cyclic adenosine 3´:5´-monophosphate

cGMP Cyclic guanosine 3´:5´-monophosphate

CICR Ca2+-induced Ca2+ release

CNS Central nervous system

DAG Diacylglycerol

DMSO Dimethyl sulfoxide

DTT Dithiothreitol

EGTA Ethylene glycol tetraacetic acid

EM Electron microscope

FRET Fluorescence resonance energy transfer

GPCR G protein-coupled receptor

HEK Human embryonic kidney cells

HRP Horseradish peroxidase

IC50 The half maximal inhibitory concentration

IP3 Inositol 1, 4, 5-trisphosphate

IP3R IP3 receptor

KATP ATP-sensitive potassium channel

KIR Inward rectifier potassium channel

KV Voltage-gated potassium channel

MLC20 20-kDa myosin regulatory light chain

MLCK Myosin light chain kinase

MLCP Myosin light chain phosphatase

MYPT1 Myosin phosphatase targeting subunit 1

Page 15: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

xiv

NCX Na+/Ca2+ exchanger

PBS Phosphate-buffered saline

PIP2 Phosphatidylinositol-4,5-bisphosphate

PKA cAMP-dependent protein kinase (protein kinase A)

PKC Protein kinase C

PKG cGMP-dependent protein kinase (protein kinase G)

PLC Phospholipase C

PLCγ1 Phospholipase Cγ1 isoform

PP1c Phosphatase type I catalytic subunit

PSS Physiological salt solution

ROK Rho-associated protein kinase

ROCs Receptor-operated channels

RyR Ryanodine receptor

SACs Stretch-activated channels

SMC Smooth muscle cells

SDS-PAGE Sodium dodecyl sulfate-polyacrylamide gel electrophoresis

SOCCs Store-operated cation channels

SR Sarcoplasmic reticulum

STIC Spontaneous transient inward current

STIM1 Stromal interaction molecule 1

STOC Spontaneous transient outward current

TRP Transient receptor potential

TIC Transient inward current

TMEM16A Ca2+-activated Cl- conductance

VGCC Voltage-gated Ca2+ channel

Vm Membrane potential

Page 16: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

1

Chapter One: INTRODUCTION

1.1 Overview.

The cardiovascular system consists of the heart, blood, and blood vessels. The principal

function of blood vessels is to transport blood throughout the body. There are three main

types of blood vessels: the arteries, which carry blood away from the heart; the capillaries,

where exchange of water and chemicals between the blood and the tissues occurs; and the

veins that transmit blood from the capillaries back to the heart. Arterial blood vessels can be

further divided into: 1) conduit arteries (e.g. aorta) that conduct blood to peripheral tissues;

and 2) resistance vessels (e.g. cerebral and feed arteries) that regulate blood flow to various

organs depending upon their physiological needs (Segal & Jacobs, 2001). Conduit vessels

are thought to change their diameter passively in response to the volume of blood entering or

leaving them. Resistance arteries, on the other hand, actively change diameter in response to

a variety of stimuli including: 1) shear stress and intraluminal pressure (Davies, 2009; Knot

& Nelson, 1995; Segal, 2000); 2) sympathetic neuronal outputs (Brayden & Bevan, 1985;

Lee, 1982; Si & Lee, 2002); and 3) hormonal and metabolic factors (Filosa et al., 2006;

Zhang et al., 2005). Arteries are comprised of three layers including: 1) the tunica adventitia,

an outer layer of connective tissue; 2) the tunica media, a layer primarily containing smooth

muscle cells; and 3) the tunica intima, an inner layer of endothelial cells lining the blood

vessel wall (Figure 1; Faraci & Heistad, 1990).

Page 17: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

2

Figure 1: Structure of arteries. The arterial wall is composed of three layers: an inner layer, the

tunica intima, composed mainly of endothelial cells; a middle layer, the tunica media, with

smooth muscle cells as the main component; and an outer layer, the tunica adventitia, composed

mainly of collagenous tissue (Davis et al., 2011).

Page 18: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

3

1.2 Role of Ca2+ in smooth muscle (Ca2+ homeostasis).

Ca2+ is an essential second messenger involved in controlling and regulating numerous

cellular functions and reactions including muscle contraction, fertilization, cell locomotion, gene

expression, and apoptosis. Because high levels of Ca2+ are toxic to cells and can lead to cell

death, intracellular Ca2+ is carefully regulated (Wray et al., 2005). Ca2+ signaling in vascular

smooth muscle is complex, however, in recent years, advanced technologies have provided tools

to better understand the roles and regulation of Ca2+ in different cells and tissues (Carafoli et al.,

2001; Wray et al., 2005). Improved cell culture methods, the availability of high-affinity

fluorescent Ca2+ indicators (including genetically-encoded indicators based on green fluorescent

proteins and membrane-permeant small molecules, e.g. Fura-2, and Fluo-4) combined with high-

resolution fluorescent microscopy have allowed significant insights into intracellular Ca2+

signaling (Knot et al., 2005). The intracellular free Ca2+ concentration [Ca2+]i ranges from resting

levels of 100-200 nM up to levels of ~700 nM during contraction while extracellular [Ca2+]free is

1-2 mM. Thus Ca2+ will be favoured to enter the cell and [Ca2+]i must be regulated by various

ion channels, exchangers and pumps found in the plasmalemmal and sarcoplasmic reticulum

(SR) membranes of smooth muscle cells (Berridge et al., 2000; Carafoli et al., 2001; Wray et al.,

2005). Smooth muscle cells are responsible for the contractility of all hollow organs, including

blood vessels, the gastrointestinal tract, the bladder and the uterus. In resistance arteries, smooth

muscle cells actively relax and contract to a variety of stimuli that effectively change vessel

diameter and consequently regulate tissue blood flow. Increases in [Ca2+]i elicit smooth muscle

contraction initiated by the binding of Ca2+ to calmodulin (CaM) and the activation of myosin

Page 19: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

4

light chain kinase (MLCK). In general, the smooth muscle literature recognizes two principal

mechanisms responsible for the rise in [Ca2+]i. The first involves the activation of a G protein-

coupled receptor (GPCR) by either a neurotransmitter or hormone, where the stimulation of the

receptor activates phospholipase C (PLC) that in turn evokes the production of inositol 1,4,5-

trisphosphate (IP3) and diacylglycerol (DAG). The binding of IP3 to IP3 receptors (IP3R) on the

SR leads to Ca2+ release from the SR through the receptors. The second mechanism involves an

increase in resting membrane potential and the opening of voltage-gated Ca2+ channels

(VGCCs). This induces an influx of extracellular Ca2+ (Figure 2; Knot & Wilson, 1998). This

mechanism has long been proposed to play an essential role in evoking contraction of vascular

smooth muscle (Knot & Nelson, 1998; Welsh et al., 2002; Welsh et al., 2000). There are several

types of VGCC including L-type (e.g. CaV1.2) and T-type channels (e.g. CaV3.1 and CaV3.2) that

are believed to be expressed in vascular smooth muscle cells (Abd El-Rahman et al., 2013; Kuo

et al., 2010; Nikitina et al., 2007).

Page 20: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

5

Figure 2: Regulation of smooth muscle contraction through the phosphorylation and

dephosphorylation of the myosin II regulatory light chain (LC20) by MLCK and MLCP,

respectively. The influx of Ca2+ through VGCC elevates global cytosolic [Ca2+] and can also

induce Ca2+ release from the SR. Ca2+ binding to calmodulin activates MLCK and

phosphorylates MLC20 to induce contraction by cross-bridge cycling between actin and myosin.

Meanwhile, dephosphorylation of MLC20 by MLCP elicits relaxation. Signaling pathways via G

protein-coupled receptors (GPCR): 1) Gq/11 activates phospholipase Cβ (PLCβ), which converts

Page 21: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

6

phosphatidylinositol 4,5-bisphosphate (PIP2) into inositol 1,4,5-trisphosphate (IP3) and

diacylglycerol (DAG). IP3 activates Ca2+ release from the SR via inositol 1,4,5-trisphosphate

receptors (IP3R). DAG that activates protein kinase C (PKC); 2) G12/13 is coupled to RhoGEF,

which activates RhoA, which in turn stimulates Rho-associated kinase (ROK). PKC and ROK

phosphorylate CP1-17, and the myosin phosphatase targeting subunit 1 (MYPT1) of MLCP,

respectively, to inhibit MLCP and elicit Ca2+ sensitization of contraction seen as a leftward shift

in the force vs. [Ca2+]i curve in the lower left in the figure (Cole & Welsh, 2011).

Page 22: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

7

1.3 Ca2+ channels.

Ca2+ ion entry into the cytosol mediates a wide range of cellular responses, including the

activation of muscle contraction, regulation of Ca2+-dependent enzymes, regulation of gene

transcription, and apoptosis (Dolmetsch et al., 2001; Kisilevsky & Zamponi, 2008; Martin-

Moutot et al., 1996; Sutton et al., 1999). Ca2+ may enter cells by a number of different

mechanisms, including non-voltage gated Ca2+ channels and voltage-gated Ca2+ channels

(VGCCs).

1.3.1 Non-voltage-gated Ca2+ channels.

Non-voltage-gated Ca2+ channels include receptor-operated channels (ROCs), store-

operated Ca2+ channels (SOCCs), and mechano-sensitive or stretch-activated channels (SACs)

(Burnstock, 1972; Benham et al, 1985; Loirand et al, 1991; Fleischmann et al, 1997). ROCs are

regulated by agonist–receptor interaction and they are activated by many different agonists (e.g.

acetylcholine, angiotensin II, endothelin-1, norepinephrine, serotonin and vasopressin) acting

through plasma membrane receptors. They also include the ligand-gated channel P2X that is

activated by ATP (Sanders, 2001; Benham et al, 1985). ROC current is a non-selective cation

current (Icat) that is carried by both mono- and divalent cations with differing degrees of Ca2+

selectivity (Sanders, 2001). SOCCs are activated by the emptying of intracellular Ca2+ stores

(Berridge, 1995; Hirota et al., 2007; Jones & Braun, 2009; Lee et al., 2005). Growing evidence

suggests that proteins, such as Orai1 and STIM1, are components of SOCCs. Orai1 has been

Page 23: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

8

identified as the main pore-forming subunit of the SOCCs (Guibert et al., 2008; Muik et al.,

2011).

Stretch-activated channels (SACs) are activated by mechanical stimulation of smooth

muscle cells such as compression, stretch and shear stress. SACs are mechano-transducers that

convert physical forces into biological signals and cell responses. Cation influx through SACs

induces membrane depolarization that subsequently activates VGCCs to increase [Ca2+]i (Yin &

Kuebler, 2010). Several types of transient receptor potential (TRP) channels are expressed in

smooth muscle cells and have been linked to SAC activity (Gonzales et al., 2010b; Gonzalez-

Cobos & Trebak, 2010; Ramsey et al., 2006; Yang et al., 2006). TRP channels facilitate Ca2+

and Na+ influx into the cells, raise [Ca2+]i and [Na+]i respectively, and depolarize the plasma

membrane. The superfamily of TRP channels can be divided into several families and

subfamilies: three main subfamilies, TRPC (canonical), TRPV (vanilloid) and TRPM

(melastatin), and several subfamilies including TRPP (polycystin), TRPML (mucolipin), TRPA

(ankyrin) and TRPN. TRPC, TRPM, and TRPV display mechanosensitive properties while

TRPN channels do not (Earley, 2010; Gonzalez-Cobos & Trebak, 2010; Muraki et al., 2003;

Ramsey et al., 2006; Yang et al., 2006; Yin & Kuebler, 2010). These channels exhibit different

relative permeabilities to cations (Ca2+, Na+, K+, Mg2+, Mn2+, Cs+, Li+). While all TRP channels

are permeable to cations, only two TRP channels are impermeable to Ca2+ (TRPM4 and

TRPM5), and two others are highly Ca2+ permeable (TRPV5 and TRPV6; Owsianik et al.,

2006). TRP channel pharmacology is complex, and is different in different tissues and species

(Earley et al., 2004; Ramsey et al., 2006; Welsh et al., 2002).

Page 24: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

9

1.3.2 Voltage-gated Ca2+ channels (VGCCs).

Voltage-gated Ca2+ channels are expressed in many tissues and cells, e.g. the central and

peripheral nervous system and smooth muscle cells, where they demonstrate a wide range of

physiological, pharmacological and electrophysiological properties. Many different factors have

been identified that control VGCCs activity. These include alternate splicing, auxiliary subunit

associations, peptide and small organic blockers, GPCRs, protein kinases and Ca2+-binding

proteins (Braunstein et al., 2009; Catterall, 2000; Hofmann et al., 1999). There are several

different kinds of VGCCs that have different physiological functions and electrophysiological

and pharmacological properties. VGCCs include: 1) the CaV1 group (L-type), which has four

members (CaV1.1-1.4); 2) the CaV2 group, which has three members: CaV2.1 (P/Q-type), CaV2.2

(N-type), CaV2.3 (R-type); and 3) the CaV3 group (T type), which has three members (CaV3.1-

3.3). The CaV1 and CaV2 channels are classified as high-voltage-activated (HVA) channels, due

to their high voltage thresholds of activation, while the CaV3 channels are classified as low-

voltage-activated (LVA) channels (Catterall, 2000; Cheng et al., 2009).

1.3.3 General structure of VGCCs.

VGCCs are hetero-oligomeric protein complexes that have a main pore-forming α1-

subunit of 190–250 kDa that is in association with auxiliary β-, α2δ- and γ-subunits. The α1-

subunit has four membrane-spanning domains (I–IV) that are linked together in a single

polypeptide chain. Each domain contains six transmembrane segments (S1–S6); S4 is the voltage

sensor; a pore-forming loop (P-loop) is formed between S5 and S6. The C- and N-termini are

both intracellular (Figure 3; Catterall, 2000; Altier & Zamponi, 2008). The selectivity filter

Page 25: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

10

consensus sequence for the Ca2+ pore-forming loop has four glutamic acid repeats (EEEE). The

cytoplasmic loops connecting the four domains are structurally essential for interaction with the

β-subunits and membrane-binding proteins. They are also essential for channel gating and

interaction with second messengers. An auxiliary β-subunit and a transmembrane disulphide-

linked α2δ subunit complex are important components of most types of VGCCs. They are

required for optimal surface expression, channel regulation and gating kinetics (Catterall, 2000;

Hofmann et al., 1999; Marcantoni et al., 2008).

1.3.4 L-type channel structure.

In addition to the pore-forming α1-subunit of CaV1.2 in vascular smooth muscle cells

(VSMCs), other auxiliary subunits have been identified. There are four β subunit isoforms (β1b,

β2, β3, β4), and four α2δ subunit isoforms. The β and α2δ subunits directly interact with the α1-

subunit, the β-subunit modulates activation and inactivation of the channel, while the α2δ-subunit

is necessary for plasma membrane expression of CaV1.2 (Bielefeldt, 1999). There are a number

of splice variants of the α1-subunit of CaV1.2 channels that have been extensively investigated,

and a numbers of diseases have been linked to aberrant splicing. Previous work indicated that the

gene encoding the α1 subunit of CaV1.2 consists of 55 exons and that at least 19 exons undergo

alternative splicing (Liao et al., 2005; Tang et al., 2004).

Page 26: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

11

Figure 3: The topology of the high voltage-gated calcium channels. The pore-forming α1-

subunit is composed of four domains (I-IV), each consisting of six transmembrane segments (S1-

S6). Domain S4 contains the voltage sensor (labeled +). S5 and S6 segments form the pore of the

channel. β, α2-δ, and γ auxiliary subunits are also shown. The N- and C-termini of the channel

are intracellular (Altier & Zamponi, 2008).

Page 27: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

12

With the recent discovery of exon 1c at the N-terminus (Cheng et al., 2009), the CaV1.2

gene now is known to contain at least 56 exons of which 20 exons are subject to alternative

splicing. Based on splicing, CaV1.2 channels can be generally segregated into cardiac or smooth

muscle isoforms (Liao et al., 2005; Liao et al., 2007; Liao et al., 2004).

L-type channels are high-voltage-activated (HVA) Ca2+ channels, that activate at voltages

much more positive than the resting membrane potential. Their activation is fast and voltage

dependent, while inactivation is slow (> 1s), but it is more rapid in the presence of Ca2+ and

comparatively slow in the presence of Ba2+ (Catterall, 2000). Deactivation is also quick, ensuring

that the channels close rapidly when the membrane is repolarized to resting levels (˂ 1 ms;

Catterall, 2000). CaV1.2 is two to three times more permeable to Ba2+ than Ca2+, and lacks Ca2+-

dependent inactivation when Ba2+ is used as the charge carrier in electrophysiological recording

solutions (Nikitina et al., 2007). In most studies of vascular smooth muscle, the high Ba2+

concentrations that were used (10-20 mM, compared to the use of a physiological Ca2+

concentration of ~1.8 mM) induced a hyperpolarized shift in the current-voltage (IV) relationship

of CaV1.2 of ~10 mV. This occurs because Ba2+ obstructs the influx of ions. This is referred to

as surface charge screening and is due to the binding of positive ions to the membrane, on a

selective binding site on the channel protein (Kuo et al., 2010; Salemme et al., 2007; Li et al.,

2010). L-type channels are distinct from other CaV channels in their high sensitivity to 1, 4-

dihydropyridines (DHPs). Their blockage by DHPs occurs in a state-dependent manner as a

result of the binding of DHPs to inactivated channels and increases with membrane

depolarization (Dolphin, 2006). Previous studies also showed that the IIIS5, IIIS6 and IVS6

Page 28: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

13

domains determine the binding of DHPs to the α1-subunit of L-type Ca2+ channels (Liao et al.,

2007).

1.3.5 T-type channel structure.

The structure of the α1-subunit of T-type channels is related to the structure of the α1-

subunit of CaV1.2 channels. Unlike other VGCCs, T-type channels (CaV3.x) do not require

auxiliary subunits for functional expression (Lambert et al., 1997; Leuranguer et al., 1998), and,

to date, the existence of auxiliary subunits (β, α2δ, γ) with CaV3 α1-subunits in vascular smooth

muscle cells has not been reported. The α1-subunit of CaV3.1 consists of 38 exons (Mittman et

al., 1999), and a recent study showed that at least 15 sites undergo alternative splicing in human

brain. The α1-subunit of the CaV3.2 gene consists of 36 exons, and a more recent study revealed

over 12 alternative splicing sites (Zhong et al., 2006). Almost all CaV3 splice variants have been

found in non-smooth muscle cells (Shcheglovitov et al., 2008; Zhuang et al., 2000). As noted

above, the selectivity filter consensus sequence for the pore loops of CaV1.x and CaV2.x has four

repeats of glutamic acid (EEEE), whereas, in the CaV3.x pore loop, two glutamates are replaced

by aspartic acid residues (EEDD). The difference in the amino acid composition of the repeat

segments in the pore loops of L- and T-type channels is thought to account for the differences in

Ca2+ permeability of the channels (Jan & Jan, 1990; Talavera et al., 2001). L-and T-type Ca2+

channels have similar overall sizes, with long loops connecting membrane spanning domains I

and II as well as II and III, and short loops connecting membrane spanning domains III and IV.

In T-type Ca2+ channels, the functions of the loops are unknown, but may contain essential

regulatory sites (e.g. for protein kinases); this is in contrast to HVA channels in which these

Page 29: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

14

loops are essential for binding auxiliary subunits (Marcantoni et al., 2008; Perez-Reyes, 2003).

Short amino acid sequences connect to the N-terminus, while long amino acid sequences connect

to the C-terminus of T-type channels. T-type Ca2+ channels differ from other voltage-gated Ca2+

channels in their low threshold for activation and inactivation. These channels are characterized

by their activation at relatively hyperpolarized potentials, their fast activation and inactivation

and their slow deactivation. Unlike CaV1.2 channels, they have Ca2+-independent inactivation

kinetics (Perez-Reyes, 2003). T-type Ca2+ channels are almost equally permeable to Ca2+ and

Ba2+, thus, the use of Ba2+ as a charge carrier helps to differentiate between L- and T-type Ca2+

channels when both are expressed in the same cell type (Kuo et al., 2010; Nikitina et al., 2007).

High Ba2+ concentrations (10-20 mM) result in a hyperpolarized shift in voltage-dependence of

T-type channels (Perez-Reyes, 2003). Generally, T-type Ca2+ channels are characterized by their

greater sensitivity to agents such mibefradil, kurtoxin, R(-)efonidipine and Ni2+ (Lee et al., 1999;

Marcantoni et al., 2008; Perez-Reyes, 2003).

1.3.6 General functions of VGCCs.

VGCCs play a key role in mediating many cellular functions. Like other ion channels,

they contribute not only to shaping the action potential and to oscillations of membrane voltage,

but they also control Ca2+ influx through the plasma membrane and are involved in excitation–

secretion, excitation–contraction and excitation-transcription coupling (Dolphin, 2006;

Marcantoni et al., 2008).

Page 30: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

15

1.3.7 L-type channel (CaV1.2) function.

In cardiac myocytes, Ca2+ influx through L-type Ca2+ channels shapes the plateau phase

of the cardiac action potential. Ca2+ influx triggers Ca2+ release from SR stores via ryanodine

receptors (RyR), a process known as Ca2+-induced Ca2+-release, CICR. L-type Ca2+ channels can

mediate global increases in [Ca2+]i in smooth muscle cells (Lambert et al., 1997; Moosmang et

al., 2003). In vascular smooth muscle, L-type Ca2+ channels (CaV1.2) play an important role in

the development of myogenic tone. Pressurization of cerebral arteries leads to graded membrane

depolarization and opening of CaV1.2 channels. This causes further depolarization and steady

state Ca2+ entry due to the slow inactivation of the channels (Lambert et al., 1997; Knot &

Nelson, 1998; Moosmang et al., 2003). The high selectivity of DHPs for CaV1.2 L-type channels

(Table1) over CaV3 channels has allowed the functional role of CaV1.2 in vascular smooth

muscle cells to be studied. DHPs cause arterial dilation in a state-dependent manner (Dolphin,

2006; Perez-Reyes, 2003) supporting the conclusion that CaV1.2 channels are the major source of

Ca2+ entry in response to membrane potential depolarization. Thus they are key players in the

development of myogenic tone, which is essential for proper smooth muscle contraction and

function (Knot & Nelson, 1998; Moosmang et al., 2003).

1.3.8 T-type channel (CaV3.1 and CaV3.2) function.

The T-type Ca2+ channels produce the pacemaker potential in the SA node of the heart.

Moreover, in the central nervous system (CNS), T-type Ca2+ channels contribute to tonic bursting

activity in the thalamus (Dolphin, 2006; Perez-Reyes, 2003). Their role in vascular smooth

muscle has not yet been completely identified. Recent studies have provided evidence for the

Page 31: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

16

presence of T-type Ca2+ channels (CaV3.1 and CaV3.2) in resistance arteries (e.g. mesenteric

arteries and renal microcirculation; Abd El-Rahman et al., 2013; Jensen et al., 2004; Kuo et al.,

2010; Navarro-Gonzalez et al., 2009) and suggest that Ca2+ influx through T-type Ca2+ channels

could contribute to a global rise in [Ca2+]i and the development of myogenic tone, especially

when vessels are hyperpolarized. However, experimental support for this is limited and

incomplete (Kuo et al., 2010; Nikitina et al., 2007). Meanwhile the presence of two subtypes of

T-type Ca2+ channels in vascular smooth muscle (CaV3.1, and CaV3.2) raises the question: in

smooth muscle cells do both subtypes function in the same way? Published research regarding T-

type channel function suggests multiple roles for T-type channels. Some results suggest that Ca2+

influx through T-type channels facilitates arterial constriction and other researchers propose that

T-type channels may mediate dilation (Braunstein et al., 2009; Chen et al., 2003; Hansen et al.,

2011; Kuo et al., 2010; Poulsen et al., 2011). The latter possibility is important to investigate

since CaV3.2 knockout (α13.2-/-) mice showed a constricted arterial phenotype (Chen et al.,

2003). This study suggested that Ca2+ influx through CaV3.2 channels activates a Ca2+-dependent

K+ conductance (BKCa) that induces hyperpolarization and, therefore, vasodilation. This

suggested a possible mechanism in which VGCCs induce K+ channel activity and cause

vasodilation in smooth muscle cells. Such a mechanism needs to be extensively examined. At the

present time Ni2+ is the best pharmacological tool available to distinguish between Cav3.1 and

CaV3.2. Ni2+ blocks both subtypes of T-type channels, however, the IC50 of Ni2+ for CaV3.1 is

250 µM and is 12 µM for CaV3.2 (Table1; Lee et al., 1999; Perez-Reyes, 2003); thus a low

concentration of Ni2+ can be used to selectively inhibit the CaV3.2 channel subtype.

Page 32: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

17

VGCCs

Inhibitor

Channel

IC50

Cell type

Reference

Nifedipine

CaV1.2

CaV3.1

CaV3.2

10-15 nM

3 µM

1.2 µM

CHO

HEK293

HEK

Morel et al. 1998

Akaike et al. 1989

Lee et al. 2006

Mibefradil

CaV1.2

CaV3.1

CaV3.2

3 µM

0.012-0.014 µM

0.069 µM

HEK

HEK293

HEK293

Jimenez et al. 2000

Morita et al. 2002

Martin et al. 2000

R(-)efonidipine

CaV1.2

CaV3.1

CaV3.2

1000 µM

0.1 µM

0.1 µM

BHK

BHK

BHK

Furukawa et al. 2004

Furukawa et al. 2004

Furukawa et al. 2004

Kurtoxin

CaV1.2

CaV3.1

CaV3.2

>10 µM

0.07 µM

0.07 µM

Xenopus oocytes

Cardiomyocytes

Cardiomyocytes

Chuang et al. 1998

Horiba et al. 2008

Horiba et al. 2008

Ni2+

CaV1.2

CaV3.1

CaV3.2

324 µM

250 µM

12 µM

rbU-SMC

HEK293

HEK293

Bradley et al. 2004

Lee et al. 1999

Lee et al. 1999

Table 1. IC50 values for inhibition of VGCCs by various blockers.

Note: VGCCs, voltage-gated Ca2+ channels; CHO, Chinese hamster ovary; HEK, human

embryonic kidney; BHK, baby hamster kidney; rbU, rabbit urethra.

Page 33: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

18

1.4 Potassium channels in smooth muscle.

In general, potassium (K+) channels are comprised of a pore-forming α-subunit with

cytoplasmic N- and C-termini and contain pore-forming six transmembrane domains (S1–S6).

The S4 domain contains the voltage sensor (Korovkina & England, 2002). Each α-subunit is

associated with auxiliary β-subunits, which influence the biophysical characteristics of the

channel (Bahring et al., 2001). In BKCa channels the α-subunits contain an additional seventh

transmembrane region (S0) at the exoplasmic N-terminus (Tanaka et al., 2004). Potassium

channels are the main channels involved in mediating hyperpolarizing responses and the

induction of relaxation in smooth muscle cells (Ko et al., 2008). In vascular smooth muscle there

are four general types of K+ channels: large conductance Ca2+-activated (BKCa), voltage-gated

(KV), inward rectifier (KIR) and ATP-sensitive (KATP) K+ channels. Although hyperpolarization

is induced by all K+ channels, each K+ channel is regulated differently (Nelson and Quayle,

1995). BKCa channels are both voltage and Ca2+ sensitive (Cox & Petrou, 1999; Nelson &

Quayle, 1995). These channels are involved in inducing vasodilation through a negative

feedback mechanism (Jaggar et al., 1998b; Knot & Nelson, 1995). KV channels open in response

to membrane depolarization and allow K+ efflux, repolarization and a return to the resting

membrane potential. The function of KV channels, therefore, is to limit membrane depolarization

and maintain resting vascular tone (Korovkina & England, 2002). KIR channel activity increases

with membrane hyperpolarization, in contrast to KV and BKCa channels, and it is thought to play

a role in determining the resting membrane potential of vascular smooth muscle (Ko et al.,

2008). KATP channel current is linear in nature, and channel gating is voltage independent; it

Page 34: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

19

links cell metabolism to electrical activity of the plasma membrane by controlling membrane

potential (Thorneloe & Nelson, 2005; Teramoto, 2006; Ashcroft, 2005).

1.5 Sarcoplasmic reticulum (SR).

The SR in smooth muscle is the major intracellular Ca2+ store and plays a key role in

regulating Ca2+ homeostasis in smooth muscle cells. When the SR is close to the plasma

membrane it is defined as superficial or peripheral and when it is positioned away from the

plasma membrane it is described as deep or central. Ca2+ homeostasis, local Ca2+ release, and

interactions with plasma membrane ion channels, and hence cell excitability, have been

associated with the peripheral SR (Wray & Burdyga, 2010). It has been proposed that the central

SR is involved in providing Ca2+ to the myofilaments to induce contraction (Wray & Burdyga,

2010). The SR elicits relaxation when Ca2+ is sequestered into the SR by Ca2+-ATPase (SERCA)

activity, or constriction through the release of Ca2+. Two major Ca2+ channels are found in the

SR membrane, IP3R and RyR, which contribute to Ca2+ homeostasis, and are involved in

generation of a variety of Ca2+ signals (e.g. Ca2+ waves, sparks and puffs; Wray & Burdyga,

2010) .

1.5.1 Inositol 1, 4, 5-trisphosphate receptors (IP3R).

In addition to Ca2+, IP3Rs require IP3 for activation and SR Ca2+ release. IP3 is produced

by the activation of Gq/11- coupled GPCRs in the plasma membrane. IP3R regulation is complex

as both Ca2+ and IP3 can activate and/or inhibit IP3Rs (Bootman et al., 1995; Chalmers et al.,

2007; Sanders, 2001). There are three isoforms of the IP3R (IP3R1, IP3R2 and IP3R3; Furuichi et

Page 35: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

20

al., 1989). The IP3R protein has a molecular weight of ~310 kDa and displays three distinctive

regions: 1) the N-terminal ligand-binding region, with a suppressor domain and IP3-binding core;

2) the C-terminal channel-forming region with six transmembrane helices and a pore loop

between TM5 and TM6; and 3) the regulatory region connecting the N- and C-termini (Furuichi

et al., 1989). In vascular smooth muscle, IP3Rs are the main SR Ca2+ release channels. When

GPCRs are activated, IP3 and DAG are generated from PLC-mediated cleavage of PIP2; IP3

opens IP3R channels (Jaggar et al., 2000). Furthermore, Ca2+ release events have been observed

from single IP3Rs and named Ca2+ “blips” (Amberg et al., 2007; Lee et al., 2002; Taylor et al.,

2008). Similar to Ca2+ sparks, produced from RyRs, Ca2+ puffs are generated by the concerted

opening of a cluster of IP3Rs (Furuichi et al., 1989; Mironneau et al., 1996). Ca2+ puffs may

combine with Ca2+ sparks to generate oscillations which propagate as Ca2+ waves through the

cell (Mironneau et al., 1996).

1.5.2 Ryanodine receptors (RyR).

Molecular and biochemical analyses have identified three different RyR isoforms (RyR1,

RyR2 and RyR3), encoded by three different genes on different chromosomes (Coronado et al.,

1994; Fill & Copello, 2002). Expression of RyR1, RyR2 and RyR3 isoforms are predominant in

skeletal muscle, cardiac muscle and brain, respectively (Coronado et al., 1994; Fill & Copello,

2002). All three isoforms are expressed in vascular smooth muscle cells, where RyR2 is the

predominant isoform (Chalmers et al., 2007; Nelson et al., 1995; Wellman & Nelson, 2003).

RyRs are the largest known ion channel proteins (~560 kDa/subunit) and form homotetramers

(Chalmers et al., 2007; Fill & Copello, 2002). The most widely accepted model of the RyR

Page 36: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

21

describes the RyR channel as an assembly of 6 transmembrane (TM) domains, in which TM5

and TM6 form the pore region with a pore loop having a conserved sequence motif,

GXRXGGGXGD, that contains a ryanodine-binding site and a selectivity filter (Takeshima et

al., 1989; Zhao et al., 1999). About 80% of the mass of the RyR is in the cytosolic domain,

which contains a number of protein-binding sites including sites for calmodulin (CaM) and

protein kinase A (PKA) binding (Bers, 2004). [Ca2+]i has a biphasic effect on RyR activity, it

activates RyRs at micromolar [Ca2+]i (~10 μM) and inhibits RyRs at higher [Ca2+]i (>100 μM;

Chalmers et al., 2007). It has been proposed that upon membrane depolarization in vascular

smooth muscle, Ca2+ enters the cytosol via VGCCs and activates RyR channels by a Ca2+-

induced Ca2+ release (CICR) process, in a manner similar to that described in cardiac muscle

(Chalmers et al., 2007; Sanders, 2001). Ca2+ efflux from activated IP3Rs may also diffuse and

trigger Ca2+ release from neighboring RyRs through CICR (Chalmers et al., 2007; Sanders,

2001). SR Ca2+ through CICR has been proposed as a mechanism for eliciting contraction in

vascular smooth muscle cells (Jaggar et al., 1998b; Nelson et al., 1995). The discovery of local

Ca2+ signaling revealed another mechanism through which RyRs can affect vascular

contractility. Ca2+ sparks constitute one of the main local events that involve Ca2+ release

through RyRs (Nelson et al., 1995; Perez et al., 2001). Ca2+ sparks have been suggested to

activate BKCa channels to elicit hyperpolarization and prevent the vessel from becoming over-

constricted. Thus RyRs may be part of a negative feedback mechanism to control vessel diameter

(Nelson et al., 1995; Perez et al., 2001; Wellman & Nelson, 2003).

Page 37: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

22

1.5.3 Ca2+ sparks.

Ca2+ sparks are local Ca2+ events caused by openings of RyRs. They were initially

identified in cardiac muscle and subsequently in skeletal and smooth muscle (Jaggar et al., 2000;

Nelson et al., 1995). Improved detection of Ca2+ sparks has allowed these events to be

characterized as having a rise time of ~20 ms, a half-time of decay of 50-60 ms, a spatial spread

of 2.4 μm and a frequency of 1 Hz (Nelson et al., 1995; Perez et al., 2001; Perez et al., 1999).

Modulation of Ca2+ sparks by ryanodine and caffeine indicates that RyR openings are involved

in these events (Knot et al., 1998; Mironneau et al., 1996; Nelson et al., 1995; Perez et al., 2001;

Perez et al., 1999). It is clear in cardiac muscle that the influx of Ca2+ via VGCCs plays a key

role in the generation of Ca2+ sparks via CICR. Evidence for a similar mechanism in vascular

smooth muscle is less clear (Chalmers et al., 2007). In vascular smooth muscle, several studies

indicate that Ca2+ sparks can occur for a period of time in the presence of L-type Ca2+ channel

blockers or in the absence of extracellular Ca2+ (Nelson et al., 1995; Zhao et al., 1999).

However, Ca2+ spark activity can be modulated by L-type channel activity. In intact cerebral

arterial segments, Ca2+ spark frequency and amplitude were seen to be augmented by high

extracellular [K+], suggesting that Ca2+ influx via VGCCs can directly or indirectly regulate Ca2+

sparks (Jaggar et al., 1998a). Ca2+ influx through VGCC could also be important for refilling the

SR store without having a direct effect on the RyR channels. Ca2+ sparks can be generated at a

low frequency in un-stimulated smooth muscle cells, and can activate BKCa channels to produce

a spontaneous transient outward current (STOC; Knot et al., 1998; Nelson et al., 1995; Perez et

al., 1999). In pressurized rat cerebral arteries at physiological membrane potential of -40 mV,

simultaneous measurements of Ca2+ sparks and STOCs indicated that most Ca2+ sparks are

Page 38: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

23

associated with the initiation of STOCs (Perez et al., 1999). This association occurs despite the

fact that both Ca2+ spark and STOC amplitudes vary (Perez et al., 1999). Although a rise in

[Ca2+] is normally associated with constriction of vascular smooth muscle Ca2+ sparks and

STOCs cause relaxation (Knot et al., 1998). Ca2+ sparks can, therefore, regulate myogenic tone

by activating BKCa channels, which in turn regulates membrane potential, leading to the

inhibition of L-type Ca2+ channels. This limits Ca2+ entry and reduces contraction. Thus, RyR-

mediated Ca2+ sparks control vascular tone through a feedback mechanism that promotes

vasodilation.

1.6 Summary.

L-type Ca2+ channels were the first VGCCs identified in vascular smooth muscle cells.

Subsequently other voltage gated Ca2+ channel types (especially T-type Ca2+ channels) were also

identified. Different functional roles of the T-type Ca2+ channel subtypes have been proposed in

smooth muscle cells, either augmenting the myogenic response by increasing global [Ca2+] or

reducing it through a negative feedback mechanism. In the latter case, it has been suggested that

T-type Ca2+ channels regulate BKCa activity to elicit vessel dilation, however, this has not been

well established. The work described below was designed to identify the T-type Ca2+ channels

present in cerebral arterial smooth muscle cells and to understand their functional role in the cells

especially with regard to tone development. Identification of the different types of VGCCs

expressed and their electrophysiological characteristics and functional roles in cerebral arterial

smooth muscles, will improve our understanding of Ca2+ homoeostasis and of tone in these

arteries.

Page 39: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

24

1.7 Hypothesis.

Specific L- and T-type Ca2+ channels are expressed in cerebral arterial smooth muscle

and contribute to myogenic tone development and control of localized Ca2+ signaling.

A variety of techniques were employed to test the hypothesis, including molecular, biochemical,

electrophysiological, functional, structural, and computational modeling approaches. The

following two principal objectives were used to guide experiments designed to test this

hypothesis:

1) To identify the different subtypes of voltage-gated Ca2+ channels that are expressed

in cerebral vascular smooth muscle and determine their roles in myogenic tone

development.

2) To determine the functional role of T-type Ca2+ channels, specifically CaV3.2, in

cerebral vascular smooth muscle.

Page 40: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

25

Chapter Two: Identification of L- and T-type Ca2+ Channels in Rat Cerebral Arteries:

Role in Myogenic Tone Development

Rasha R. Abd El-Rahman, Osama F. Harraz, Suzanne E. Brett, Yana Anfinogenova, Rania E.

Mufti, Daniel Goldman, and Donald G. Welsh

This work was published in Am J Physiol Heart Circ Physiol 304: 58–71, 2013.

I was responsible for the biochemical and functional data collection and analysis, and I wrote

the drafts of the manuscript. Yana Anfinogenova and Osama Harraz carried out the patch clamp

experiments. Daniel Goldman performed the blood flow modeling.

Page 41: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

26

2.1 INTRODUCTION

Arterial smooth muscle cells actively respond to vasoactive stimuli, altering arterial

diameter and tissue blood flow. Changes in vessel tone are the result of myosin light chain

(MLC20) phosphorylation, a process dynamically controlled by Ca2+ influx from the extracellular

space (Deng et al., 2002). Voltage-gated Ca2+ channels are the principal conductances that

regulate extracellular Ca2+ influx. These membrane channels are hetero-oligomeric complexes

comprised of a pore-forming α1-subunit and accessory proteins that influence gating

characteristics and protein trafficking (Kisilevsky & Zamponi, 2008). The α1-subunit is

composed of four domains, each of which contains six transmembrane segments, an S4 voltage

sensor and a P loop that confers ion selectivity (Jan & Jan, 1990; Talavera et al., 2001).

Molecular studies have identified three classes of α1-subunits (CaV1-3), and within each category

there are several subtypes. CaV1/CaV2 subunits display electrical properties characteristic of high

voltage-activated Ca2+ channels (i.e. L-, P/Q-, N- & R- types; Catterall, 1995). In contrast, CaV3

subunits encode for Ca2+ channels activated by lower voltages (i.e. T-type; Iftinca & Zamponi,

2009; Marcantoni et al., 2008).

CaV1.2 is a key pore-forming subunit of L-type Ca2+ channels in cerebral arterial smooth

muscle. There are at least 3 splice variants and, upon depolarization, its graded activation plays a

central role in setting cytosolic [Ca2+] (Cheng et al., 2009; Liao et al., 2007). L-type Ca2+

channel activation is important for tone development; channel blockade does not eliminate all

arterial responsiveness. Case-in-point is the work of Mufti et al. (Mufti et al., 2010) who

showed that ~20% of the myogenic response was insensitive to diltiazem, an L-type Ca2+

channel inhibitor. Such observations have fostered the view that there are additional Ca2+ influx

Page 42: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

27

pathways that help set cytosolic [Ca2+] and thus the level of myogenic tone (Dimopoulos et al.,

2007; Walsh et al., 1995). In this context, recent studies have noted that CaV3.1 and CaV3.2, α1-

subunits of T-type Ca2+ channels, are present in resistance arteries (Jensen et al., 2004; Nikitina

et al., 2007). T-type Ca2+ channels activate ~10-15 mV negative to L-type Ca2+ channels and

thus are positioned to regulate Ca2+ influx at more hyperpolarized potentials (Perez-Reyes,

2003). While Ca2+ influx through T-type Ca2+ channels could facilitate myogenic tone

development, experimental findings are limited and incomplete (Navarro-Gonzalez et al., 2009).

The present study determined which voltage-gated Ca2+ channels are expressed in rat

cerebral arteries and what role they play in myogenic tone development. To meet this objective,

we combined the experimental strengths of molecular biology, patch clamp electrophysiology,

vessel myography and computational modelling. Our RT-PCR and western blot analysis revealed

that the α1-subunits of CaV1.2 (L-type) and CaV3.1/3.2 (T-type) were expressed at the mRNA

and protein level in isolated smooth muscle cells and whole cerebral arteries, respectively.

Consistent with these observations, patch clamp electrophysiology revealed the presence of two

distinct Ba2+ currents, with electrical and pharmacological fingerprints characteristic of L- or T-

type Ca2+ channels. Functionally, L-type Ca2+ channel inhibition attenuated myogenic tone

development, particularly at higher pressures where arteries are depolarized. In comparison, the

contribution of T-type Ca2+ channels was more limited and best observed at lower pressure in

hyperpolarized vessels. While their contribution to tone development is limited, computational

modeling indicates that T-type Ca2+ channel activity could possibly alter resting blood flow by

20-50%. In closing, our findings demonstrate that α1-subunits of L- and T-type Ca2+ channels

are both expressed in arteries and that they encode for a distinct voltage-dependent conductance.

Page 43: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

28

They further reveal that each current contributes, albeit unequally, to the maintenance of

myogenic tone, a key physiological response in the cerebral circulation.

2.2 MATERIALS AND METHODS

2.2.1 Animal procedures.

Animal procedures were approved by the Animal Care and Use Committee at the

University of Calgary. Briefly, female Sprague-Dawley rats (10-12 weeks of age) were

euthanized via carbon dioxide asphyxiation. The brain was carefully removed and placed in cold

phosphate-buffered (pH 7.4) saline solution containing (in mM): 138 NaCl, 3 KCl, 10 Na2HPO4,

2 NaH2PO4, 5 glucose, 0.1 CaCl2, and 0.1 MgSO4. Middle and posterior cerebral arteries were

carefully dissected out of surrounding tissue and cut into 2-3 mm segments.

2.2.2 Vessel myography.

Arterial segments were mounted in a customized arteriograph and superfused with warm

(37 °C) physiological salt solution (PSS; pH 7.4; 21% O2, 5% CO2, balance N2) containing (in

mM): 119 NaCl, 4.7 KCl, 20 NaHCO3, 1.1 KH2PO4, 1.2 MgSO4, 1.6 CaCl2, and 10 glucose (Kuo

et al., 1991; Welsh et al., 2000). To limit the tonic dilatory influence of the endothelium on

myogenic tone development (Knot et al., 1999; Kuo et al., 1991), these cells were removed by

passing air bubbles through the vessel lumen (1 to 2 min); successful removal was confirmed by

the loss of bradykinin-induced dilation. Arteries were equilibrated for 30 min at 15 mmHg and

contractile responsiveness assessed by briefly exposing (∼10 s) arteries to 60 mM KCl. After

equilibration, intravascular pressure was increased incrementally from 15 to 20, or 80 mmHg.

Page 44: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

29

Vessels were then subjected to one of three experimental protocols: 1) nifedipine (50 and 300

nM) + mibefradil (1 and 5 µM), 2) nifedipine + kurtoxin (1 µM), or 3) nifedipine + efonidipine

(3 µM). Each protocol was followed by the addition of a Ca2+-free physiological salt solution (0

externally added Ca2+ + 2 mM EGTA). In one additional set of experiments, arterial tone was

sequentially measured at 20, 40, 60, 80, and 100 mmHg under control conditions and in the

presence of nifedipine (300 nM) ± mibefradil (5 µM). Arterial diameter was monitored using an

automated edge detection system (IonOptix).

2.2.3 Isolation of arterial smooth muscle cells.

Smooth muscle cells from middle and posterior cerebral arteries were enzymatically

isolated as described previously (Luykenaar et al., 2004; Talavera et al., 2001). Briefly, arterial

segments were placed in an isolation medium (37 °C, 10 min) containing (in mM): 60 NaCl, 80

Na+ glutamate, 5 KCl, 2 MgCl2, 10 glucose, and 10 HEPES with 1 mg/ml albumin (pH 7.4).

Vessels were then exposed to a two-step digestion process that involved: 1) a 15-min incubation

in isolation medium (37 °C) containing 0.6 mg/ml papain and 1.8 mg/ml dithioerythritol and 2) a

15-min incubation in isolation medium containing 100 µM Ca2+, 0.7 mg/ml type F collagenase

and 0.4 mg/ml type H collagenase. After treatment, tissues were washed repeatedly with ice-cold

isolation medium and triturated with a fire-polished pipette. Liberated cells were stored in ice-

cold isolation medium for use the same day.

Page 45: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

30

2.2.4 PCR analysis.

Smooth muscle cells (~200) were isolated from middle and posterior cerebral arteries.

Individual cells were collected manually with a suction pipette and pooled together in RNase-

and DNase-free collection tubes to extract total RNA (RNeasy mini kit with DNAase treatment;

Qiagen, Valencia, CA). Total RNA was extracted (RNeasy mini kit with DNAase treatment;

Qiagen, Valencia, CA) and first-strand cDNA synthesized using the Sensi-script RT kit (Qiagen)

with oligo d(T) primer. Subsequently, a sample of total volume of 50 µl was prepared as follows:

2 µl of each first-strand cDNA reaction was used as templates in a PCR reaction containing 0.25

µM forward and reverse primers (see below), 0.2 mM deoxynucleoside triphosphates, and 2.5

µM recombinant Taq DNA polymerase. The PCR reaction protocol includes a denaturation step

at 92 °C for 5 min; 35 cycles of 92 °C for 45 s, 55° C for 60 s, and 72 °C for 2 min; and a final

extension step at 72 °C for 5 min. Forward and reverse specific primers (made by the University

of Calgary Core DNA Services) were as follows: CaV1.1 (400 bp), (forward) 5´-

TGCTGATCGTCATGCTCTTC-3´, (reverse) 5´-ATGGCCTTGAACTCATCCAG-3´; CaV1.2

(400 bp), (forward) 5´-GCCTCTTCACGGTGGAG-3´, (reverse) 5´-

TCCCAATCACTGCATAGATAA-3´; CaV1.3 (399 bp), (forward) 5´-

TCCGAAGAGCCTGCATTAGT-3´, (reverse) 5´-GCGGTCTTAACACTCGGAAG-3´;

CaV1.4 (500 bp), (forward) 5´-GCAAAGTGGCTCTTCAGGAC-3´, (reverse)

5´-GTTGAGGGGAACATTCATGG-3´; CaV2.1 (501 bp), (forward)

5´-CTGCTTTGAAGAGGGGACAG-3´, (reverse) 5´-CGAAGAGCTCCATCAAAAGG-

Page 46: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

31

3´; CaV2.2 (504 bp), (forward) 5´-TGTTCAGCAAGTGGCTTTTG-3´, (reverse) 5´-

TTTCAGGGGGAACAGACAAC-3´; CaV2.3 (502 bp), (forward) 5´-

TCCTGCATGACAACCAACAT-3´, (reverse) 5´-GGTCTCGGAAGTACGATCCA-3´;

CaV3.1 (501 bp), (forward) 5´-TCTTCCAGGACAGGTGGAAC-3´, (reverse) 5´-

TCAGCACGAAGGACACAAAG-3´; CaV3.2 (402 bp), (forward) 5´-

AGTTTCCTCTTTGGGGGCTA-3´, (reverse) 5´-CAGGAAAACCCAAACCTGAA-3´; CaV3.3

(501 bp), (forward) 5´-CCCTGGAGATGATCCTGAAA-3´, (reverse) 5´-

AGTTGCCAAAGGTCATGAGG-3´; β-actin (130 bp), (forward) 5´-

TATGAGGGTTACGCGCTCCC-3´, (reverse) 5´-ACGCTCGGTCAGGATCTTCA-3´.

After electrophoresis on a 1.5% agarose gel, the PCR products were excised, extracted, and

purified with a kit purchased from Qiagen (Valencia, CA). Sequencing was performed at the

University of Calgary Core DNA Services. Smooth muscle cell samples were screened for

template and endothelial cell contamination, as previously described (Wu et al., 2007).

2.2.5 Western blotting.

Cortex or cerebral arterial segments from two rat brains, or tissues from heart, retina, or

kidney arterioles were placed in 100 µl of lysis buffer (pH 7.4) containing (in mM): 150 NaCl, 1

CaCl2, 1 MgCl2, 10 HEPES, 0.5% Tween, and 10% mammalian protease inhibitor cocktail

(Sigma-Aldrich, St. Louis, MO). Samples were mechanically disrupted, exposed to three freeze-

thaw cycles, and then centrifuged (10 min; 13,000 rpm). Supernatant was placed in a clean tube,

assayed for total protein, and stored at -20 °C for up to 1 week. Samples were prepared for

electrophoresis by placing 15 µl of supernatant in 5 µl of 4X sample buffer plus 2 µl of DTT.

Page 47: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

32

After heating (10 min; 90 °C) was completed, 20 µg of protein were loaded on a 5.6%

polyacrylamide gel. Proteins were transferred to a polyvinylidene difluoride (PVDF) membrane

at 4 °C for 1 h; blotted membranes were initially washed (PBS, 5 min) and then exposed to a

0.5% glutaraldehyde PBS solution (45 min) to cross link and fix proteins. After a second set of

washes in Tris-buffered saline (TBS), the membrane was blocked with 1.0% ECL blocking agent

(GE Healthcare, Buckinghamshire, UK) in TBS containing 0.02% Tween-20 (TBST) for 1 h and

subsequently incubated overnight (4 °C) in a TBST (0.1%) solution containing a primary rabbit

antibody [CaV1.2 (1:200), CaV3.1 (1:200), CaV3.2 (1:200); Alomone Laboratories]. The next

morning, the membrane was washed (TBST, 0.02%), and incubated in a TBST (0.1%) solution

containing horseradish peroxidase-conjugated secondary antibody (HRP anti-rabbit antibody)

(1:10,000 dilution, Jackson Laboratories). The membrane was washed for a final time in TBST

(0.02%; 1 h, 20-22 °C) and HRP detected with the Amersham ECL Advance Western Blotting

Detection Kit (GE HealthCare). Emitted light was detected with a chemiluminescence image

analyzer and analyzed with Multi Gauge v3.0 software (Fujifilm, ON, CA).

2.2.6 Electrophysiology.

Conventional patch-clamp electrophysiology was used to measure whole-cell currents in

isolated smooth muscle cells. Briefly, recording electrodes (pipette resistance, 4–7 MΩ) were

fashioned from borosilicate glass and back-filled with pipette solution containing (in mM): 135

CsCl, 2.5 Mg-ATP, 10 HEPES, and 10 EGTA (pH 7.2). This pipette was then gently lowered

onto the cell and negative pressure applied to rupture the membrane. With series resistance ˂10

MΩ and input resistance >10 GΩ, cells were then voltage clamped (-60 mV) and equilibrated for

Page 48: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

33

5 min in a bath solution containing (in mM): 110 NaCl, 135 CsCl, 4 TEA-Cl, 1.2 MgCl2, 10

BaCl2, 10 HEPES, and 10 glucose (pH 7.4). The voltage protocol consisted of a prepulse to -90

or -60 mV (200 ms), followed by a series of voltage steps from -80 mV to +40 mV (200–300 ms,

10-mV increments). This protocol was run under control conditions and in the presence of

nifedipine (300 nM) ± mibefradil (5 µM), kurtoxin (5 µM), or efonidipine (3 µM). Whole-cell

currents were recorded on an Axopatch 200B amplifier (Axon Instruments), filtered at 1 kHz,

digitized at 5 kHz, and stored on a computer for subsequent analysis with Clampfit 10.2

software. Cell capacitance ranged between 14 and 18 pF and was measured with the cancellation

circuitry in the voltage-clamp amplifier. A 1M NaCl-agar salt bridge between the reference

electrode and the bath solution was used to minimize offset potentials (< 2 mV). All experiments

were performed at room temperature (20 –22 °C).

2.2.7 Computational blood flow modeling.

A previously developed theoretical model (Goldman & Popel, 2000; Pries et al., 1990;

Pries et al., 1994) was used to calculate two-phase [red blood cell (RBC) and plasma] steady-

state flow in the arteriolar networks considered. This model implements conservation of blood

and RBC volume flow at each node joining three arterial branches and includes known blood

rheology, specifically the Fahraeus and Fahraeus-Lindqvist effects in unbranched segments and

phase separation at diverging bifurcations. Besides geometric information, the model requires

specification of boundary conditions in the form of hematocrit at the inlet segment (upstream end

of the 1st-order posterior cerebral artery) and pressure changes between the inlet node and outlet

nodes (downstream ends of the 3rd-order posterior arteries). All flow simulations used an inlet

Page 49: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

34

hematocrit of 0.4, and inlet and outlet pressures of 75 mmHg and 50 mmHg, respectively.

Arterial diameter/length values were set according to functional measurements from posterior

cerebral vessels pressurized to 60 mmHg. Values were as follows: 1st order: 164 µm diameter,

3.6 mm length; 2nd order: 150 µm diameter, 3.5 mm length; and 3rd order: 125 µm diameter, 1.1

mm length.

2.2.8 Statistical analysis.

Data are expressed as means ± SE, and n indicates the number of vessels or cells. No

more than two experiments were performed on vessels from a given animal. Where appropriate,

paired, unpaired t-tests, and one way ANOVA were performed to compare the effects of a given

condition/treatment on arterial diameter, or whole cell current (see figure legends for specific

details). P values ≤ 0.05 were considered statistically significant.

2.2.9 Solutions and chemicals.

All buffers, chemicals, and reagents were purchased from Sigma-Aldrich unless

otherwise stated.

2.3 RESULTS

2.3.1 Ca2+ channel α1-subunit expression in middle and posterior cerebral arteries.

This study began with an mRNA characterization of CaV α1-subunit expression in whole

cerebral arteries and isolated smooth muscle cells. Primers were designed against rat sequences

and efficacy confirmed by screening a range of appropriate control tissues. Skeletal muscle was

Page 50: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

35

used as the control for CaV1.1 and brain tissue was used for CaV1.2 and CaV1.3. Retina was the

control tissue for CaV1.4. Brain tissue was used as control for all CaV2.x channels. Afferent

arterioles from kidney were used as control for CaV3.1 and CaV3.2. Cardiac tissue was the

control for CaV3.3 channel. All α1-subunit products were sequenced to confirm their identities.

With the use of whole cerebral arteries, this study confirmed that α1-subunits for high (CaV1.1,

CaV1.2, CaV1.3, and CaV2.2) and low (CaV3.1, CaV3.2, and CaV3.3) voltage-activated Ca2+

channels were observable at the mRNA level (Fig. 4B). This expression profile was reduced to

CaV1.2, CaV3.1, and CaV3.2 in isolated smooth muscle cells (Fig. 4C). Note that isolated smooth

muscle cells were pre-screened for endothelial contamination by using primers for ET-1; no

preps that screened positive for ET-1 were used. With the use of the mRNA observations as a

guide, western blot analysis was subsequently performed to assess protein expression. With the

use of brain as the positive control, this study confirmed that the α1-subunits of L-type (CaV1.2)

and T-type (CaV3.1 and CaV3.2) Ca2+ channels were present at the protein level in whole cerebral

arteries (Fig. 5).

2.3.2 Whole-cell Ba2+ currents in cerebral arterial smooth muscle cells.

Building on the preceding observations, we next ascertained whether whole-cell patch

clamp electrophysiology could delineate L- and T-type Ca2+ channels from one another.

Cerebral arterial smooth muscle cells were enzymatically isolated, and whole-cell currents were

monitored in the presence of 10 mM Ba2+ to accentuate charge flow through L-type Ca2+

channels. Stepping from a prepulse of -90 mV, voltage protocols first revealed an inward Ba2+

current that activated at ~ -30 mV and peaked at +20 mV (Fig. 6), however, when the prepulse

Page 51: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

36

was changed to -60 mV to facilitate T-type Ca2+ channel inactivation, no statistically significant

change in current was detected at any voltage. Further experiments showed that nifedipine,

applied at concentrations approximately five- to 30-fold higher than the IC50 value for the

smooth muscle variant of CaV1.2 (Jensen & Holstein-Rathlou, 2009), blocked ~75% of the

whole-cell Ba2+ current (Fig. 7, A and B). Plots in Fig. 7C indicate that the current-voltage

relationship of the nifedipine-insensitive current is leftward shifted in relation to the nifedipine-

sensitive component when 50 nM or 300 nM nifedipine was used. Note that a nifedipine

concentration of 300 nM is predicted, based on previous work, to block >98% of all L-type Ca2+

channels in this vascular preparation (Morel et al., 1998). Higher micromolar concentrations

were avoided since they interfere with CaV3.1 and CaV3.2 channels (Akaike et al., 1989; Lee et

al., 2006). Further work was performed to determine whether the nifedipine-insensitive current

reflected T-type Ca2+ channel activation. The experimental approach consisted of two elements.

The first centered on whether the nifedipine-insensitive component was blocked by T-type Ca2+

channel inhibitors including mibefradil, efonidipine, and kurtoxin. These agents were applied at

concentrations 10- to 20-fold above the IC50 value for CaV3.1 and CaV3.2 and should block >

95% of all T-type Ca2+ channels (Furukawa et al., 2004; Horiba et al., 2008; Martin et al., 2000;

Morita et al., 2002). The second involved an electrical analysis in which the time constants of

activation/inactivation (τactivation, τinactivation) were calculated for the nifedipine- and the T-type

Ca2+ channel inhibitor-sensitive components. Whole cell currents were measured using a

prepulse of -90 mV at a series of voltages from -40 mV to +40 mV. Currents were measured for

control and in the presence of 300 nM nifedipine. Currents were then measured after 5 µM

mibefradil was added to the bath. The nifedipine-sensitive current at each voltage was

Page 52: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

37

determined by subtracting the current in the presence of nifedipine from that measured for

control. The mibefradil-sensitive component at each voltage was determined by subtraction of

the current recorded in the presence of both nifedipine and mibefradil from that recorded in the

presence of nifedipine. Figure 8, A and B, indicate that the nifedipine-insensitive component of

the whole-cell Ba2+ current was largely blocked by 5 µM mibefradil, a T-type Ca2+ channel

blocker. The time constants for activation/inactivation, calculated at +10 to +20 mV, were lower

for the mibefradil-sensitive component than the nifedipine-sensitive component, indicating faster

activation and inactivation. This finding is consistent with mibefradil blocking activated T-type

Ca2+ channels (Fig. 8C).

Similar experiments were carried out using two other T-type channel blockers: kurtoxin

and efonidipine. Whole cell currents were measured using a prepulse of -90 mV at a series of

voltages from -40 mV to +40 mV. Currents were measured for control and in the presence of 300

nM nifedipine. Currents were then measured after 5 µM kurtoxin (Fig. 9, A and B) or 3 µM

efonidipine (Fig. 10, A and B) was added to the bath. The nifedipine-sensitive current at each

voltage was determined by subtracting the current in the presence of nifedipine from that

measured for control. The kurtoxin/efonidipine-sensitive components at each voltage were

determined by subtraction of the current recorded in the presence of both nifedipine and kurtoxin

or efonidipine from that recorded in the presence of nifedipine. The activation/inactivation time

constants were also smaller for the kurtoxin/efonidipine-sensitive than the nifedipine-sensitive

component of the whole-cell Ba2+ current (Figures 9C and 10C). Reverse-order experiments (i.e.

addition of the T-type channel blockers before nifedipine) were not performed since these T-type

Ca2+ channel inhibitors will block L-type Ca2+ channels when applied at micromolar

Page 53: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

38

concentrations (Jensen & Holstein-Rathlou, 2009). Note, in these longer recordings, a small,

linear current characteristic of leak was at times evident after the addition of T-type Ca2+ channel

blockers.

2.3.3 Ca2+ channel blockers and myogenic tone.

Having determined that L- and T-type Ca2+ channels are electrically present in cerebral

arterial smooth muscle cells, we next determined their role in myogenic tone development. The

approach consisted of the sequential application of L- and then T-type Ca2+ channel blockers to

middle/posterior cerebral arteries pressurized to 20 or 80 mmHg. At 20 mmHg, arterial

membrane potential will rest near -55 mV, whereas at 80 mmHg, tissues will depolarize to -35

mV (Lacinova, 2005; Smirnov & Aaronson, 1992). At these different voltages, Ca2+ influx

through L- and T-type Ca2+ channels should vary along with their relative contributions to the

myogenic response. At 20 mmHg, cerebral arteries displayed 43 ± 4 µm of tone (Fig. 11, A and

C). Change in diameter was determined by subtraction of the diameter measured for control (no

blockers presents) from the diameter measured in Ca2+ free solution. The addition of nifedipine,

50 nM and 300 nM, to the superfusate modestly attenuated tone development, as did 1 and 5 µM

mibefradil, a T-type Ca2+ channel blocker (see Fig. 11A). Pressurizing middle or posterior

cerebral arteries to 80 mmHg augmented resting tone to 104 ± 6 µm (Fig. 11, B and D). In this

depolarized and constricted state, nifedipine application had a disproportionately greater effect

on myogenic tone development than did mibefradil. Nifedipine-sensitive change in diameter (Δ

Diameter in Fig. 11E) was calculated as the vessel diameter measured in the presence of 300 nM

nifedipine minus the value measured for control (absence of any channel blockers). The

Page 54: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

39

mibefradil-sensitive change in diameter was calculated as the diameter recorded in the presence

of both 300 nM nifedipine and 5 µM mibefradil minus the value recorded in the presence of 300

nM nifedipine. These values were determined from measurements made with arterial pressures

of 20 mmHg and 80 mmHg, see Fig. 11E. The nifedipine-sensitive and mibefradil-sensitive

changes in diameter were also expressed as percentages of the maximal change in diameter for

the artery. Maximal change was calculated as the difference between artery diameters recorded

for control (no blockers present) and in Ca2+ free buffer. These results are shown in Fig. 11E,

which indicates, by using both absolute and relative representations, that L-type Ca2+ channel

activation plays an increasing role in tone development with pressurization. In contrast, the T-

type Ca2+ channel component did not substantively change in absolute terms with vessel

pressurization; it did, however, significantly decrease when expressed relative to the maximal

response (see Fig. 11E). A similar pattern was observed when the preceding experiment was

repeated with kurtoxin (Fig. 12) and efonidipine (Fig. 13), two structurally dissimilar T-type

Ca2+ channel blockers (Furukawa et al., 2004; Horiba et al., 2008). In these experiments

nifedipine-sensitive, kurtoxin-sensitive and efonidipine-sensitive changes in diameter were also

calculated as described above for the experiments in which mibefradil was used. These values

were also expressed as percentages of the maximum vessel response as described above.

The relative importance of L- and T-type Ca2+ channels was better revealed in Figure 14

where myogenic tone was monitored over a more complete pressure range (20–100 mmHg)

before and after the sequential addition of 300 nM nifedipine and 5 µM mibefradil. As above, the

functional contribution of L-type Ca2+ channels to arterial tone increased with pressurization,

whereas that of T-type Ca2+ channels decreased. Although T-type Ca2+ channel blockade elicits

Page 55: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

40

only limited vasomotor responses, these alterations are likely sufficient to influence blood flow,

as evident from blood flow modeling (Fig. 15). This computational work was used to predict

how changes in vessel diameter could affect blood flow. This work revealed that a 5–15%

change in resting diameter (comparable to that induced by T-type Ca2+ channel blockade at 60

mmHg) could elicited a 20–50% change in network perfusion. With these simulations, we

assume that T-type Ca2+ channel expression is constant among the three vessel orders.

2.4 DISCUSSION

In this study of the rat cerebral circulation, we determined which voltage-gated Ca2+

channels are expressed and whether they are important to myogenic tone development. In this

regard, we used 1) molecular/biochemical approaches to determine which α1-subunits were

present, 2) patch clamp electrophysiology to characterize Ca2+ channel properties, and 3) vessel

myography to ascertain function. An mRNA assessment revealed that the α1-subunits of L-type

(CaV1.2) and T-type (CaV3.1 and CaV3.2) Ca2+ channels were present in isolated rat cerebral

arterial smooth muscle cells. Western analysis subsequently confirmed protein expression in

intact cerebral arteries. Whole-cell electrophysiology demonstrated that an inward Ba2+ current

was readily observable and divisible into a nifedipine-sensitive (L-type) and a nifedipine-

insensitive component. The latter was inhibited by mibefradil, kurtoxin, and efonidipine and

displayed activation/inactivation properties consistent with T-type Ca2+ channels. Functional

experiments further revealed that L-type Ca2+ channels are principally responsible for driving

myogenic tone, although a limited role for T-type Ca2+ channels is suggested in hyperpolarized

Page 56: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

41

vessels. Blood flow modeling indicates that T-type Ca2+ channel modulation could alter blood

flow through the posterior cerebral circulation by ~ 20–50%.

2.4.1 Background.

Voltage-gated Ca2+ channels are expressed in a diverse range of excitable and

nonexcitable tissues (Catterall et al., 2005). Structurally, the pore-forming α1-subunit comprises

four membrane-spanning domains (I-IV) linked together in a single polypeptide chain. Each

domain contains six transmembrane segments plus a P loop that dips incompletely into the pore

to confer selectivity (Dolphin, 2006; Hofmann et al., 1999). Molecularly, voltage-gated Ca2+

channels are categorized according to α1-subunit expression, and there are, at present, three main

classifications including CaV1.x (L-type), CaV2.x (P/Q-, N-, and R types), and CaV3.x (T-type).

CaV1.x and CaV2.x encode for Ca2+ channels that are more active at depolarized voltages

compared with the low voltage-activated CaV3.x subunits (Ball et al., 2009; Hansen et al., 2011).

In vascular smooth muscle, CaV1.2 principally controls steady-state Ca2+ entry and, therefore,

myosin light chain phosphorylation and tone development (Harder, 1984; Knot & Nelson, 1998).

Although important to tone development, recent studies have suggested that this α1-subunit is not

singularly expressed in arterial smooth muscle (Braunstein et al., 2009; Kuo et al., 2010;

Nikitina et al., 2007; Pluteanu & Cribbs, 2011). For example, immunohistochemical

observations have suggested that α1-subunits of T-type Ca2+ channels are present in the renal,

mesenteric, and cerebral circulation (Braunstein et al., 2009; Feng et al., 2004; Navarro-

Gonzalez et al., 2009). Supporting electrophysiological findings are sparse, although selected

studies have shown that a nifedipine-insensitive Ba2+ current with properties reminiscent of T-

Page 57: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

42

type Ca2+ channels can be isolated in arterial smooth muscle (rat and dog basilar arteries)

(Navarro-Gonzalez et al., 2009; Nikitina et al., 2007). At present, it is not clear what role T-type

Ca2+ channels play in arterial tone development. Past functional findings have been decidedly

mixed with only a small number indicating a limited to prominent contribution, depending on the

vascular bed and stimulus (Braunstein et al., 2009; Hansen et al., 2011; Kuo et al., 2010; Poulsen

et al., 2011). Others, however, have argued that T-type Ca2+ channels have no role in the

contractile process and that their principal function centers on cell cycle regulation and

proliferation (Oguri et al., 2010; Perez-Reyes, 2003; Rodman et al., 2005).

2.4.2 Molecular and electrical characterization.

This study began by examining which voltage-gated Ca2+ channels are present in cerebral

arterial smooth muscle. Work was initiated at a molecular level with a PCR analysis of whole

arteries and isolated smooth muscle cells from rat middle/posterior cerebral arteries. Consistent

with past findings (Braunstein et al., 2009), multiple α1-subunits were observed in whole arteries

including CaV1.1-1.3, CaV2.2, and CaV3.1–3.3 (Fig. 4). The breadth of the mRNA profile was

expected since whole arteries comprise three major cell types (smooth muscle, endothelium, and

perivascular nerves), all of which can express CaV α1-subunits (Cribbs, 2001). Limiting mRNA

analysis to isolated smooth muscle cells reduced the expression pattern to CaV1.2 (L-type) and

CaV3.1/3.2 (T-type). Western blot analysis subsequently confirmed that each α1-subunit was

expressed at the protein level in cerebral arteries (Fig. 5). Although this study did not pursue an

immunohistochemical analysis, Kuo et al. have shown that all three α1-subunits are indeed

present in rat cerebral arterial smooth muscle (Kuo et al., 2010). With molecular/biochemical

Page 58: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

43

evidence of L- and T-type Ca2+ channels expressed in cerebral arterial smooth muscle we next

used patch clamp electrophysiology to functionally isolate and identify the two general

conductances. Like past studies, experimentation began by monitoring Ba2+ currents in cells held

at -90 or -60 mV, the concept being that the hyperpolarized holding potential would relieve

voltage-dependent inactivation of T-type Ca2+ channels and thus augment whole-cell current

(Navarro-Gonzalez et al., 2009; Nikitina et al., 2007). We did not observe a significant change in

current. Therefore, we next monitored inward Ba2+ currents in the presence and absence of

nifedipine to determine whether a dihydropyridine-insensitive current with T-type Ca2+ channel

properties could be functionally isolated (Fig. 7). Nifedipine blocks L-type Ca2+ channels in a

state-dependent manner by binding to inactivated pores (Dolphin, 2006; Liao et al., 2007).

Concentrations were set 5- to 30-fold above the IC50 for the smooth muscle splice variants of

CaV1.2 (Jensen & Holstein-Rathlou, 2009) and below levels that interfere with T-type Ca2+

channels (Akaike et al., 1989; Lee et al., 2006). At levels that should block >98% of all L-type

Ca2+ channels (300 nM), whole-cell inward current decreased by ~75%, a finding in agreement

with Nikitina et al. (Nikitina et al., 2007) in dog basilar artery and Kuo et al. (Kuo et al., 2010) in

the rat cerebral circulation. The residual nifedipine-insensitive current displayed a leftward

shifted current-voltage relationship characteristic of T-type Ca2+ channels (Iftinca & Zamponi,

2009; Perez-Reyes, 2003). A more detailed examination of the nifedipine-insensitive current

followed with experiments monitoring the effects of T-type Ca2+ channel blockers and

characterizing key electrical properties. In regards to the former (the effects of the blockers),

Figs. 8–10 show that mibefradil, efonidipine, and kurtoxin effectively blocked much of the

nifedipine-insensitive Ba2+ current. The analysis of mibefradil/efonidipine/kurtoxin-sensitive

Page 59: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

44

currents revealed that the activation/inactivation time constants were smaller than those

determined for the nifedipine-sensitive current and corresponded well with published values for

CaV3.1/3.2 and CaV1.2, respectively (Kuo et al., 2010; Perez-Reyes, 2003). Overall, the

preceding patch clamp observations indicate L- and T-type Ca2+ channels are not only

electrically present in cerebral arterial smooth muscle but distinguishable from one another.

2.4.3 T-type Ca2+ channels and myogenic tone.

When current through T-type Ca2+ channels is recorded in solutions containing

physiological concentrations of Ca2+, the channels display a V0.5 for steady-state

activation/inactivation that is 10-15 mV negative to CaV1.2 (Nikitina et al., 2007; Perez-Reyes,

2003). As such, there will be a subsequent leftward shift in the window current (Kuo et al., 2010;

Navarro-Gonzalez et al., 2009; Nikitina et al., 2007; Perez-Reyes, 2003). From this knowledge,

it is logical to predict that the contribution of T-type Ca2+ channels to tone development should

be more prominent at hyperpolarized potentials. To test this hypothesis, this study examined

myogenic tone development and the effect of sequential addition of L- and then T-type Ca2+

channel blockers on cerebral arteries pressurized to 20 or 80 mmHg. It was theorized that at 20

mmHg, vessels would reside in a hyperpolarized state (~-55 mV) and that T-type Ca2+ channel

activity would be greater given their activation/inactivation properties (Perez-Reyes, 2003). In

contrast, at 80 mmHg, cerebral arteries would be more depolarized (~-35 mV) and L-type Ca2+

channel activity would be predicted to dominate to a greater extent over T-type (Bannister et al.,

2009; Knot & Nelson, 1995). Findings in Figs. 11–13 clearly illustrate that T-type Ca2+ channel

blockers, such as mibefradil, kurtoxin and efonidipine, do impair the nifedipine-insensitive

Page 60: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

45

component of the myogenic response. When expressed on a relative scale, the contribution of T-

type Ca2+ channels was greatest at low intravascular pressures where the vessel is hyperpolarized

(-55 mV) (Kuo et al., 2010; Nikitina et al., 2007; Perez-Reyes, 2003). As expected, the

functional significance of L-type Ca2+ channels increased with pressurization, a finding

consistent with the literature and the electrophysiological properties of high voltage-activated

Ca2+ channels (Bannister et al., 2009; Knot & Nelson, 1998). The unequal contribution of L- and

T-type Ca2+ channels to myogenic tone development is better observed in Fig. 14 when cerebral

arteries were stepped sequentially from 20 to 100 mmHg. Although previous studies have

attempted similar measurements (Hansen et al., 2011; Kuo et al., 2010; Nikitina et al., 2007),

this study is the first to show clear evidence that T-type Ca2+ channels contribute to myogenic

tone development in a manner consistent with their voltage-dependent properties (Perez-Reyes,

2003).

In regard to the preceding findings and their interpretation, this study implicitly assumed

that T-type Ca2+ channels operate as a general influx pathway, providing a portion of the

extracellular Ca2+ required for contractile activation. Although reasoned, it is plausible that T-

type Ca2+ channels might also facilitate arterial tone through alternative mechanisms. For

example, Ca2+ flux via CaV3.1 or CaV3.2 channels might be sufficient to activate a Ca2+-

dependent inward current such as TRPM4, a transient receptor potential channel that contributes

to pressure-induced depolarization (Earley, 2010). They might also initiate and/or maintain Ca2+

waves, asynchronous events dependent on Ca2+ release from the SR (Braunstein et al., 2009;

Martin-Cano et al., 2009). Like T-type Ca2+ channels, these events are more important in setting

myogenic tone at hyperpolarized rather than depolarized voltages (Mufti et al., 2010).

Page 61: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

46

2.4.4 Implications to blood flow control.

The observations in Figures 11–14 suggest that T-type Ca2+channels contribute to

myogenic tone development in the cerebral circulation. Although their overall role is limited, it

would be imprudent to overlook their impact on tissue perfusion since flow is proportionally

related to blood pressure and vessel radius raised to the 4th power. To better illustrate the impact

of T-type Ca2+ channels on tissue perfusion, blood flow simulations were performed on a

network structure that mimicked the proximal aspects of the posterior cerebral circulation (Fig.

15). This network structure consisted of seven segments whose diameter and length were set

according to measurements on 1st- to 3rd-order posterior cerebral arteries pressurized to 60

mmHg. Blood and RBC volume flow were conserved, whereas hematocrits along with inlet and

outlet pressure were set to 0.4, 75, and 50 mmHg, respectively. With this model, we observed

that 5–15% changes in resting arterial diameter, generally akin to dilatory responses induced by

T-type Ca2+ channel inhibition at 60 mmHg, augmented network perfusion by ~20–50%. From

such findings, we deduce that T-type Ca2+ channel modulation could be functionally important

and have a measurable impact on tissue perfusion.

2.4.5 Summary.

This study demonstrated that multiple voltage-gated Ca2+ channels are present in the rat

cerebral circulation. More specifically, molecular and electrophysiological approaches indicated

that both L-type (CaV1.2) and T-type (CaV3.1 and CaV3.2) Ca2+ channels are expressed and

electrically discernible in smooth muscle cells from middle/posterior cerebral arteries. Functional

analysis indicated that T-type Ca2+ channels contribute to the genesis of the myogenic response.

Page 62: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

47

Their functional significance would be more evident at low intravascular pressure where the

hyperpolarized state of the vessel would optimize steady-state Ca2+ influx, given the voltage-

dependent properties of the channel (Perez-Reyes, 2003). Although the overall contribution of T-

type Ca2+ channels to tone control is modest, blood flow modeling indicates that T-type Ca2+

channel modulation could alter resting blood flow by as much as 50%.

2.4.6 Limitations and future directions.

This study provided evidence for the presence of CaV1.2 (L-type) and CaV3.1 and CaV3.2

(T-type) voltage-gated Ca2+ channels in rat cerebral arteries and the results presented also

suggest functional roles for CaV3.1 and CaV3.2 T-type VGCCs in myogenic tone development.

However, there are some experimental limitations and additional experiments that should be

considered.

(i) Molecular and biochemical approach.

The semi-quantitative reverse transcription PCR technique was performed on isolated

smooth muscle cells to scan for mRNA message of VGCCs. In these experiments posterior and

middle cerebral arteries were isolated as described previously (Luykenaar et al., 2004; Talavera

et al., 2001). The tissues were treated enzymatically, then washed repeatedly with ice-cold

isolation medium and triturated with a fire-polished pipette to liberate the cells. Individual cells

were collected manually with a suction pipette and pooled together in RNase- and DNase-free

collection tubes to extract total RNA (RNeasy mini kit with DNAase treatment; Qiagen,

Valencia, CA). First-strand cDNA was synthesized using the Sensi-script RT kit (Qiagen). The

Page 63: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

48

results showed the presence of the α1–subunit of L-type (CaV1.2) and T-type (CaV3.1 and

CaV3.2) channels in isolated cerebral arterial smooth muscle cells.

Although the preparations were pre-screened for endothelial contamination (ET-1), a

screening for neuronal contamination in the preparations was not done. As the arteries are

comprised of three cell types (smooth muscle, endothelial, perivascular nerves), any

contamination from perivascular nerves could cause misleading results regarding the expression

of VGCCs in smooth muscle; this is especially true for T-type VGCCs, which are abundantly

expressed in the neuronal tissues.

One way to screen the preparations for neuronal cell contamination would be to use semi-

quantitative reverse transcription PCR with primers that will allow for the detection of DNA that

codes for neuronal cell specific marker protein (e.g. neurofilament (NF), subtypes of NF, NF-L,

NF-M, NF-H, or α-internexin). A second way to screen the preparations for a neuronal cell

contamination would be to use real-time reverse transcription PCR, which quantitates differences

in mRNA expression, to run two parallel reactions that would amplify a smooth muscle specific

marker (e.g. smooth muscle myosin) and a neuronal marker (e.g. neurofilament) by using SYBR

green core reagents (Applied Biosystems, Carlsbad, CA, USA). SYBR green is a dye that binds

to double stranded DNA but not to single-stranded DNA and it fluoresces very brightly, and can

be used to detect the amplified DNA. A third method to check for neuronal contaminations

would also use reverse transcription PCR. Forward and reverse primers for a smooth muscle

specific marker (e.g. smooth muscle myosin) and a for a neurofilament marker (e.g. NF or any

subtype of NF) would be added to a single sample and used to amplify DNA. SYBR green would

be used to label all amplified DNA. A melt curve would then be generated to indicate the

Page 64: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

49

presence of one or two components. If only one component was found it would presumably be

from smooth muscle and would indicate the sample had no neuronal contamination. If two

components were found in the melt curve that would indicate the sample is contaminated by

neurons. If the results of the screening show there is no neuronal contamination, then our results

indicate the presence of CaV3.1 and CaV3.2 RNA in smooth muscle cells, which would be in

agreement with the results of Kuo and Navarro-Gonzales et al. who applied real-time PCR, and

detected the presence of the two subtypes (CaV3.1 and CaV3.2) of T-type VGCCs in rat cerebral

arteries (Kuo et al., 2010; Navarro-Gonzalez et al., 2009).

A test could also be done to check if the primers used in PCR reactions for CaV3.1 and

CaV3.2 resulted in amplification of any other types of DNA (e.g. splice variants or DNA from

another type of channel). In this test PCR reaction could be run using the primers for one of the

channels and SYBR green could be used to label all of the amplified DNA. A melt curve could

be generated and would indicate whether one or more components were amplified; additional

components could be splice variants for example for the target channels.

Western blot analysis was performed with whole cerebral arteries. However, western blot

analysis could also be carried out on samples prepared from isolated smooth muscle cells. This

would involve preparing a suspension of isolated cells from cerebral artery segments, collecting

individual smooth muscle cells from the cell suspension using a suction pipette and then pooling

the collected cells from multiple preparations as was done for the PCR analysis.

The previous approaches (biochemical, molecular) have indicated the presence of

different types and subtypes of VGCCs in rat cerebral arteries. Immunofluorescence microscopy

involving labelling of isolated smooth muscle cells with antibodies to the channels would be

Page 65: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

50

another approach that could be used to further investigate the presence of both L-and T-type

VGCCs in these cells. Experiments of this type are described in Chapter 3, section 3.4.4.

Primary antibodies against selected α1-subunits of L- and T-type VGCCs (CaV1.2,

CaV3.1, and CaV3.2) were used and the results indicated the presence of these channels in rat

cerebral arteries; however, the specificity of the primary antibodies could be better tested to

make sure there is not any non-specific binding for the antibodies used. There are a number of

ways that could be used to look at specificity of the primary antibodies. 1) The labelling of the

whole blot can be examined to look for extraneous bands (at molecular weights different than

that expected for the channels) that might indicate non-specific binding. 2) Competition assays

using the peptide antigens for the primary antibodies can also be used to examine specificity of

the antibodies. 3) Primary antibodies from different sources could be compared and used to

determine which has the highest specificity. 4) A small interference RNA (siRNA) approach

could be used to selectively reduce the expression of the protein that is targeted by a particular

antibody. siRNA interference is a methodology whereby small segments of double-stranded

RNA are introduced into isolated cerebral arteries using reversible permeabilization in organ

culture media, to degrade the mRNA required for protein translation. The degradation is

controlled by an endogenous RNA-induced silencing complex, DICER and RISC (Hamilton and

Baulcombe, 1999). Western blot analysis of protein from tissue treated with siRNA should show

less labeling of the target protein band(s) if the primary antibody is specific for the protein.

Small interfering RNA (siRNA) is a small segment of double stranded RNA that binds to

and degrades the mRNA essential for protein translation. Therefore, it greatly reduces or

eliminates the expression of the targeted protein, and induces a type of protein knockout.

Page 66: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

51

However, the ability to deliver siRNA to cells is limited as many mammalian cell types are

resistant to the transfection methods needed to introduce the synthetic siRNA into them. The

technique requires organ culture for approximately about 3 days. This could expose the vessels

to contamination or tissue damage. Furthermore, the siRNA technique can also be non-specific;

off-target effects of siRNA can also occur if it targets mRNA that is not the intended one.

Application of the siRNA technique is also expensive for large experiments and it mediates

short-term knockdown, so it is not useful for long term studies.

As part of applying the siRNA technique, control tissues have to be tested to confirm the

validity of the technique itself, as both the controls and the treated siRNA tissues are subject to

organ culture for about 3 days. One control tissue would be non-treated (i.e. have no siRNA) and

the second control tissue would be treated with a scrambled RNA sequence that would not affect

the expression of the target protein. If western blot analysis of samples from the control and

siRNA treated tissues do not show changes in the bands labeled by the primary antibody being

tested then that would indicate that those bands are due to non-specific labelling by the antibody.

(ii) Patch-clamp electrophysiological approach.

- Patch-clamp measurements using Ba2+

Conventional patch clamp electrophysiology results demonstrated the presence of two

inward VGCC channel currents (nifedipine-sensitive and nifedipine-insensitive). In cerebral

arterial smooth muscle cells, Ba2+ (10 mM) was used as the charge carrier because the current

carried by Ba2+ is larger than the current carried by Ca2+ therefore it is easier to measure and

differentiate between L- and T-type channels (Perez-Reyes, 2003, Harraz & Welsh, 2013).

Page 67: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

52

Although using Ba2+ makes it easier to measure currents through VGCCs, Ba2+ also causes the L-

type channels to lose their Ca2+-dependent inactivation properties (Nikitina et al., 2007) and the

high concentration of Ba2+ used leads to a shift in the current-voltage (IV) curve toward more

hyperpolarized values by ~10 mV. This shift has to be considered when comparing channel types

and their properties, as it could give a false indication about the identity of the channels being

investigated (Abd El-Rahman et al., 2013; Kuo et al., 2010). For example it would make it hard

to distinguish between different types of VGCCs and splice variants of VGCCs (Abd El-Rahman

et al., 2013; Kuo et al., 2010). One way to address these possible problems would be to use Ca2+

as the charge carrier (1.8 mM, as higher concentration will cause the cells to contract). Although

the current amplitude will be approximately three times less than the current amplitude when

Ba2+ is used (Kuo et al., 2010) this method has been used to record L- and T-type currents

(Harraz and Welsh, 2013) but it was not possible to differentiate between different types of

channels because of the small current.

- Protocol with different holding potentials.

Figure 6, shows results of experiments in which Ba2+ currents were measured at different

voltages from starting prepulses of -90 mV or -60 mV. The prepulse of -60 mV was used to

maximize inactivation of T-type channels. In these experiments we did not find statistically

significant differences in the currents recorded at any voltage when the two prepulses were

compared. Therefore these results could not be used as evidence of the presence of two different

inward currents. These results could be further investigated by repeating the experiments to

increase the n numbers for the two different prepulses. We did notice a possible small (but not

Page 68: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

53

significant) decrease in currents when -60 mV was used as a prepulse. These experiments could

be redone by repeating the two protocols several times in a row and comparing results sets (i.e.

of 1st set of traces to 2nd ; 2nd set of traces to 3rd etc…) to eliminate the possibility that the

currents change over time (due to run-down for example). Control experiments in which a series

of traces are compared all at the same prepulse (i.e. at -90 mV prepulse or -60 mV prepulse)

could also be done and analyzed the same way to determine if the rundown of the currents with

time affected the results.

- Use of T-type channel blockers.

One difficulty with the use of pharmacological blockers in the present study is that the

available blockers are not as specific as one would like. For example mibefradil shows high

affinity for blocking T-type Ca2+ channels if Ca2+ is used as the charge carrier but competes with

Ba2+ more than Ca2+ for permeation through the pore, thus both concentration and divalent cation

type affect the efficacy of the drug (Morita et al., 2002; Martin et al., 2000; Lee et al., 1999;

Perez-Reyes, 2003). These results should be taken into consideration when investigating the

presence of different VGCCs in arterial smooth muscle, because it could affect the interpretation

of the data. For example it might be difficult to differentiate between splice variants of a

particular type of a channel and the different types of channels. Another way to test for blockers

specificity would be to measure the currents that remain in the presence of one T-type channel

blocker and then measure it again after the addition of a 2nd T-type blocker along with the first

one (different combinations of the blockers could be used e.g. mibefradil then kurtoxin,

mibefradil then efonidipine, kurtoxin then efonidipine, or vice versa,…etc). If the second blocker

Page 69: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

54

causes additional effects on the channel that might be an indication that additional channels types

are affected; if the second blocker addition causes no further change in the remaining current that

would strengthen the conclusion that the same channels are being affected by the channel

blockers.

Also, the siRNA approach discussed above could be employed to address the problem of

using non-specific blockers. For these experiments Ca2+ currents from cells isolated from

untreated or tissue treated with scrambled siRNA could be compared with currents from cells

treated with siRNA to knock down CaV3.1 and CaV3.2 T-type channels. The effects of L- and T-

type channel blockers could also be compared in these cells; siRNA treatment should eliminate

effects of the T-type channel blockers if they are specific.

(iii) Vessel myography approach.

The experimental scheme for the vessel myography relied mainly on the use of different

pharmacological blockers (mibefradil, kurtoxin, efonidipine) that block T-type VGCCs (CaV3.1

and CaV3.2) when used at a specific range of concentrations (see Table 1; Lee et al., 1999;

Perez-reyes, 2003). However, these pharmacological blockers also affect other channels, at

higher concentrations (K+, Na+, L and N-type Ca2+ channels, and Cl- channels) which could

mislead the vessel myography results. Furthermore, the blockers block T-type Ca2+ channels in

different ways from one another. Mibefradil inhibits the channels in both dose-dependent and

use-dependent manner, as the channels become more active, inhibition increases because it

blocks the channels through the open pore (Morita et al., 2002; Jimenez et al, 2000; Lee et al,

1999; Perez-Reyes, 2003). Block by kurtoxin, a 63 residue peptide toxin, is voltage-dependent so

Page 70: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

55

a minor depolarization causes inhibition as kurtoxin binds to the channel’s voltage sensor, and

produces a complex modification of gating and accelerated deactivation (Horiba et al., 2008; Lee

et al., 1999; Perez-Reyes, 2003). R (-) efonidipine acts the same way as the DHPs act on L-type

channels. It stabilizes the inactivated states of T-type VGCCs (Furukawa et al., 2004; Lee et al.,

1999; Perez-Reyes, 2003).

Those properties suggest that the results we achieved could be complex because the

pressure-dependent effects we observe could be influenced by the ways in which the drugs bind

to different states of the channels and the parts of the channel that they interact with. The

experiments in Figures 11, 12, and 13 examining diameter changes could be repeated by addition

of nifedipine, and then the addition of one of the T-type blockers followed by the addition of a

second T-type blocker in combination with nifedipine and the first T-type blocker. This would

allow us to determine if effects of the blockers are additive or not. Additive effects would

suggest that other types of channels might also be affected by the blockers.

To overcome the non-specificity of the blockers the siRNA approach outlined above

could be introduced to rat cerebral arteries to investigate VGCCs in control and siRNA treated

arteries.

(iv) Computational approach.

The mathematical model that was constructed to predict blood flow in a cerebral arterial

network demonstrated, on a theoretical level, that T-type VGCCs are capable of significantly

affecting blood flow. This mathematical model permits many different questions to be

investigated on a theoretical level and could open the door for more wet lab experiments and

Page 71: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

56

investigations. Two questions that could be investigated and addressed with the model are: 1)

how might different expression levels of the channels at different vessel orders affect blood flow

and 2) At what order would a given dilation have the greatest effect. At the same time,

experiments could be done to screen for the quantitative expression of the VGCCs at different

vessel orders using reverse transcriptase PCR. Such experiments could indicate a relatively

greater role for T-type channels, compared to L-type channels, at different vessel orders. This

could also suggest that vessel myography could be used to investigate the role of T-type channels

in different vessel orders.

Page 72: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

57

2.5 Figures.

Page 73: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

58

Figure 4: mRNA expression of the α1-subunits of voltage-gated Ca2+ channels. Voltage-

gated Ca2+ channel expression in control tissues (skeletal muscle [CaV1.1], brain [CaV1.2,

CaV1.3, CaV2.1, CaV2.2, and CaV2.3], heart [CaV3.3], afferent arteriole [CaV3.1 and CaV3.2], and

retina [CaV1.4]) (A). Whole arteries or isolated smooth muscle cells from middle/posterior

cerebral arteries were collected and processed for RT-PCR. CaV α1-subunit expression in whole

cerebral arteries (B) and in enzymatically isolated smooth muscle cells (C), which indicate the

presence of L-type (CaV1.2), and T-type channels (CaV3.1 and CaV3.2). “M” indicates the size

marker lane (Performed by Rasha Abd El-Rahman).

Page 74: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

59

Page 75: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

60

Figure 5: Protein expression of the α1-subunits of voltage-gated Ca2+ channels. Rat cortex

along with vessel segments from middle/posterior cerebral arteries were collected and separately

prepared for western blot analysis. Protein extracts were electrophoresed on 5.6%

polyacrylamide gel, transferred to PVDF and then labeled with primary antibodies against

CaV1.2, CaV3.1, and CaV3.2 (Performed by Rasha Abd El-Rahman).

Page 76: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

61

Page 77: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

62

Figure 6: Inward Ba2+ currents in cerebral arterial smooth muscle cells. Whole cell patch

clamp electrophysiology was used to measure Ba2+ currents in smooth muscle cells isolated from

middle/posterior cerebral arteries. The voltage protocol consisted of a prepulse at -60 or -90 mV

(200 ms) followed by a series of steps (-80 to +40 mV, 200 ms, 10 mV increments). A)

Representative Ba2+ currents from a smooth muscle cell exposed to a prepulse at -90 mV (left)

or -60 mV (right). B) IV plot of peak inward Ba2+ current (n = 6) recorded with a prepulse at -90

mV (ο) or -60 mV (•). (Performed by Osama Harraz; cells were prepared by Rasha Abd El-

Rahman).

Page 78: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

63

Page 79: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

64

Figure 7: Nifedipine-sensitive and -insensitive Ba2+ currents in cerebral arterial smooth

muscle cells. Whole cell patch clamp electrophysiology was used to measure Ba2+ currents in

smooth muscle cells isolated from middle/posterior cerebral arteries. The voltage protocol

consisted of a pre-pulse from -60 to -90 mV (200 ms) followed by a series of steps (-80 to +40

mV (300 ms, 10 mV increments). A) Representative Ba2+ current under control conditions and in

the presence of 50 nM nifedipine (left two panels), and representative Ba2+ current under control

conditions and in the presence of 300 nM nifedipine (right two panels). B) IV plot of peak

inward Ba2+ current along with the nifedipine-sensitive and –insensitive components (n = 6). C)

IV plot of the nifedipine-sensitive and -insensitive Ba2+ current expressed relative to maximal

current at +10 mV for 50 nM nifedipine (left) and 300 nM nifedipine (right). (Performed by

Osama Harraz; cells were prepared by Rasha Abd El-Rahman).

Page 80: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

65

Page 81: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

66

Figure 8: Effects of mibefradil on the nifedipine-insensitive component of the inward Ba2+

current. Whole cell patch clamp electrophysiology was used to measure Ba2+ currents in

smooth muscle cells isolated from middle/posterior cerebral arteries. The voltage protocol

consisted of a pre-pulse from -60 to -90 mV (200 ms) followed by a series of steps (-80 to +40

mV, 200 ms, 10 mV increments). This protocol was run under control conditions and in the

presence of nifedipine (300 nM) + mibefradil (5 µM). A) Representative traces of Ba2+ current

under control conditions and in the presence of nifedipine and nifedipine plus mibefradil. The

nifedipine-sensitive component and the mibefradil-sensitive component were determined by

subtraction of measured currents as described in the Methods section. B) IV plot of nifedipine-

and mibefradil-sensitive currents (n = 9). C) Time constant (τ) for inactivation (left & center

panels) and activation (right panel) for the nifedipine-sensitive and mibefradil-sensitive

components at 10 mV (left) and 20 mV (center). The time constant for activation at 10 mV for

these current components is shown in the right panel (n = 9). * denotes significant difference

from the nifedipine-sensitive current (paired t-test, and one way ANOVA, P < 0.05) (Performed

by Yana Anfinogenova).

Page 82: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

67

Page 83: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

68

Figure 9: Effects of kurtoxin on the nifedipine-insensitive component of the inward Ba2+

current. Whole cell patch clamp electrophysiology was used to measure Ba2+ currents in smooth

muscle cells isolated from middle/posterior cerebral arteries. The voltage protocol consisted of a

pre-pulse from -60 to -90 mV (200 ms) followed by a series of steps (-80 to +40 mV, 200 ms, 10

mV increments). This protocol was run under control conditions and in the presence of

nifedipine (300 nM) + kurtoxin (5 µM). A) Representative traces of Ba2+ current under control

conditions and in the presence of nifedipine and nifedipine plus kurtoxin. The nifedipine-

sensitive component and the kurtoxin-sensitive component were determined by subtraction of

measured currents as described in the Methods section. B) IV plot of nifedipine- and kurtoxin-

sensitive currents (n = 6). C) Time constant (τ) for inactivation (left & center panels) and

activation (right panel) for the nifedipine-sensitive and kurtoxin-sensitive components at 10 mV

(left) and 20 mV (center). The time constant for activation at 10 mV for these current

components is shown in the right panel (n = 6). * denotes significant difference from the

nifedipine-sensitive current (paired t-test, and one way ANOVA, P < 0.05) (Performed by Yana

Anfinogenova).

.

Page 84: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

69

Page 85: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

70

Figure 10: Effects of efonidipine on the nifedipine-insensitive component of the inward Ba2+

current. Whole cell patch clamp electrophysiology was used to measure Ba2+ currents in smooth

muscle cells isolated from middle/posterior cerebral arteries. The voltage protocol consisted of a

pre-pulse from -60 to -90 mV (200 ms) followed by a series of steps (-80 to +40 mV, 200 ms, 10

mV increments). This protocol was run under control conditions and in the presence of

nifedipine (300 nM) + efonidipine (3 µM). A) Representative traces of Ba2+ current under

control conditions and in the presence of nifedipine and nifedipine plus efonidipine. The

nifedipine-sensitive component and the efonidipine-sensitive component were determined by

subtraction of measured currents as described in the Methods section. B) IV plot of nifedipine-

and efonidipine-sensitive currents (n = 6). C) Time constants (τ) for inactivation (left & center

panels) and activation (right panel) for the nifedipine-sensitive and efonidipine-sensitive

components at 10 mV (left) and 20 mV (center). The time constant for activation at 10 mV for

these current components is shown in the right panel (n = 6). * denotes significant difference

from the nifedipine-sensitive current (paired t-test, and one way ANOVA, P < 0.05) (Performed

by Yana Anfinogenova).

.

Page 86: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

71

Page 87: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

72

Figure 11: The effects of nifedipine and mibefradil on myogenic tone. Cerebral arteries were

pressurized to 20 or 80 mmHg while diameter was monitored in the absence and presence of

nifedipine (50 and 300 nM) and nifedipine plus mibefradil (1 and 5 μM). A & B) Representative

traces revealing the effect of nifedipine and mibefradil on cerebral arteries pressurized to 20 or

80 mmHg. C & D) Summary data denoting the influence of nifedipine and mibefradil on

cerebral arteries pressurized to 20 (n = 7) or 80 (n = 6) mmHg. * and ** denote significant

increases from control and nifedipine (300 nM), respectively (paired t-test and one way

ANOVA, P < 0.05). E) The nifedipine- and mibefradil-sensitive vasomotor responses are plotted

as an absolute change or as a percent of maximal response. * denotes significant difference

from 20 mmHg (unpaired t-test, and one way ANOVA, P < 0.05) (Performed by Rasha Abd El-

Rahman).

.

Page 88: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

73

Page 89: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

74

Figure 12: The effects of nifedipine and kurtoxin on myogenic tone. Cerebral arteries were

pressurized to 20 or 80 mmHg while diameter was monitored in the absence and presence of

nifedipine (50 and 300 nM) and nifedipine plus kurtoxin (1 μM). A & B) Representative traces

revealing the effect of nifedipine and kurtoxin on cerebral arteries pressurized to 20 or 80

mmHg. C & D) Summary data denoting the influence of nifedipine and kurtoxin on cerebral

arteries pressurized to 20 (n = 5) or 80 (n = 6) mmHg. * and ** denote significant increases

from control and nifedipine (300 nM), respectively (paired t-test and one way ANOVA, P <

0.05). E) The nifedipine- and kurtoxin-sensitive vasomotor responses are plotted as an absolute

change or as a percent of maximal response. * denotes significant difference from 20 mmHg

(unpaired t-test, and one way ANOVA, P < 0.05) (Performed by Rasha Abd El-Rahman).

Page 90: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

75

Page 91: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

76

Figure 13: The effects of nifedipine and efonidipine on myogenic tone. Cerebral arteries

were pressurized to 20 or 80 mmHg while diameter was monitored in the absence and presence

of nifedipine (50 and 300 nM) and nifedipine plus efonidipine (3 μM). A & B) Representative

traces revealing the effect of nifedipine and efonidipine on cerebral arteries pressurized to 20 or

80 mmHg. C & D) Summary data denoting the influence of nifedipine and efonidipine on

cerebral arteries pressurized to 20 (n = 5) or 80 (n = 6) mmHg. * and ** denote significant

increases from control and nifedipine (300 nM), respectively (paired t-test, P < 0.05). E) The

nifedipine- and efonidipine-sensitive vasomotor responses are plotted as an absolute change or as

a percent of maximal response. * denotes significant difference from 20 mmHg (unpaired t-test,

and one way ANOVA, P < 0.05) (Performed by Rasha Abd El-Rahman).

Page 92: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

77

Page 93: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

78

Figure 14: The effects of nifedipine and mibefradil on myogenic tone through a full range

of intravascular pressures. Cerebral arteries were sequentially pressurized from 20 to 100

mmHg while arterial diameter was monitored in the absence and presence of nifedipine (300

nM) and nifedipine plus mibefradil (5 µM). A) Representative traces revealing the effect of

nifedipine and mibefradil on cerebral arteries sequentially pressurized between 20 and 100

mmHg. B & C) The nifedipine- and mibefradil-sensitive vasomotor responses are plotted as an

absolute change or as a percent of maximal response. Absolute diameter (in μm) for control and

in Ca2+-free solution were as follows (n = 6): 20 mmHg, 175 ± 4, 188 ± 3; 40 mmHg, 171 ± 2,

211 ± 4; 60 mmHg, 153 ± 3, 220 ± 4; 80 mmHg,142 ± 4, 227 ± 4; and 100 mmHg, 135 ± 4 , 234

± 4. * denotes significant difference from the preceding pressure step (paired t-test and one way

ANOVA, P < 0.05) (Performed by Rasha Abd El-Rahman).

Page 94: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

79

Page 95: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

80

Figure 15: Computational modeling predicts that T-type Ca2+ channel modulation alters

tissue blood flow. A) A computational model (Goldman & Popel, 2000; Pries et al., 1990;

Pries et al., 1994) was used to calculate two-phase steady-state flow through an arterial network

consisting of 7 arterial branches whose diameter/ length approximates the posterior cerebral

arterial network diverging off the Circle of Willis. This model implements conservation of blood

and RBC volume flow at each node joining vessel segments and includes known blood rheology.

In addition to geometric information, this model required specification of boundary conditions in

the form of hematocrit (0.4) and pressure changes between the inlet (75 mmHg) and outlet (50

mmHg) segments. B) The impact of different dilation (5-15 %) on blood flow for the different

vessel orders. C) Represents the percentage change in blood flow at different percentages of

dilation (5, 10, and 15 %). (Computer programming by Daniel Goldman and simulations

performed by Daniel Goldman and Rasha Abd El-Rahman).

Page 96: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

81

Chapter Three: CaV3.2 Channels and the Induction of a Negative Feedback Response

In Cerebral Arterial Smooth Muscle

Rasha R. Abd El-Rahman, Osama F. Harraz, Kamran Bigdely-Shamloo, Sean M. Wilson,

Monica Rubalcava, Rania E. Mufti, Albert L. Gonzales, Scott Earley, Edward J. Vigmond

and Donald G. Welsh

I was responsible for structural and functional data collection and analysis and I wrote drafts of

the manuscript to be submitted for publication. Osama Harraz and I carried out the perforated

patch clamp experiments. Suzanne B. Welsh performed membrane potential measurements.

Computational modeling was performed by Kamran Bigdely-Shamloo and Edward J. Vigmond,

and Sean M. Wilson and Monica Rubalcava were responsible for the Ca2+ spark experiments.

Page 97: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

82

3.1 INTRODUCTION

Cerebral resistance arteries control tissue blood flow through alterations in smooth

muscle contractility initiated by neuronal stimulation, hormonal and metabolic state, and

intraluminal pressure (Wray et al., 2005). These factors initiate change in vessel diameter in part

by increasing cytosolic [Ca2+], a response that augments myosin light chain (MLC20)

phosphorylation via the regulation of myosin light chain kinase (MLCK) and myosin light chain

phosphatase (MLCP). The elevation in cytosolic [Ca2+] is sustained by the influx of extracellular

Ca2+ and by Ca2+ release from the sarcoplasmic reticulum (SR) (Berridge et al., 2000; Carafoli et

al., 2001). Extracellular Ca2+ influx is in turn driven by resting membrane potential and the

activity of voltage-gated Ca2+ channels (VGCCs).

VGCCs are heteromultimeric protein complexes which consist of a pore-forming α1-

subunit containing four domains (I-IV) of six transmembrane segments. In vascular smooth

muscle, CaV1.2 (L-type) is the key channel regulating extracellular Ca2+ influx, ultimately

leading to MLC20 phosphorylation and consequently tone development (Knot & Nelson, 1998;

Lambert et al., 1997). While CaV1.2 channels dominate from a functional perspective, recent

studies have additionally noted the presence of T-type Ca2+ channels including CaV3.1 and

CaV3.2. Arguably Ca2+ influx through T-type channels could, like their L-type counterpart,

facilitate the global rise in cytosolic [Ca2+] needed to drive smooth muscle cell contraction and

arterial constriction. Oddly, however, studies centered on CaV3.2 have noted that blockade of

these channels results in arterial constriction rather than dilation (Chen et al., 2003). To explain

these paradoxical observations, the investigators suggested that Ca2+ influx through CaV3.2 may

act in a localized manner to regulate defined Ca2+-dependent targets. It was speculated that these

Page 98: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

83

targets could range from nitric oxide synthase in endothelial cells to BKCa channels in vascular

smooth muscle (Chen et al., 2003).

In this study, we investigated whether Ca2+ influx through CaV3.2 channels can indeed

operate in a localized manner to activate a Ca2+-dependent target to influence cerebral arterial

tone. In this regard, we show through the use of vessel myography, confocal and electron

microscopy, computational modelling, and electrophysiology that flux through these channels

within a discrete domain activates the cytosolic gate of RyRs. This in turn leads to the repetitive

generation of Ca2+ sparks, finite SR Ca2+ release events that activate BKCa channels and provide

negative feedback to depolarized and constricted arteries, resulting in dilation of the arteries.

Overall, this is the first study to indicate that T-type Ca2+ channels can drive a Ca2+-induced Ca2+

release (CICR) event that is key to setting arterial tone and tissue perfusion in the cerebral

circulation.

3.2 MATERIALS AND METHODS

3.2.1 Animal procedures.

Animal procedures were approved by the Animal Care and Use Committee at the

University of Calgary. Briefly, female Sprague–Dawley rats (10–12 weeks of age) were killed

via carbon dioxide asphyxiation. The brain was carefully removed and placed in cold phosphate-

buffered (pH 7.4) saline solution containing (in mM): 138 NaCl, 3 KCl, 10 Na2HPO4, 2

NaH2PO4, 5 glucose, 0.1 CaCl2 and 0.1 MgSO4. Middle and posterior cerebral arteries were

carefully dissected out of surrounding tissue and cut into 2–3 mm segments.

Page 99: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

84

3.2.2 Vessel myography and measurements of membrane potentials.

Arterial segments were mounted in a customized arteriograph and superfused with warm

(37 oC) physiological salt solution (PSS; pH 7.4; 21% O2, 5% CO2, balance N2) containing (in

mM): 119 NaCl, 4.7 KCl, 20 NaHCO3, 1.1 KH2PO4, 1.2 MgSO4, 1.6 CaCl2 and 10 glucose (Kuo

et al., 1991; Welsh et al., 2000). To limit the endothelium’s tonic dilatory influence on myogenic

tone development (Knot et al., 1999; Kuo et al., 1991), air bubbles were passed through the

vessel lumen (1–2 min); successful removal of the endothelium was confirmed by the loss of

bradykinin-induced dilation. Arteries were equilibrated for 30 min at 15 mmHg and contractile

responsiveness assessed by briefly exposing (~10 s) the tissue to 60 mM KCl. Following

equilibration, intravascular pressure was increased incrementally from 20 to 100 mmHg and

arterial tone was monitored under control conditions and in the presence of paxilline (1 µM,

BKCa inhibitor) and/or Ni2+ (50 µM, CaV3.2 blocker). Maximal arterial diameters were

subsequently ascertained in Ca2+-free PSS (zero externally added Ca2+ + 2 mM EGTA). Smooth

muscle membrane potential (Vm) was assessed by inserting a glass microelectrode backfilled

with 1 M KCl (tip resistance =120–150 MΩ) into the vessel wall. Independent measurements of

Vm were first made at 60 mmHg and then in the presence of Ni2+ (50 µM), a CaV3.2 blocker. The

criteria for successful cell impalement included: (1) a sharp negative Vm deflection upon entry;

(2) a stable recording for at least 1 min following entry; and (3) a sharp return to baseline upon

electrode removal.

Page 100: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

85

3.2.3 Immunohistochemistry.

Rats were anaesthetized with sodium pentobarbital and perfused intra-cardially with 250

ml of 0.1 M phosphate-buffered saline solution (PBS, pH 7.4), followed by 100 ml of 4%

paraformaldehyde (pH 7.4) in PBS at room temperature. Brains were then removed and post-

fixed in 4% paraformaldehyde for 1 h at 22 °C. Posterior and middle cerebral arteries (MCAs)

were subsequently excised and placed in a conical vial with a standard working PBS solution

containing 3% goat serum, 0.1% Tween, and 1% dimethylsulphoxide (DMSO). Primary

antibodies against smooth muscle actin (monoclonal antibody, 1:25 dilution), CaV1.2 (1:25

dilution), CaV3.1 (1:100 dilution), CaV3.2 (1:100 dilution), BKCa (1:100 dilution), IP3R (1:100

dilution), neurofilament (1:100 dilution) and RyR2 (1:100 dilution) were added to the working

solution and incubated for 48 h at 4 °C. The Ca2+ channel and RyR antibodies were polyclonal.

Next, the tissues were washed 3 times with PBS (15 min each, 22 °C) and then incubated with

working PBS solution containing secondary antibodies for 4 h at 22 °C. Following a final set of

washes with PBS (3X), sections were mounted on gel-coated slides and covered with antifade

medium (90% glycerol/ PBS/ 0.1% p-phenylenediamine; pH 10). Coverslips were sealed with

nail polish and slides were stored at -20 °C. Controls consisted of omitting the primary

antibodies or preabsorbing them with an excess of purified antigen peptide. All reactions

involved the use of fluorophore-conjugated secondary antibodies (Alexa Fluor 488-goat anti-

mouse IgG (1:1000 dilution), Alexa Fluor 555-goat anti-rabbit IgG (1:1000 dilution), Alexa

Fluor 405-goat anti-chicken IgG (1:1000 dilution)). Immunolabelling was assessed using an

Olympus FV300 BX50 confocal microscope equipped with (405 λ, blue), (555 λ, red) and (488

λ, green) filter sets.

Page 101: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

86

3.2.4 Electron tomography

Rats were anaesthetized with sodium pentobarbital and perfused intra-cardially with 250

ml of 0.1 M phosphate-buffered saline solution (PBS, pH 7.4), followed by 100 ml of 4%

paraformaldehyde (pH 7.4) in PBS at room temperature. Brains were then removed and post-

fixed in 4% paraformaldehyde (22 °C) for 1 h. Cerebral arteries were dissected free from

surrounding tissues. Samples were immersed in 1.6% paraformaldehyde and 2.5%

glutaraldehyde in 0.1 M sodium cacodylate buffer (Sabatini et al.,1962) at pH 7.4 at 4 °C

overnight. After washing three times with the buffer (0.1 M sodium cacodylate), the samples

were post-fixed in 1% osmium tetroxide buffered with sodium cacodylate for 1 h at room

temperature, dehydrated through a graded series of acetone concentrations (30 – 100%) and

embedded in Epon 812 mixture resin. Thick sections (300-400 nm) were cut on a Reichert-Jung

Ultracut E microtome using a diamond knife and collected on single hole grids with Formvar

supporting film. The sections were first stained with 20 % aqueous uranyl acetate/Reynolds’s

lead citrate (Reynolds, 1963) and then placed on one side of a transmission electron microscopy

slot grid (1 × 2 mm slot) covered with a continuous Formvar film (∼40 nm) to dry (10 min).

Colloidal gold particles (10 nm diameter) were then placed on both sides of the grid to serve as

fiducial markers, and a thin carbon coating was applied for mechanical stabilization and to

reduce electric charging. Once prepared, sections were viewed on a Tecnai F20 transmission

electron microscope (200 keV), regions of interest were defined, and images were captured with

a 1,024 × 1,024 charge-coupled device camera (GIF 794, Gatan, Pleasanton, CA). To perform

dual-axis transmission electron-microscopic tomography, Serial EM software (Mastronarde,

2005) was employed to capture one image per degree of sample rotation (136 degrees in total).

Page 102: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

87

Tomographic reconstruction was performed by weighted back projection using the IMOD

software package (Kremer et al., 1996; Mastronarde, 1997). This yielded a contiguous stack of

two-dimensional photomicrographs with ∼4-nm resolution. The same software was used to trace

subcellular structures on each section of the contiguous stack. We then compiled the traces to

produce a 3-D rendition of the microdomain structure.

3.2.5 Immunogold-labeling.

Rats were anaesthetized with sodium pentobarbital and perfused intra-cardially with 250

ml of 0.1 M phosphate-buffered saline solution (PBS, pH 7.4), followed by 100 ml of 4%

paraformaldehyde (pH 7.4) in PBS at room temperature. Brains were then removed and post-

fixed in 4% paraformaldehyde (22 °C) for 1 h and cerebral arteries were dissected free from

surrounding tissues. In accordance with directions provided with the Aurion immunogold reagent

kit, fixed arteries were exposed to 0.1% sodium borohydride in PBS (15 min), 0.05% Triton X-

100 (30 min), Aurion blocking solution 1 h at 4 oC and then PBS (2 x 10 min). Primary

antibodies (1:100 dilution) were subsequently added to the buffer and incubated for 48 h at 4 °C.

Tissues were subsequently washed with buffer (6 x 10 min), exposed to incubation medium

containing ultra-small gold labelled secondary antibody particles (0.8 nm labeled goat anti-

rabbit, 1:100 dilution, 48 h at 4 °C) and then washed with incubation buffer (6 x 10 min) and

PBS (2 x 10 min). Prepared arteries were fixed again in PBS containing 2.5% glutaraldehyde (2

h) and then washed in PBS (2 x 10 min). Arteries were sequentially processed as follows: 1)

Enhancement conditioning solution (ECS, 4 x 10 min); 2) Silver enhancement solution (2 h); 3)

0.03 M sodium thiosulphate in ECS (10 min); 4) ECS wash (4 x 10 min); and 5) PBS wash (2 x

Page 103: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

88

10 min). Tissues were then postfixed for 1 h in a 1% osmium tetroxide-PBS solution, dehydrated

in ethanol, and embedded in Epon resin. Ultrathin sections of about ~70 nm were cut, lightly

stained with 2% aqueous uranyl acetate and Reynolds lead citrate (15 min) (Reynolds, 1963),

and viewed/photographed using a Hitachi H7650 transmission electron microscope (80 keV) and

an AMT 16000 digital camera.

3.2.6 Isolation of arterial smooth muscle cells.

Smooth muscle cells from middle and posterior cerebral arteries were enzymatically

isolated as previously described (Luykenaar et al., 2004; Talavera et al., 2001). Briefly, arterial

segments were placed in an isolation medium (37 °C, 10 min) containing (in mM): 60 NaCl, 80

Na-glutamate, 5 KCl, 2 MgCl2, 10 glucose and 10 HEPES with 1 mg/ml albumin (pH 7.4).

Vessels were then exposed to a two-step digestion process that involved: 1) 15 min incubation in

isolation medium (37 °C) containing 0.6 mg/ml papain and 1.8 mg/ml dithioerythritol; and 2) a

15 min incubation in isolation medium containing 100 μM Ca2+, 0.7 mg/ml type F collagenase

and 0.4 mg/ml type H collagenase. Following treatment, tissues were washed repeatedly with

ice-cold isolation medium and triturated with a fire-polished pipette. Liberated cells were stored

in ice-cold isolation medium for use the same day.

3.2.7 Proximity ligation assay (PLA).

Freshly isolated smooth muscle cells were studied using the Duolink in situ PLA

detection kit. Briefly, cells were fixed in phosphate buffered saline (PBS) containing 4%

paraformaldehyde for 15 min, permeabilized in PBS containing 0.1% Tween for 15 min and

Page 104: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

89

quenched in PBS containing 100 mM glycine for 5 min. Cells were then washed with PBS,

blocked for 30 min at 37 °C in Duolink blocking solution, and incubated overnight at 4 °C with

pairs of primary antibodies in Duolink antibody diluent solution (rabbit anti-CaV3.2 and mouse

RyR2). Control experiments were done with no primary antibody or only one primary antibody.

Cells were labelled with Duolink PLA PLUS and MINUS probes for 2 h at 37 °C. The secondary

antibodies of the PLA PLUS and MINUS probes are attached to synthetic oligonucleotides that

hybridize when in close proximity (i.e. ≤ 40 nm separation). The hybridized oligonucleotides are

then ligated prior to rolling circle amplification. The concatemeric amplification products

extending from the oligonucleotide arm of the PLA probes were detected using red fluorescent

fluorophore-tagged, complementary oligonucleotide sequences and a Zeiss Apotome

epifluorescence microscope. Thus the red fluorescence product is localized to sites where the two

different primary antibodies bind with separation ≤ 40 nm.

3.2.8 Computational modeling.

We have constructed a mathematical model, which incorporates important ultrastructure

and Ca2+ handling features of the vascular smooth muscle cell in cerebral arteries. The smooth

muscle cell is described as a cylinder, 80 µm in length and 5 µm in diameter, as shown in Fig.

21A. For simulation purposes, the cell is subdivided into segments 8.5 µm in length. Based on

microscopic data, in each segment there is a circumferential band, which is interrupted in

approximately two locations, thus forming two discrete sections circumferentially. In the model,

each segment is therefore further sub-divided into two semi-cylindrical “slices”. Simulations

presented in this paper are based on the behaviour of one of these slices. The frequencies of Ca2+

Page 105: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

90

spark-like events have been scaled to represent the whole cell, based on the assumption that

individual slices act independently. The model includes mechanisms responsible for Ca2+

dynamics in the cerebral arterial smooth muscle cell. This includes the Na+-Ca2+ exchanger

(NCX), plasma membrane Ca2+ ATPase (PMCA), sarco/endoplasmic reticulum Ca2+ ATPase

(SERCA), ryanodine receptor (RyR), calmodulin, calsequestrin, and the Ca2+ channels CaV1.2,

CaV3.1 and CaV3.2. In its present form, the model is thus limited to investigations of Ca2+

dynamics in response to voltage-clamp stimuli. As illustrated in Fig. 21, the model explicitly

includes the interaction between CaV3.2 and RyR in a restricted microdomain (“subspace”),

representing the 10-15 nm space between caveolae in the plasma membrane and the SR.

Mathematically, the model comprises 12 ordinary differential equations, which were solved

using the “ode15s” ODE solver in MATLAB (The MathWorks, Natick, MA, USA).

3.2.9 Electrophysiology.

Perforated patch-clamp electrophysiology was used to measure whole-cell K+ currents in

isolated smooth muscle cells. The bathing solution contained (in mM): 134 NaCl, 4 KCl, 2

MgCl2, 2 CaCl2, 10 glucose, and 10 HEPES (pH 7.4). The pipette solution contained (in mM):

110 K aspartate, 30 KCl, 10 NaCl, 2 MgCl2, 10 HEPES, and 0.05 EGTA (pH 7.2) with 200

mg/ml amphotericin B. Membrane currents were recorded while the cells were held at a steady

membrane potential of -40 or -20 mV. Whole-cell currents were recorded on an Axopatch 200B

amplifier (Axon Instruments, Inc), filtered at 1 kHz, digitized at 5 kHz and stored on a computer

for subsequent analysis with Clampfit 10.2 software. Cell capacitance ranged between 14-18 pF

and was measured with the cancellation circuitry in the voltage-clamp amplifier. A 1 M NaCl–

Page 106: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

91

agar salt bridge between the reference electrode and the bath solution was used to minimize

offset potentials (< 2 mV). All experiments were performed at room temperature (20 –22 °C).

3.2.10 Ca2+ sparks.

Ca2+ sparks were measured in rat myocytes in situ using the en-face preparation

technique (Hadley et al., 2012) with the Ca2+-sensitive dye Fluo-4 AM using a Zeiss LSM 710

NLO laser scanning confocal imaging workstation on an inverted microscope platform (Zeiss

Axio Observer Z1). Fluo-4 AM was dissolved in DMSO and added from a 1 mM stock to the

arterial suspension at a final concentration of 10 µM, along with 0.1% pluronic F127 for 1 - 1.5 h

at room temperature in the dark in balanced salt solution. Arterial segments were then washed for

30 min to allow for dye de-esterification and then cut into linear strips. The arterial segments

were pinned to Sylgard blocks and placed in an open bath imaging chamber mounted on the

confocal imaging stage. Cells were illuminated at 488 nm with a krypton argon laser and the

emitted light was collected using a photomultiplier tube (PMT). Line scans were imaged at 529

frames generated every 1 s. The acquisition period for Ca2+ spark recordings was 18.9 s. The

resultant pixel size ranged from 0.11 to 0.068 µm per pixel. To ensure that sparks within the cell

were imaged, the pinhole was adjusted to provide an imaging depth of ~2.5 µm. This depth is

roughly equivalent to the width of 50% of the cell based on morphological examination of live

preparations, and was equivalent to that used in previous studies (Hadley et al., 2012; Janiak et

al., 2001). Ca2+ sparks were analyzed by visual inspection for the frequency of firing. The

number of sparks per 100 micrometer per second was computed from these observed sparks. The

Spark Master plug-in for Image J (Picht et al., 2007) was used to determine the fractional

Page 107: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

92

fluorescence intensity for the Ca2+ spark events. In these studies, the threshold for Ca2+ spark

event detection was 3.2 times the standard deviation of the background noise above the mean

background level. Prior to analysis the background fluorescence was subtracted from each image

assuming homogeneous background levels in each cell (Hadley et al., 2012).

3.2.11 Statistical analysis.

Data are expressed as means ± S.E. and n indicates the number of vessels or cells. No

more than two experiments were performed on vessels from a given animal. Where appropriate,

paired, unpaired t-tests, and one way ANOVA were performed to compare the effects of a given

condition/treatment on arterial diameter, or whole-cell current (see figure legends for specific

details). P values ≤ 0.05 were considered statistically significant.

3.2.12 Solutions and chemicals

All buffers, chemicals and reagents were purchased from Sigma-Aldrich (St. Louis, MO)

unless otherwise stated. Donkey and goat sera were purchased from Jackson Immuno-Research

(West Grove, PA). The primary antibodies against smooth muscle actin and neurofilament were

obtained from Abcam Inc. (Cambridge, MA) whereas those primary antibodies directed against

CaV1.2, 3.1, 3.2, and RyR were purchased from Alomone (Jerusalem, Israel). Secondary

antibodies, which included Alexa Fluor 488-goat anti-mouse IgG, Alexa Fluor 405-goat anti-

chicken and Alexa Fluor 555-goat anti-rabbit IgG, were obtained from Invitrogen Life

Technologies (Eugene, OR). Aurion immunogold reagents and the PLA Duolink detection kit

Page 108: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

93

were purchased from Electron Microscopy Sciences (Hatfield, PA) and Olink (Uppsala,

Sweden), respectively.

3.3 RESULTS

3.3.1 Vasomotor and electrical effects of CaV3.2 blockade.

We started our investigation by assessing the vasomotor effect of Ni2+, a selective CaV3.2

inhibitor at micromolar concentrations (Lee et al., 1999; Perez-Reyes, 2003), on pressurized

cerebral arteries. Under control conditions, arterial tone progressively increased as vessels were

sequentially pressurized from 20 to 100 mmHg (Figure 16A & 16B). Interestingly, Ni2+

blockade of CaV3.2 channels enhanced myogenic tone at pressures < 80 mmHg. This Ni2+-

induced rise in tone coincided with an arterial depolarization of 5 ± 0.9 mV (Figure 16C, 16D &

1E). This latter result suggests that Ca2+ influx through CaV3.2 channels indirectly moderates

constrictor responses by enhancing the activity of a K+ conductance involved in establishing

negative membrane potentials to oppose constrictor responses. This could include BKCa channels

whose activity is gated by the transient release of Ca2+ from the SR via ryanodine receptors

(RyR).

3.3.2. Ca2+ channel localization.

To explore the potential spatial relationship between CaV channels and RyR, we

performed immunohistochemistry on whole fixed cerebral arteries. Tissues were labeled with

antibodies directed against smooth muscle actin and a single channel of interest (CaV3.2, RyR2,

or CaV1.2). In Figure 17A, left panel, the upper composite panel indicates that smooth muscle

Page 109: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

94

actin runs lengthwise in cells and periodically disappears as this structural protein leaves the

focal plane. Intriguingly, CaV3.2 labeling ran perpendicular to the long axis of the cells and was

observed in the regions devoid of smooth muscle actin. This CaV3.2 labeling pattern is

particularly evident in the gallery subpanels. A similar labelling pattern was observed for RyR

(Figure 17A), an observation suggestive of RyR and CaV3.2 lying in close apposition to one

another. In striking contrast, CaV1.2 labeling ran lengthwise in a ribbon like fashion along

smooth muscle cells (Figure 17B, center and magnified right panel). Note that a portion of the

topical labeling observed in the latter preparations was attributed to CaV1.2 protein expression in

perivascular nerves (Figure 18D). Secondary antibody and peptide control experiments were

performed for all experiments and were negative (Figure 18). The preceding results suggest a

possible close association between CaV3.2 and RyR; we therefore performed a Proximity

Ligation Assay (PLA). Isolated smooth muscle cells were probed with primary antibodies

against CaV3.2 and RyR and then exposed to secondary antibodies tagged with synthetic

oligonucleotide strands. If the two proteins are within 40 nm of one another, the strands

hybridize and a multiplication reaction produces a visible red fluorescent product. Figure 19A-

19C indicates that a red fluorescent product was found at discrete sites in rat cerebral arterial

smooth muscle cells. This fluorescent product was observed both near and distal to nuclei. In

striking contrast, PLA product was absent from cells exposed to both secondary antibodies alone

and in experiments where one of the primary antibodies (CaV3.2 or RyR) was omitted (Figure

19D-19F; Figure 19G-19I; & Figure 19J-19L). With evidence of CaV3.2 and RyR channels

located close to one another, we performed electron tomography to better ascertain the structural

relationship between the SR and the plasma membrane. Image analysis and model

Page 110: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

95

reconstructions of tomograms revealed a repetitive microstructure along the length of the smooth

muscle cell, comprised of caveolae and SR-like structures (Figure 20A-20C). Immuno-gold

labeling was subsequently performed on whole mounted tissues to further delineate CaV3.2 and

RyR localization. As indicated in Figure 20D-20F, CaV3.2 labeling was observable in caveolae.

In comparison, a portion of RyRs was localized to internal membranous structures that lay in

close apposition to caveolae. In this study, all the figures presented have the same magnification

values of 12000X or 30000X. For each antibody 5 figures were chosen. The images were

selected for the analysis if they showed a complete smooth muscle cell. Table 2 shows results of

analysis for each image. For RyR labelling 27± 2 % of the labeling occurred near the caveolae;

for CaV3.2 32 ± 4 % of the labelling was in the PM in caveolae. Cumulatively, the preceding

results indicate the presence of microdomains in cerebral arterial smooth muscle comprised of

caveolae and SR and which contain T-type Ca2+ (CaV3.2) and RyR channels, respectively.

Page 111: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

96

Image #

RyR CaV3.2

Number of

labelled

points

Number of

points near

caveolae

% of points

near

caveolae

Number of

labelled

points

Number of

points in

caveolae

% of points

in caveolae

1 45 13 29 124 29 23

2 70 19 27 114 30 26

3 30 9 30 107 35 33

4 43 9 21 148 49 33

5 37 10 21 155 70 45

Average 27 32

Standard

Error

2 4

Table 2: Immunogold-labeling for RyR and CaV3.2

Page 112: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

97

3.3.3 Computational modeling.

To predict whether Ca2+ flux through CaV3.2 channels could possibly initiate a CICR-like

response by activating RyR, we built a microdomain model (Figure 21A) using our structural

data, existing electrophysiological measurements of CaV channels (Aslanidi et al., 2010; Liao et

al., 2005) and published mathematical representations of RyR (Keizer & Levine, 1996). Figure

21B indicates that when voltage is stepped from -60 to -40 mV, repetitive Ca2+ spark-like events,

with an amplitude of 30 µM and a frequency of 0.12 Hz, were predicted to occur in the plasma

membrane subspace. These Ca2+ spark-like events were eliminated or reduced in frequency by ~

50-60 % in simulations in which RyR and CaV3.2 were respectively blocked (Figure 21C &

21D). Additional simulations in Figure 21E show that the frequency of Ca2+ spark-like events is

graded with depolarization. Our computational work supports the view that CaV3.2 channels

could drive a CICR response. In theory, these discrete events could activate BKCa channels and

moderate cerebral arterial tone.

3.3.4 BKCa currents and Ca2+ sparks.

To explore the relationship between CaV3.2, RyR and BKCa channels, perforated patch

clamp electrophysiology was used to monitor spontaneous transient outward current (STOCs),

events that arise when SR-driven Ca2+ sparks activate a discrete pool of BKCa channels. Under

control conditions (in the absence of any blockers), repetitive STOC generation was observed in

rat cerebral arterial smooth muscle cells with the frequency increasing as the holding potential

was depolarized from -40 to -20 mV. In keeping with CaV3.2 channels facilitating STOC

generation, 50 µM Ni2+ (a selective CaV3.2 inhibitor) reduced the frequency of these events by

Page 113: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

98

~60% (Figure 22A). In comparison, STOCs were eliminated by 1 µM paxilline, a BKCa channel

inhibitor (Figure 22B). Fluo-4 confocal imaging was subsequently used to monitor Ca2+ sparks,

with line scan analysis revealing that these events were present in rat cerebral arteries (Figure

22C). In regions where Ca2+ sparks were detectable, Ni2+ application was found to reduce event

frequency by 50% (Figure 22D). Thus 50 µM Ni2+ reduces the frequency of both Ca2+ sparks and

STOCs. Cumulatively, these results indicate that Ca2+ influx through CaV3.2 channels can indeed

enhance BKCa channel activity via a CICR-driven process that involves RyR and Ca2+ spark

induction.

3.3.5 Functional implications of CaV3.2 blockade.

If CaV3.2 can drive STOC activity, the functional consequences of this event would be

the generation of a modest hyperpolarization that feeds back negatively upon constriction as

indicated by the data in Figure 16A. To further test this possibility additional experiments were

performed and provided evidence to support this concept. Like the Ni2+-induced constriction

observed in Figure 16A and 16B and reprinted for clarity in Figure 23A and 23B, paxilline (1

µM, BKCa channel blocker) enhanced myogenic constriction at intravascular pressures < 80

mmHg (Figure 23C and 23D). When Ni2+ and paxilline were sequentially added on top of one

another, the first agent induced constriction while the second had no additive effect (Figure 23E-

23H). These latter results are consistent with CaV3.2 and BKCa channels working in series with

one another within a single signaling pathway to limit constriction.

Page 114: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

99

3.4 DISCUSSION

T-type Ca2+ channels are present in the cerebral circulation but their role in tone

development is controversial. Functional observations vary widely with some results indicating

they play no significant role (Oguri et al., 2010; Perez-Reyes, 2003; Rodman et al., 2005) while

others have linked Ca2+ influx to vasoconstriction (Akaike et al., 1989; Hansen et al., 2011;

Poulsen et al., 2011) and to vasodilation (Chen et al., 2003). Past work has noted that the role of

CaV3.2 in mediating either vasoconstriction or vasodilation is not clear. In this study, we

employed functional, structural, computational and electrical approaches to show that CaV3.2

facilitates cerebral arterial dilation. Our work suggests CaV3.2 channels are part of a restricted

microdomain along with RyRs. When CaV3.2 opens, the influx of Ca2+ into the subspace

between the plasma membrane and the SR would lead to RyR gating and the generation of Ca2+

sparks. These discrete Ca2+ events initiate STOCs, BKCa channel events that induce

hyperpolarization and vasodilation. Our results indicate that Ca2+ flux through specific T-type

channels (CaV3.2) is not directed to the bulk cytosol but rather to microdomains where they

evoke a CICR and subsequently K+ channel activity to limit arterial constriction.

3.4.1 Background.

Increases in [Ca2+]i elicit smooth muscle constriction by enhancing MLC20

phosphorylation, a process controlled by the balanced activation of myosin light chain kinase

(MLCK) and myosin light chain phosphatase (MLCP) (Deng et al., 2002). Ca2+ influx through

voltage-gated Ca2+ channels is the principal regulator of [Ca2+]i and the principal voltage-gated

Ca2+ channel α1-subunits expressed in cerebral arterial smooth muscle cells are CaV1.2 (L-type),

Page 115: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

100

CaV3.1 and CaV3.2 (T-type), as shown in Chapter 2. It is well recognized that CaV1.2 plays a key

role in controlling steady-state Ca2+ entry and consequently tone development. This can be

readily observed by applying dihydropyridines and monitoring the change in [Ca2+]i and arterial

diameter (Harder, 1984; Knot & Nelson, 1998). The function of T-type Ca2+ channels in arterial

smooth muscle has yet to be convincingly determined and it is a topic of active debate (Perez-

Reyes, 2003; Rodman et al., 2005). In this regard, several studies have argued that T-type

channels are unimportant in tone generation and that their primary role centers on the regulation

of cell proliferation (Oguri et al., 2010; Perez-Reyes, 2003; Rodman et al., 2005). Others,

however, have presented alternative findings, suggesting that Ca2+ influx through T-type

channels in some cases facilitates arterial constriction but also can interestingly facilitate dilation

(Braunstein et al., 2009; Chen et al., 2003; Hansen et al., 2011; Kuo et al., 2010; Poulsen et al.,

2011). The latter result is particularly intriguing, and the authors (Chen et al., 2003)

hypothesised that the channels involved in facilitating vasodilation might be targeted to a smaller

cellular domain, where they are able to potentially control the activity of BKCa channels.

3.4.2 Mechanism of CaV3.2-induced vasodilation.

It is in this context that we began to mechanistically examine CaV3.2 and its unexpected

role in arterial tone development. Using a Ni2+ concentration that selectively targets CaV3.2, we

first observed, analogous to the findings of Chen et al. (Chen et al., 2003), the ability of this

inhibitor to constrict pressurized cerebral arteries. This Ni2+-induced constriction was

particularly prominent when vessels were pressurized between 20-60 mmHg, and was

diminished at higher pressure. With this constrictor effect being inconsistent with Ca2+ influx

Page 116: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

101

causing a bulk rise in [Ca2+]i, we hypothesized that flux might be more localized and directed

towards the regulation of a Ca2+-sensitive conductance. As such, immunohistochemistry was

performed on whole fixed arteries, and we observed CaV3.2 labeling in regions devoid of smooth

muscle actin. A similar labelling pattern was observed for RyR, an SR Ca2+ release channel that

can generate Ca2+ sparks. This labeling pattern contrasted with that of CaV1.2, which showed a

labelling pattern distinct from CaV3.2 with labelling running parallel to the long axis of the

smooth muscle cell in thin ribbon-like threads. This pattern is in keeping with the role of CaV1.2

in elevating global [Ca2+]i. A similar labelling pattern was observed for CaV3.1, BKCa, and IP3R

(see appendix). With immunohistochemical patterns indicating a potentially close association

between CaV3.2 and RyR, a proximity ligation assay (PLA) was performed. This assay entails

the application of two primary antibodies to isolated smooth muscle cells followed by the

addition of secondary antibodies conjugated to oligonucleotides. When in close proximity, the

oligonucleotide tails interact forming a circular construct that can be replicated and fluorescently

visualized as a discrete red product. Consistent with CaV3.2 and RyR residing within 40 nm of

one another, distinctive red fluorescent labeling could be observed particularly in regions of the

cell surrounding the nucleus. Some cells showed labelling closer to the surface, but they tended

to be contracted. Similar labeling was absent from cells exposed to secondary antibodies alone or

where one of the primary antibodies had been omitted from the ligation assay. Further

experiments would need to be carried out to investigate whether the methods for cell fixation

might affect the labelling and, if so, to find an optimal method. As CaV3.2 resides on the plasma

membrane and RyR resides on the SR, we subsequently employed electron microscopy to better

define regions of close association between the SR and the plasma membrane. Electron

Page 117: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

102

tomography enables the construction and visualization of 3-D model structures from a

contiguous stack of 2-D photomicrographs (~3.5 nm resolution) rendered from a thick tissue

section. In regions peripheral to the nucleus, the SR came into close proximity with the plasma

membrane and, in particular, in close proximity to the caveolae. Immunogold labelling suggested

that some CaV3.2 resides in caveolae and some RyRs are found underneath caveolae.

We built a computational model to first address on a theoretical level whether Ca2+ influx

through T-type Ca2+ channels could be predicted to cause spontaneous repetitive opening of RyR

to produce Ca2+ spark-like events. The computational model was built based upon the literature,

structural measurements, localization, and measurements of channel activity (Aslanidi et al.,

2010; Bondarenko et al., 2004; Cannell & Soeller, 1997; Keizer & Levine, 1996; Liao et al.,

2005). It consists of an 8.5 µm slice of an arterial smooth muscle cell: a microdomain 15 nm in

length is comprised of SR and a subspace of appropriate volume, and membrane proteins have

been distributed accordingly. Our model shows that depolarization could indeed elicit repetitive

spark-like events, which can be graded as a function of voltage, and these events are completely

blocked by RyR blockade, and attenuated by about 50-60 % by CaV3.2 blockade. The

computational model provides theoretical support for the idea that the CaV3.2 can gate the

opening of RyR channels. This provided motivation for further investigations to determine if we

could find biological evidence to support this conclusion.

To determine whether there is an experimentally observable, functional connection

between the influx of Ca2+ via T-type channels and 1) BKCa channel activation and 2) Ca2+ spark

production, we first measured spontaneous transient outward currents (STOCs) using the

perforated patch technique in the absence and presence of CaV3.2 and BKCa channel blockers.

Page 118: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

103

Our results showed a reduction or elimination of STOCs in the presence of Ni2+ or paxilline at -

40 and -20 mV, suggesting that CaV3.2 channels might regulate BKCa channel activity. Ca2+

spark events have been detected in pressurized arteries and found to be triggered from RyR

channels in the SR (Brayden & Nelson, 1992; Knot & Nelson, 1998; Perez et al., 2001; Wellman

et al., 2002). Ni2+ (50 µM) a selective CaV3.2 blocker, reduced Ca2+ spark event frequency in

pressurized vessels by 50%, consistent with regulation of RyR activity by CaV3.2 channels. We

conclude that CaV3.2 channels could indirectly drive BKCa channel activity via SR Ca2+ release

through RyR. To investigate how modulation of signalling between CaV3.2 and BKCa channels

could affect arterial function we performed additional myography experiments with Ni2+, a

CaV3.2 blocker, and paxilline, a BKCa channel blocker. Paxilline induced a similar augmentation

of tone as did Ni2+ blockade of CaV3.2, particularly at pressures below 80 mmHg. When Ni2+ was

first applied to the preparation, paxilline did not induce further tone, and conversely tissues pre-

treated with paxilline, were unaffected by Ni2+. These results are consistent with the idea that

these two channels are linked sequentially.

3.4.3 Summary.

In summary, we provide eveidence that Ca2+ influx through one subtype of T-type Ca2+

channels operates to facilitate vasodilation. More specifically, our results are consistent with the

interpretation that localized Ca2+ flux through CaV3.2 is sufficient, within a restricted spatial

domain, to activate RyR and sequentially elicit Ca2+ sparks, leading to BKCa channel activation

and arterial hyperpolarization and vasodilation. Our results suggest that this mechanism

attenuates myogenic tone in rat vascular cerebral arteries, the key constrictor response that sets

Page 119: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

104

basal tone and flow to the brain. This negative feedback response that counters constrictor

responses possibly extends to other critical organs such as the kidney where efferent arterioles

not only display myogenic tone but also express CaV3.2. This mechanism could also moderate

tone development in vascular beds under the control of the sympathetic nervous system, such as

the mesenteric circulation. We conclude that voltage-gated Ca2+ channels should not be viewed

as a simple influx pathway that directly elevates [Ca2+]i. Selected subtypes of VGCCs could

work in a localized manner to discretely regulate signaling pathways essential to setting the

contractile state of vascular smooth muscle.

3.4.4 Limitations and future directions.

The current research suggests a unique role for a VGCC in a negative feedback

mechanism. There are several concerns with the experiments presented in this chapter that need

to be considered however.

(i) Vessel myography approach.

Ni2+ was the only CaV3.2 blocker used in the studies of the specific role of CaV3.2 in

mediating BKCa channel activity due to the lack of other specific blockers that can differentiate

between the two subtypes of T-type Ca2+ channels that we found to be present in cerebral arteries

(Lee et al., 1999; Perez-Reyes, 2003). Ni2+ blocks CaV3.2 channels not only by binding within

the permeation pathway of the pore, as with all T-type Ca2+ channels, but also by binding to an

extracellular histidine amino acid near S4 (voltage sensor), which allosterically affects the

Page 120: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

105

channel and hence the current. This explains the high binding specificity of Ni2+ to CaV3.2

channels.

Thus the results presented in this chapter could not be confirmed by using additional

CaV3.2 blockers as there are no better specific blockers of CaV3.2 available. The availability of

more selective blockers would facilitate further differentiation and understanding of the

functional roles of CaV3.2 in vascular smooth muscle. As an alternative to the use of more non-

specific blockers, the use of siRNA or/and knockout (KO) mice (see Chapter 4) for CaV3.2 could

be applied. An siRNA approach as discussed in Chapter 2 could be used to selectively block the

expression of CaV3.2 in artery segments in culture. Arteries treated with siRNA and from control

arteries (treated with no siRNA or treated with scrambled siRNA) could be compared in vessel

myography experiments in which vessel diameter is measured as a function of increasing

intravascular pressure. If CaV3.2 functions as described above and has a vasodilatory influence,

then in the siRNA treated vessels, increasing pressure should lead to greater myogenic tone than

in vessels not treated with siRNA or in those treated with a scrambled siRNA. Similarly vessels

from cerebral arteries of CaV3.2 KO mice would be expected to develop greater myogenic tone

than those from control mice if CaV3.2 has the same function in mice as that proposed here for

rat cerebral arteries. The KO approach is discussed further in Chapter 4.

(ii) Structural approach.

In order to investigate the functional role of CaV3.2 VGCCs in the plasmalemma and

their relationship with RyRs in the SR, several different structural approaches were used.

Immunohistochemistry using whole mounted vessels showed that both CaV3.2 and RyR have

Page 121: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

106

overlapping distributions that indicate a possible relationship between them. However, several

kinds of experiments could be done to strengthen the immunohistochemistry data. For example

immunohistochemistry experiments could be carried out using isolated smooth muscle cells and

with a higher power microscope objective to better visualize the labelling in individual cells.

Primary antibodies to CaV1.2, CaV3.1 (see supplemental Figure 27), CaV3.2, and RyR could be

used to study the presence and distributions of these channels, it would also be of interest to

examine the distribution of other relevant proteins such as caveolin, BKCa (see supplemental

Figure 28) and IP3R (see supplemental Figure 29). The distributions of antibody labelling could

confirm our previous results but could also reveal distributions of the channels that were

different or more extensive than was seen when we labelled whole mounted vessels.

Experiments using isolated cells could help to rule out the possibility that the labeling seen in the

whole arteries was due to other cell types present in the arteries (e.g. endothelial cells or

perivascular nerve cells).

Dual labeling experiments (both standard and immunofluorescence and PLA

experiments) could be carried out using isolated single cells and higher magnification.

Antibodies to CaV3.2 and caveolin could be used to double label cells to further test for the

presence of CaV3.2 channels in caveolae. Moreover, a dual labelling experiment and

conventional immunofluorescence could be carried out using antibodies to CaV3.2 and RyR to

also examine possible co-localization of the two channels. Co-localization of CaV3.2 and RyR in

these experiments would strengthen our evidence of a possible relationship between the

channels. Other dual labelling experiments could be done using antibodies to CaV3.1 and CaV3.2,

CaV1.2 and RyR, or CaV1.2 and caveolin. Results of dual labeling experiments that showed co-

Page 122: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

107

localization of CaV3.2 with caveolin combined with results showing co-localization of CaV3.2

with RyR would support the hypothesis that Ca2+ influx through CaV3.2 could trigger Ca2+

release through RyR. If we found co-localization of CaV1.2 and caveolin it might suggest that

CaV3.2 is not the only VGCC that could influence RyR function. From our hypothesis we might

expect not to see co-localization of CaV3.1 with CaV3.2 or of CaV3.1 with caveolin. However, if

we do see some co-localization, this might indicate additional functions for the two channel

types. Furthermore, triple labelling for different channels could also be done (e.g. CaV3.2, RyR,

caveolin; CaV3.1, RyR and caveolin; CaV3.1, CaV3.2 and caveolin; CaV1.2, RyR, and caveolin)

and could provide more information about the possible function of these channels in arterial

smooth muscle.

A PLA assay was used to determine if CaV3.2 channels and ryanodine receptors existed

in cells in close proximity to each other. Results of the assay indicate that there are populations

of both channels in close proximity to each other (≤ 40 nm). PLA results did not show discrete

punctate signals as it might have been expected to however. The manufacturers of the assay

(Olink Bioscience) suggest that this may be because of high expression levels of the studied

proteins. This could cause a high amount of labelling and lead to blurred images (Olink

Bioscience).

Electron tomography, using a series of images generated with a transmission electron

microscope, was used to provide a picture of a possible arrangement of the cell plasma

membrane and near membrane intracellular structures. Our results suggest the presence of a

microdomain structure in vascular smooth muscle cells. In these experiments cerebral arteries

were isolated, fixed in paraformaldehyde/glutaraldehyde, dehydrated and then embedded in Epon

Page 123: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

108

resin prior to being cut into thick sections for analysis. It is possible that the preparation of the

sample introduces artefacts in the structures actually present in the cells of the arteries. Therefore

the electron tomography experiments could possibly be enhanced by using frozen-hydrated

samples (Cryo-EM), where samples are rapidly frozen at a very low temperature (near -196 οC)

so that water becomes vitrified and not crystallized inside the tissue and cells; this could give

better preservation of structures within the cells, and the cells will be less prone to damage by the

electron beam of the electron microscope.

Electron microscopy using immunogold labelled secondary antibodies and primary

antibodies to CaV3.2 and to RyR was used to look at the distribution of the two proteins in thin

sections of cerebral arteries. My results indicated that some CaV3.2 channels are localized in the

plasma membrane in caveolae and in close proximity to RyR in the SR and that about one third

of the labelling of CaV3.2 and RyR was associated with caveolae-like or SR-like structures,

respectively.

. The specificity of the primary antibodies was tested in competition experiments in

which the antibodies were pre-incubated with peptide antigens. However, further work could be

done to investigate the distributions of CaV3.2 and RyR.

To investigate the possibility that the distributions of CaV3.2 and RyR labelling that we

saw could be due to random labelling by the two antibodies, an EM image could be artificially

(via computer) labelled with random distributions for the two antibodies using the same amounts

of labelling as measured experimentally. Different size points could be used to represent the

immunogold labelling by the two antibodies and the degree of co-localized points and the

association of points with caveolae and SR structures could be determined. These experiments

Page 124: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

109

would provide an estimate of the apparent co-localization or association with caveolae and SR in

the same areas that would be expected to occur due to random labelling by the antibodies. If the

average of the results of repeated random distribution experiments showed the same degree of

co-localization that was observed in the electron microscopy experiments this would argue

against the idea that CaV3.2 and RyR are co-localized. Similarly if the random distribution

experiments showed the same degree of association of points with caveolae- or SR-like structure

as were observed in EM experiments this would argue against the idea that the proteins are

localized in areas where the SR and caveolae are in close proximity.

Co-immunoprecipitaion (Co-IP) is a technique used to investigate protein-protein

interactions, and could be used to investigate the presence of CaV3.2 and proteins associated with

CaV3.2. The presence of caveolin in the Co-IP samples would be a strong indication that CaV3.2

could be found in caveolae. Secondly, the co-localization of Cav3.2 and RyR could be examined

by fluorescence resonance energy transfer (FRET), where the two proteins of interest are labeled

with different fluorochromes, such that the emission spectrum of one overlaps the absorption

spectrum of the other. If the two fluorochromes are within a certain distance of each other (∼ 10

nm) a FRET signal would be observed.

(iii) STOCs and Ca2+ sparks.

STOCS were measured by using perforated patch clamp at both - 40 and -20 mV for

control experiments, and in the presence of Ni2+ and paxilline. The limitation of using Ni2+ as the

only blocker available has been highlighted before. Meanwhile, paxilline was used as BKCa

channel blocker in the study to compare its effect to the effect of Ni2+. The BKCa channel

Page 125: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

110

blocker, iberiotoxin, a very highly selective inhibitor of BK channels (more selective than

paxilline; Galvez et al., 1990) could be used and compared to the results of paxilline to see if

there is any difference in the blocking between the two blockers. Furthermore, STOCS could be

measured at different voltage ranges (e.g. from -50 mV to +10 mV) to monitor their production

in control cells and in the presence of CaV3.2 blockers (Ni2+) and BKCa blockers (paxilline and/

or iberiotoxin) to test if there is any contribution by other channels in these voltage ranges.

Ca2+ spark production was measured using the fluorescent Ca2+ dye fluo-4 and confocal

microscopy. The dissociation constant (Kd) for Ca2+ binding to fluo-4 is 0.345 nM (Life

Technologies, 2010) and the fluorescence of the dye increases when the dye binds Ca2+. One

concern with the use of fluo-4 to measure Ca2+ sparks is that the dynamic range of the dye might

not be great enough to accurately measure the magnitude of some Ca2+ sparks. Recently,

Figueroa et al., 2013 adapted this method to include the simultaneous use of two fluorescent

dyes (fluo-4 and fluo-4FF) to measure Ca2+ sparks. Fluo-4FF has the same spectral properties as

fluo-4 but the Kd for Ca2+ binding to the dye is 9.7 μM (Life Technologies, 2010). The use of

the two dyes together increases the range of magnitudes of Ca2+ sparks that can be measured

(Figueroa et al., 2013).

Moreover, the limitations of using Ni2+ as CaV3.2 channel blockers could be overcome by

using a siRNA approach or by using CaV3.2 channel knockout mice to measure STOCs and Ca2+

sparks in control and treated vessels or knockout mice. The use of knockout mice is further

discussed in Chapter 4.

Page 126: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

111

(iv) Computational approach.

The mathematical model that has been constructed for Ca2+ handling in vascular smooth

muscle cells of cerebral arteries was used to test the possibility, on a theoretical level, that T-type

VGCC subtype CaV3.2 and RyR Ca2+ channels are able to induce CICR, hence Ca2+ spark

activity. The model could be further developed with information about channel distribution and

channel density, once this information is available from the results of future experiments.

Page 127: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

112

3.5 Figures.

Page 128: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

113

Figure 16: Effect of Ni2+ on pressure-induced tone and membrane potential (Vm) in rat

cerebral arteries. A&B) Middle and posterior cerebral arteries were pressurized from 20-100

mmHg while diameter was monitored under control conditions, in the presence of the CaV3.2

blocker Ni2+ (50 µM), and in Ca2+-free medium. Representative traces (A) and summary data (B,

n = 7 arteries from 6 rats) reveal the constricting effects of Ni2+. Arterial Vm in pressurized

cerebral arteries (60 mmHg) in the absence (C) and presence of 50 µM Ni2+ (D). Representative

traces are shown in (C) and (D). Summary data is shown in (E), where n = 6 arteries from 6 rats.

* denotes significant difference from control; P value ≤0.05 (paired t-test, and one way ANOVA

(panel (A) and panel (B) performed by Rasha Abd El-Rahman and panels (C-D) by Suzanne

Welsh).

Page 129: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

114

Page 130: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

115

Figure 17: Ca2+ channel localization in rat cerebral arteries. Cerebral arteries were labeled

with antibodies against smooth muscle actin (green) and CaV3.2/RyR/CaV1.2 (red). In (A, top),

CaV3.2 and RyR labeling (arrowheads) run perpendicular to the long axis of smooth muscle cells

(long white arrow) in regions devoid of actin. In (17A, bottom, the same as 17A upper) smooth

muscle actin (green) and CaV3.2/RyR (red) are presented in individual gallery panels. In (B),

CaV1.2 labeling (arrowheads) runs parallel to the long axis of smooth muscle cells (long white

arrow). Photomicrographs are representative of 3 individual experiments (Performed by Rasha

Abd El-Rahman).

Page 131: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

116

Page 132: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

117

Figure 18: Controls of smooth muscle actin, CaV3.2, RyR, and CaV1.2. Cerebral arteries were

labeled with antibodies against smooth muscle actin and CaV3.2/RyR/CaV1.2. Secondary

antibody and peptide control experiments were negative for labeling of smooth muscle actin and

CaV3.2/RyR/CaV1.2 in panels A, B, and C. In (D) cerebral arteries were labeled with antibodies

against neurofilament protein (purple) in perivascular nerves that are embedded in the vascular

smooth muscle cell layer. Secondary antibody control experiments are where both primary

antibodies are absent. Peptide control experiments utilized a peptide that blocks the specific

epitope of the primary antibody. Peptides were added to primary antibodies overnight at a

dilution 1: 1, then added to tissue samples. Photomicrographs are representative of 3 individual

experiments (Performed by Rasha Abd El-Rahman).

Page 133: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

118

Page 134: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

119

Figure 19: Proximity ligation assay and the colocalization of CaV3.2 and RyR in rat

cerebral arterial smooth muscle cells. Isolated smooth muscle cells labeled with primary

antibodies against CaV3.2 and RyR were subjected to a proximity ligation assay. A-C) A gallery

representation, where (A) is differential interference contrast (DIC), (B) nuclei labeled with

Hoechst 33342 blue stain, where (C) reveals the presence of red fluorescent product consistent

with CaV3.2 and RyR localizing within 40 nm of one another. For orientation purposes, cell

nuclei were labeled with blue stain Hoechst 33342 blue. D-F) Proximity ligation assay in the

absence of both primary antibodies but in the presence of both secondary antibodies, where (D)

is DIC, (E) is the stained nuclei, and (F) shows no red fluorescent product. G-I) Proximity

ligation assay in the absence of one primary antibody (no CaV3.2 antibody only RyR antibody)

but in the presence of both secondary antibodies, where (G) is DIC, (H) is the stained nuclei, and

(I) shows no red fluorescent product. J-L) Proximity ligation assay in the absence of one primary

antibody (no RyR antibody only CaV3.2 antibody) but in the presence of both secondary

antibodies, where (J) is DIC, (K) is the stained nuclei, and (L) shows no red fluorescent product.

Scale bars are 10 μm and the optical section depth in each image is 0.3 – 0.5 μm.

Photomicrographs are representative of ~20 smooth muscle cells in 2-3 cell preparations

(Performed by Rasha Abd El-Rahman).

Page 135: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

120

Page 136: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

121

Figure 20: Electron microscopic imaging of rat cerebral arterial smooth muscle cells. A-C)

3-D tomographic imaging was performed on posterior cerebral arteries to ascertain the presence

of microdomains. In (A), tomography performed on a 300 nm thick section of tissue to generate

a contiguous stack of 2-D photomicrographs with ~3.5 nm resolution. Subcellular structures

were traced on each section. In (B) and (C), these tracings were compiled to produce 3-D

renditions. Note that discrete membranous regions can be observed where caveolae and SR

come in close apposition to one another. Photomicrographs are representative of 3 different

arterial preparations. D-F) Immunogold labeling of RyR and CaV3.2 channels in rat cerebral

arteries. In (D), RyR labeling (red arrows) can be observed underneath the plasma membrane

near caveolae. Boxed area in top panel was magnified and placed underneath. In (E), CaV3.2

labeling (red arrows) is observed near the plasma membrane in association with caveolae (the

boxed area was magnified in the panel beneath it). In (F) control experiment (no primary

antibody) reveals an absence of electron dense particles (the boxed area was magnified in the

panel beneath it). This photomicrograph is representative of 3 different arterial preparations

(Performed by Rasha Abd El-Rahman).

Page 137: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

122

Page 138: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

123

Figure 21: A microdomain model of smooth muscle Ca2+ dynamics. Using tomographic,

immunolabeling and electrophysiological data, along with mathematical models of channel

kinetics, a microdomain model of Ca2+ dynamics has been created. In (A), the model consists of

an 8.5 µm slice of an arterial smooth muscle cell. The microdomain is 15 nm in length and is

comprised of SR and a subspace of appropriate volume. Membrane proteins have been

distributed accordingly and the level of expression set by mathematical optimization procedures.

Key proteins included were CaV1.2, CaV3.1, CaV3.2, RyR, Na+-Ca2+ exchanger (NCX), SERCA

& PMCA pumps, calmodulin and calsequestrin. With this model, Ca2+ concentrations can be

calculated in the SR, the subspace and the cytosol. In (B), simulations were performed whereby

voltage was stepped from -60 to -40 mV to induce repetitive Ca2+ spark-like events. The

frequency of these Ca2+ spark-like events was eliminated or reduced upon blockade of RyR (C)

or CaV3.2 channels (D), respectively, and increased with depolarization (E) (simulation

programming by Kamran Bigdely-Shamloo and figures by Rasha Abd El-Rahman).

Page 139: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

124

Page 140: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

125

Figure 22: Spontaneous transient outward currents (STOCs) and Ca2+ spark production in

cerebral arterial smooth muscle cells. A & B) Perforated patch clamp electrophysiology was

used to measure STOCs under control conditions and in the presence of Ni2+ (50 µM) (panel A)

and paxilline (1 µM) (panel B). Representative traces and summary data (n = 8 cells) can be

found in (A) and (B). * denotes significant difference from control; P < 0.05 (paired t-test, and

one way ANOVA). C & D) Posterior and middle rat cerebral arteries were loaded with Fluo-4

and line scan imaging performed to investigate the generation of Ca2+ sparks under control

conditions and in the presence of Ni2+ (50 µM). Representative line scan (arrows denote Ca2+

sparks) and summary data can be found in (C) and (D), respectively. * denotes significant

difference from control; P < 0.05 (paired t-test, and one way ANOVA). Experiments shown in

panels (A) and (B) were performed by Osama Harraz and Rasha Abd El-Rahman. Cell isolation

and preparation was performed by Rasha R. Abd El-Rahman, experiments shown in panels (C)

and (D) were performed by Sean Wilson and Monica Rubalcava.

Page 141: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

126

Page 142: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

127

Figure 23: Effect of Ni2+ and paxilline on myogenic tone in rat cerebral arteries. A-D) Rat

cerebral arteries were pressurized from 20-100 mmHg while diameter was monitored under

control conditions and in the presence of Ni2+ (50 µM, n = 6 arteries), paxilline (1 µM, n = 6

arteries) and Ca2+-free medium. Representative traces (A & C) and summary data (B & D) reveal

the constricting effects of Ni2+ and paxilline, where the red bars represent the difference between

the control and the addition of either Ni2+ or paxilline. * denotes significant difference from

control. E) Rat cerebral arteries were pressurized to 60 mmHg and then sequentially exposed to

Ni2+ (50 µM, n = 7) and then paxilline (1 µM, n = 6) or in the other order (G). F & H) bar graphs

demonstrate the change in diameters at different intravascular pressures (20 - 60 mmHg), where

at 60 mmHg no further change of diameter was observed when paxilline was added on top of

Ni2+ (F), or vice versa (H). * denotes significant difference from 60 mmHg; P value ≤ 0.05

(paired T-test, and one way ANOVA) (Performed by Rasha Abd El-Rahman).

Page 143: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

128

Chapter Four: DISCUSSION

4.1 Overview.

As discussed above, in vascular smooth muscle L-type voltage-gated Ca2+ channels

(CaV1.2) are believed to be the major channel regulating Ca2+ entry, subsequent contraction and

development of arterial tone (Bannister et al., 2009; Davis & Hill, 1999; Harder, 1984; Knot &

Nelson, 1998; Moosmang et al., 2003). Other studies have demonstrated, however, that other

VGCCs are present in smooth muscle cells of different vascular tissues (Braunstein et al., 2009;

Kuo et al., 2010; Nikitina et al., 2007; Oguri et al., 2010; Abd El-Rahman et al., 2013; 2009;

Feng et al., 2004; Navarro-Gonzalez et al., 2009). Although LVA Ca2+ channel currents has been

measured in resistance arteries with properties consistent with T-type channels (i.e. fast

activation/inactivation, and slow deactivation), additional information is necessary to determine

their role. It has been suggested that T-type channels function in cell proliferation and cell cycle

regulation (Oguri et al., 2010; Perez-Reyes, 2003; Rodman et al., 2005). Other investigators

have suggested that T-type Ca2+ channels play a role in myogenic tone development (Perez-

Reyes, 2003; Rodman et al., 2005). This exact role is controversial and varies between the

investigators according to the type of tissues they studied.

Evidence has been presented suggesting that Ca2+ influx through some T-type channels

facilitates arterial constriction (CaV3.1), while influx through other T-type channels (CaV3.2)

tends to dilate arteries (Braunstein et al., 2009; Chen et al., 2003; Hansen et al., 2011; Kuo et al.,

2010; Poulsen et al., 2011). With regard to the potential involvement of CaV3.2 channels in

vasodilation, some investigators have theorized that these channels might be targeted to small

Page 144: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

129

cellular domains, where they might be involved in the regulation of the activity of BKCa channels

(Chen et al., 2003).

The overall goal of this thesis was to develop a mechanistic understanding of the function

of L- and T-type VGCCs in the generation and regulation of myogenic tone in vascular smooth

muscle cells. The two main objectives were: 1) identification of the type(s) of VGCCs expressed

in vascular smooth muscle cells and their roles in the generation of myogenic tone; and 2)

determination of the specific function of one of the T-type Ca2+ channels, CaV3.2, in regulating

myogenic tone in cerebral arteries.

4.2 Objective #1: Identification of L- and T-type Ca2+ channel subtypes in rat cerebral

arterial smooth muscle cells and their roles in myogenic tone development.

Our study (Chapter 2) revealed the presence of mRNAs and all three α1-subunit proteins

of CaV1.2 (L-type) and CaV3.1/3.2 (T-type) in rat cerebral arteries (Abd El-Rahman et al., 2013),

consistent with the findings of Kuo et al. (Kuo et al., 2010). We used Ba2+ as the conducting ion

rather than Ca2+ in order to enhance the whole cell current amplitude (Abd El-Rahman et al.,

2013) and a pharmacological approach to distinguish between L- and T-type channels (Abd El-

Rahman et al., 2013). We found that nifedipine reduced the whole-cell current and the remaining

current showed a hyperpolarized shift characteristic of T-type Ca2+ channels (Iftinca & Zamponi,

2009; Perez-Reyes, 2003). This result is consistent with the findings of Nikitina et al. (Nikitina et

al., 2007) and Kuo et al. (Kuo et al., 2010). We found that T-type Ca2+ channel blockers

(mibefradil, efonidipine, and kurtoxin) reduced the nifedipine-insensitive current as expected for

T-type current. Furthermore, the activation/inactivation time constants (τ) for T-type currents

Page 145: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

130

were smaller (indicating faster activation and inactivation) than those of the L-type current,

which agreed well with previously published values for CaV3.1/3.2 and CaV1.2 (Kuo et al., 2010;

Perez-Reyes, 2003). Functional measurements were also carried out in which cerebral arteries

were pressurized and the changes in diameter of arteries were observed in control conditions and

in the presence of L-type Ca2+ channel blocker (nifedipine) and T-type Ca2+ channel blockers

(mibefradil, kurtoxin and efonidipine). The functional contribution of L-type Ca2+ channels to

arterial tone increased with pressurization, as cerebral arteries became more depolarized (-35

mV, ~80 mmHg); (Abd El-Rahman et al., 2013). This finding is consistent with the biophysical

properties of the L-type channel (Bannister et al., 2009; Knot & Nelson, 1998). Results of these

experiments are consistent with the interpretation that the contribution of T-type Ca2+ channels to

development of myogenic tone is greater when the vessel is hyperpolarized (-55 mV, ~20

mmHg) (Abd El-Rahman et al., 2013; Kuo et al., 2010; Nikitina et al., 2007; Perez-Reyes,

2003). Computational modeling of the arterial network showed that a slight change of ~ 5 to

15% in arterial diameter, similar to that observed with T-type Ca2+ channel blockade, could

possibly lead to an increase in network perfusion of ~ 20 –50%. Such results indicate that T-type

channels could have a significant effect on blood flow (Abd El-Rahman et al., 2013).

In summary, the main conclusions are: (i) both L-type (CaV1.2) and T-type (CaV3.1 and

CaV3.2) Ca2+ channels are expressed in smooth muscle cells from middle/posterior cerebral

arteries; (ii) T-type Ca2+ channels contribute to induction of the myogenic response in

hyperpolarized vessels.

Page 146: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

131

4.3 Objective #2: CaV3.2 and the induction of a negative feedback response in cerebral

arterial smooth muscle.

Based on our observation that two subtypes of T-type Ca2+ channels (CaV3.1, CaV3.2) are

present in rat cerebral artery smooth muscle cells and the suggestion that T-type channels may

mediate vasoconstriction or vasodilation we were motivated to try to separate the contribution of

these two channel types to vascular tone. In agreement with Chen et al. (Chen et al., 2003) we

observed the constriction of pressurized rat cerebral arteries upon application of the CaV3.2

blocker Ni2+. Ni2+ has an IC50 for CaV3.2 of 12 µM, and was used at a concentration (50 µM) that

would block ~ 98% of those channels, without affecting CaV3.1 or CaV1.2 (Table 1; Perez-

Reyes, 2003).

Our structural results are consistent with the presence of discrete microdomains

consisting of caveolae and SR, in which CaV3.2 and RyR, respectively, reside, which indicate

that CaV3.2 could mediate a localized Ca2+ influx rather than a global rise in [Ca2+]i.

Furthermore, computational modeling predicted, on a theoretical level, that Ca2+ influx through

T-type Ca2+ channels could possibly elicit repetitive Ca2+-induced Ca2+ release events. These in

turn could induce Ca2+ spark-like events through the activation of RyR, which initiate the

activation of large conductance Ca2+-activated K+ channels (BKCa), and hence elicit

hyperpolarization in cerebral arteries (Perez et al., 2001; Perez et al., 1999). This mechanism in

part agrees with that proposed by Kotlikoff (Kotlikoff, 2003) and Ji et al. (Ji et al., 2006), who

reported CICR in smooth muscle cells, but as “loose coupling” events (Collier et al., 2000;

Kotlikoff, 2003), where RyR opening is tied to Ca2+ channel activity in the sarcolemma.

However, CaV1.2 was proposed as the Ca2+ source mediating rises in [Ca2+]i (Collier et al.,

Page 147: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

132

2000). The biophysical properties of T-type Ca2+ channels (CaV3.2) that activate at lower

voltages, and that rapidly inactivate, suggest that they could be the source of Ca2+ that induces

CICR events. Our model predicts a complete blockade of Ca2+ sparks by blocking RyR channels

as expected. The model predicts a ~ 50 - 60% attenuation upon CaV3.2 channel blockade.

Moreover, direct measurements of spontaneous transient outward currents (STOCs) using the

perforated patch technique showed ~ 60% reduction or elimination of STOCs in the presence of

Ni2+ (CaV3.2 blocker) or paxilline (BKCa blocker). In addition, about a 50% reduction of Ca2+

spark events occurred upon the addition of Ni2+. In functional studies in which cerebral artery

diameter was measured, we found that the addition of paxilline after Ni2+ could not further

increase tone. Similarly, the addition of Ni2+ after paxilline caused no further increase in tone.

These results support and highlight the idea that these two channels are linked serially in a single

sequential pathway. These results support the hypothesis that BKCa channels are regulated by

CaV3.2 indirectly through RyR, and consequently elicit vasodilation in cerebral arteries. Our

proposed feedback mechanism is somewhat analogous to that of Earley et al., who suggested that

Ca2+ spark generation and subsequently BKCa channel activity in cerebral arterial smooth muscle

is triggered by TRPV4 channel activation (Earley et al., 2005). However, in our proposed

scenario, CaV3.2 is the main channel involved, which is compatible with its biophysical

properties of fast activation and inactivation at low voltages, and slower deactivation.

In summary, the main conclusion from Objective #2 is that CaV3.2 voltage-gated Ca2+

channels elicit vasodilation by mediating the influx of Ca2+ into a restricted microdomain to

activate RyR and sequentially cause Ca2+ sparks, which in turn activate BKCa channels and

induce arterial hyperpolarization (Figure 24).

Page 148: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

133

Figure 24: Diagram highlighting the proposed function of CaV3.2 channels in the

regulation of arterial tone. We propose that Ca2+ influx through CaV3.2 channels mediates a

dilatory response by activating ryanodine receptors (RyR) and initiating Ca2+ sparks, localized

events that activate large conductance Ca2+-activated K+ (BKCa) channels (made by Rasha Abd

El-Rahman).

Page 149: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

134

4.4 Future directions, experimental considerations, clinical relevance, and conclusion.

This thesis addressed crucial questions concerning the expression and functional role of

VGCCs in cerebral arterial smooth muscle cells. This study suggests that CaV3.1 mediates global

increases in intracellular calcium as suggested in other studies (Braunstein et al., 2009; Hansen et

al., 2011; Kuo et al., 2010; Poulsen et al., 2011) while CaV3.2 may work to alter [Ca2+]i in a

localized manner, as opposed to mediating global increases in [Ca2+]i.

In the preceding chapters additional experiments were proposed to address experimental

concerns and limitations. The future directions section below describes experiments that could be

carried out to expand on the results presented in Chapters 2 and 3. This section is followed by a

discussion of the clinical relevance of the results and a conclusion.

4.4.1 Future directions.

An additional experiment that could be added to the myography experiments in Chapters

2 and 3 would be to measure arterial diameter in the presence of an L-type channel blocker (e.g.

nifedipine) plus Ni2+ (to block CaV3.2 channels as described in Chapter 3) and to follow this by

adding one of the T-type channel blockers used in Chapter 2 (i.e. mibefradil, kurtoxin,

efonidipine). This should allow any effects of the blockers on CaV3.1 and other channels that

might be present in the preparations to be determined in the absence of the contribution of

CaV3.2 channels.

Our work could also be extended to investigate the functional role of CaV3.1 and CaV3.2

in mesenteric and pulmonary arteries and renal arterioles. Both channels have been identified in

these tissues (Braunstein et al., 2009; Hansen et al., 2011; Pluteanu & Cribbs., 2011; Poulsen et

Page 150: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

135

al., 2011). This would determine if CaV3.1 and CaV3.2 have similar roles as vasoconstrictors or

vasodilators in other vascular beds. This would also be of interest because mesenteric arteries

undergo phasic contractions compared to the tonic contraction seen in cerebral vascular and renal

arteries. The experimental approaches described in Chapters 2 and 3 (e.g. electrophysiological,

functional and structural approaches) could be applied in these studies to investigate the types of

inward Ca2+ currents in these arteries and to examine the possible expression of different T-type

channel subtypes, and their role in the development of myogenic tone. Careful experiments

might make it possible to differentiate between different splice variants of CaV3.1 and CaV3.2

and their biophysical characteristics (conductivity, channel gating, and window current

properties) in these arteries and to understand how the expression of different subtypes might be

reflected in their functional role/s.

Vessel myography could be used to study rat cerebral and other arteries that have not

been denuded of endothelium to investigate the influence that endothelial cells (e.g. through the

release of endothelium-derived hyperpolarizing factors (EDHF)) might have on the development

of myogenic tone in intact blood vessels. This work may also provide information about how the

endothelium could affect the negative feedback that we proposed might occur through the

interaction of CaV3.2 and BKCa channels in cerebral arteries.

Along with a study of the effect of the presence of the endothelium on the myogenic

response of the cerebral and other arteries discussed above and the experiments suggested above,

another line of research could be followed to investigate the presence of CaV3.1 and CaV3.2 in

endothelium. The endothelium of mesenteric and cerebral arteries expresses mRNA for CaV3.1

and CaV3.2 Ca2+ channels, whereas pulmonary micro-vascular endothelial cells express CaV3.1

Page 151: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

136

(Townsley et al., 2006; Zhou & Wu, 2006). This work could lead to further information about

the mechanism involved in the control or limitation of vasoconstriction in smooth muscle

(Harraz & Welsh, 2013; Zhou et al., 2010; Kuo et al., 2010). The influx of Ca2+ through the T-

type Ca2+ channels into the cytoplasm of the endothelial cells could activate the production of

EDHFs such as nitric oxide (NO) which would induce vasodilation of smooth muscle cells.

These proposed experiments could involve PCR, western blots, immunohistochemistry, and

electrophysiology. It should be noted, however, that patch-clamp studies of endothelial cells

from a variety of vessels failed to find any evidence for T-type Ca2+ channel current (Braunstein

et al., 2009; Kuo et al., 2010). The data available in this field of study is still incomplete,

therefore more work needs to be done, including electrophysiological and structural studies to

determine if T-type channels are present and if they have a functional role in endothelial cells.

The possible involvement of CaV3.1 Ca2+ channels in the regulation of other channels,

such as TRP channels (e.g. TRPM4 channels) would be of interest. TRP channels are involved in

membrane depolarization, activating L-type Ca2+ channels and hence they induce

vasoconstriction in vascular smooth muscle (Figures 25 and 26). TRPM4 channels are permeable

to Na2+ causing membrane depolarization; at the same time they are known to be activated by

Ca2+ release from IP3Rs (Earley et al., 2004; Gonzales et al., 2010a). CaV3.1 could activate IP3Rs

and the Ca2+ released from the SR could then elicit transient inward cation currents (TICCs) or it

could activate chloride (Cl-) channels (e.g. TMEM16A) that induce spontaneous transient inward

currents (STIC) (Harraz & Welsh, 2013). Both channels (TRPM4 and TMEM16A) can cause

membrane depolarization and hence activation of L-type VGCCs that in turn cause a global rise

in [Ca2+]i and constriction (Earley et al., 2004; Gonzales et al., 2010a) (Figures 25 and 26).

Page 152: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

137

Therefore, CaV3.1 could induce vasoconstriction while CaV3.2 could induce vasodilation.

Application of experimental approaches (patch-clamp, vessel myography) using blockers of

TRPM4 and TMEM16A (9-phenanthrol and T16Ainh-A01, respectively) could be used to

investigate the contribution of these channels to recorded currents and to investigate their role in

myogenic tone development. Structural approaches (e.g. immunogold-labelling, PLA, and

immunofluorescence) could also be used to determine the localization and possible co-

localization of TRPM4 and TMEM16A with other proteins including CaV3.1 and CaV3.2.

It would also be of interest to study the possibility of signaling complexes among T-type

VGCCs (especially CaV3.1) and members of the potassium channels superfamily (e.g. KV4

channels) to regulate the function of the latter channels. Recently it was reported that CaV3-

mediated calcium influx in cerebellar stellate cells assists in the control of the voltage-

dependence of KV4 inactivation at a nano-domain level (Turner et al., 2011). Study of this

possibility in vascular beds could thus uncover a similar nano-domain Ca2+ signaling mechanism

in these beds.

T-type Ca2+ channels might also initiate and/or maintain Ca2+ waves, asynchronous

events dependent on Ca2+ release from the SR (Braunstein et al., 2009; Martin-Cano et al.,

2009). These Ca2+ wave events are more essential in setting myogenic tone at hyperpolarized

rather than depolarized membrane potentials (Mufti et al., 2010) which might be in consistent

with T-type Ca2+ channel involvement. Using confocal microscopy to study the production of

Ca2+ waves in rat cerebral arteries under control conditions, and in the presence of T-type Ca2+

channels blockers (mibefradil, kurtoxin, efonidipine, and Ni2+) would produce information about

possible roles for CaV3.1 and CaV3.2 in Ca2+ wave production.

Page 153: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

138

- Knockout mice.

A longer term approach would be to perform experiments with knockout mice which are

already commercially available for both CaV3.1 and CaV3.2 (Chen et al., 2003; Stamboulian et

al., 2004). The knockout gene technique involves the replacement of a normal gene with a

mutant gene using a homologous recombination technique that causes either deletion or

inactivation of the normal gene of interest; the gene is said to have suffered a knockout (Zan et

al., 2003). As noted CaV3.1 and CaV3.2 knockout mice have been made, however, the vascular

smooth muscle in these animals has not been studied to any great extent. Although an abnormal

vasodilatory phenotype is observed for the aorta in CaV3.2 knockout mice, their blood pressure is

normal. Examination of CaV3.1 knockout mice indicates that the channel contributes to sinoatrial

node pacemaker activity and atrioventricular conduction (Mangoni et al., 2006), but no work has

been done in the vascular smooth muscle in these animals. To extend the present studies to the

knockouts, it would first be necessary to repeat many of the experiments that were described in

Chapters 2 and 3 on normal mice to determine if regulation of the cerebral arteries in them is

similar to that seen in rats. Depending upon the results with normal mice, experiments could then

be designed to compare normal and knockout mice. If, for example, evidence (e.g. mRNA and/or

presence of channel proteins) was obtained consistent with the presence of both L- and T-type

Ca2+ channels in the cerebral vasculature in mice, then electrophysiological and myography

experiments could be undertaken to determine the effects of the knockout. Experiments designed

to examine the localization and co-localization of these channels with other proteins of interest

could reveal structural changes in smooth muscle cells that develop as a result of the knockout of

one or both of the T-type channels.

Page 154: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

139

To illustrate some possible experiments that could be done with KO mice. I will make the

assumption that mice and rats have the same complements of VGCCs. If this actually turns out to

be the case, there are several types of experiments that could be done using knockout mice.

Cerebral artery smooth muscle cells from normal and CaV3.2 KO mice could be used in

electrophysiological experiments in which STOCs are measured. If CaV3.2 functions the way we

concluded in Chapter 3, then characteristics of STOCs (e.g. frequency) seen in normal mice

might not be the same as those seen in cerebral arterial smooth muscle cells from CaV3.2 KO

mice.

The effect of Ni2+ (at a concentration that blocks only CaV3.2) could be compared in

cerebral arteries from normal and CaV3.2 KO mice in vessel myography experiments. Ni2+ would

be expected to cause constriction in vessels from normal mice based on our conclusions from the

experiments reported in Chapter 3. In contrast, Ni2+ would be predicted to have no effect in the

cerebral arteries from CaV3.2 KO mice.

Vessel myography experiments, similar to those reported in Chapter 2, could be carried

out using cerebral arteries from normal and CaV3.2 KO mice in order to examine the effects of

T-type channel blockers. Based on the experiments reported in Chapters 2 and 3, a greater degree

of relaxation might be expected to occur in vessels from CaV3.2 KO mice in response to T-type

blockers (e.g. mibefradil, kurtoxin, and efonidipine). In vessels from normal mice, blocking

CaV3.1 channels would have a vasoconstricting effect while blocking CaV3.2 channels would

have a vasodilatory effect. As we showed in Chapter 2 the net effect of the blockers, based on the

predominant role of CaV3.1 channels, is to cause vasodilation. In vessels from CaV3.2 KO mice

only the vasodilatory effect would occur therefore the degree of relaxation might be expected to

Page 155: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

140

be greater. The results of these experiments could possibly be complicated by differences in the

resting vessel diameters and possible differences in CaV3.1 expression level between normal and

CaV3.2 KO mice.

Ca2+ sparks and Ca2+ waves could be characterized in cerebral artery smooth muscle cells

from normal, CaV3.1 KO mice and CaV3.2 KO mice, to see if Ca2+ waves occur and if their

properties vary between cells from different types of mice. The properties of the waves could

include frequency, amplitude, rate of propagation and distance travelled through the cell.

Vessel myography experiments using cerebral arteries from normal, CaV3.1 KO mice and

CaV3.2 KO mice could be compared using vessels with the endothelium present in order to

understand the role of T-type VGCCs in smooth muscle and endothelial cells. However, these

experiments might be extremely complicated by the possible presence of both CaV3.1 and

CaV3.2 in smooth muscle and endothelial cells in the normal mice.

In smooth muscle cells from both CaV3.1 and CaV3.2 KO mice the inward current

blocked by T-type channel blockers (e.g. mibefradil, kurtoxin, and efonidipine) would be

expected to be less than that blocked in cells from normal mice. If this result was not seen in

experiments then that might indicate that the blockers have non-specific effects.

In considering the use of knockout mice, however, one must be aware of some possible

limitations/difficulties with the technique. Knockout mice are made by using a technique in

which a DNA fragment containing a desired mutant gene is inserted into a vector and introduced

into a special line of embryo-derived mouse stem cells (ES) by homologous recombination. If

the mutant gene prevents the expression of the protein coded by the normal gene a knockout is

Page 156: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

141

produced. In using this technology, there are several possible problems and limitations that must

be taken into consideration. 1) In about 15% of knockout mice the knockout is either embryonic

lethal or the mice do not grow to adulthood, so the effects of the affected protein cannot be

studied in adult animals. 2) Certain genes are difficult to knockout. 3) Knocking out some genes

will affect the expression of other genes such that a specific characteristic phenotypic cannot be

distinguished. 4) In experiments using knockouts another potential problem is that cells from two

different animals (normal and knockout) are compared. 5) A major issue for the use of knockout

technology is that readily available knockout models are developed in mice. Thus results

obtained from the use of knockouts may not be readily applicable to other animals. This can be a

problem especially if work on another animal model has been extensive and it appears that the

tissues and biological systems differ significantly between these models and the mouse model. 6)

It is still an expensive approach because of the high maintenance costs needed for the mice.

Page 157: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

142

Figure 25: Diagram highlighting the different proposed possible functions of CaV3.2

channels in the regulation of arterial tone. We propose that Ca2+ influx through CaV3.2

channels mediates a dilatory response by activating ryanodine receptors (RyR) and initiating

Ca2+ sparks, localized events that activate large conductance Ca2+-activated K+ (BKCa) channels.

CaV3.2 could possibly activate TRP channels (e.g. the TRPM4 channels) to elicit transient

inward cation currents (TICCs) (made by Rasha Abd El-Rahman).

.

Page 158: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

143

Figure 26: Diagram highlighting the possible separate functions of CaV3.1 and CaV3.2

channels in the regulation of arterial tone. Ca2+ influx through CaV3.2 channels mediates a

dilatory response by activating ryanodine receptors (RyR) and initiating Ca2+ sparks, localized

events that activate large conductance Ca2+-activated K+ (BKCa) channels. Furthermore, a role for

CaV3.1 in activating Cl- channels (possibly TMEM16A channels) may be to induce spontaneous

transient inward currents (STIC) or perhaps TRP channels (e.g. TRPM4) to elicit transient

inward cation currents (TICCs) (made by Rasha Abd El-Rahman).

.

Page 159: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

144

4.4.2 Clinical relevance.

The work presented in my thesis enhanced our current understanding of the L- and T-type

voltage-gated Ca2+ channel function in cerebral arterial smooth muscle. We have proposed a

unique functional role for one of the T-type Ca2+ channels (CaV3.2) in cerebral arteries. As

discussed above there are some limitations to our experiments (the lack of specific blockers for

T-type VGCCs, the use of Ba2+ as the charge carrier when recording the inward Ca2+ current, and

the limitations of the different structural experiments) and I have suggested additional

experiments that could be done. Nevertheless, my investigation has a significant impact on

understanding the functional role of smooth muscle VGCCs in normal cerebral arteries. This new

understanding may help in comprehending the molecular basis of the function of cerebral arteries

in pathological conditions. A major pathological condition is hypertension and VGCCs are one

of the main targets in the treatment of hypertension (Godfraind et al, 1986; Belardetti, et al,

2008; Hansen et al, 2011).

Hypertension is a chronic medical condition in which blood pressure is elevated in the

arteries, hence augmenting vascular tone (Carretero & Oparil, 2000a, 2000b). Therefore,

blocking T-type Ca2+ channels per se may not be effective at treating hypertension, because at

least one of the T-type channels functions in a manner different from L-type channels, the latter

contributing to constriction while CaV3.2 contributes to dilation. To target the CaV1.2 or CaV3.2

separately, drugs that are selective in blocking CaV3.2 or CaV3.1, and CaV1.2, will have to be

developed. It may be that CaV3.2 is down-regulated in hypertension leading to the reduction in

Ca2+ spark generation. It has been noted that Ca2+ sparks are decreased in frequency in

hypertensive rat arterial cells (Amberg et al., 2003); this could be because CaV3.2 expression is

Page 160: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

145

decreased or channel regulation is altered (possibly by a mutation in the protein). Further work in

this area will be important in examining these possibilities.

4.4.3 Conclusion.

The work described in this thesis used a combination of innovative methods to

investigate the basis of Ca2+ dynamics and arterial tone regulation in vascular smooth muscle

cells. Collectively, three voltage-gated Ca2+ channels have been confirmed to be present in rat

cerebral arterial smooth muscle. Molecular and electrophysiological approaches showed that

both L-type (CaV1.2) and T-type (CaV3.1 and CaV3.2) Ca2+ channels are expressed and

electrically distinctive in cerebral arterial smooth muscle cells. This work also provided evidence

that T-type Ca2+ channels contribute to myogenic tone generation, where L-type (CaV1.2)

showed the greater contribution and T-types the lesser contribution. Furthermore, our findings

point to a unique role for a VGCC (CaV3.2) in driving local Ca2+ events (Ca2+ sparks), through a

CICR process, to regulate arterial tone in vascular smooth muscle cells. Therefore, understanding

of VGCCs should take into consideration the different roles the channels play. Furthermore, our

findings suggest that CaV3.1, identified in rat cerebral arteries, could be responsible for the

modest contribution of T-type Ca2+ channels to myogenic tone. This channel could play a more

prominent role in increasing [Ca2+]i at a more hyperpolarized state of the vessels and therefore

assists in initiating basal myogenic tone.

Page 161: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

146

APPENDIX

Figure 27. CaV3.1 protein expression in vascular smooth muscle cells.

Figure 28. BKCa protein expression in vascular smooth muscle cells.

Figure 29. IP3R protein expression in vascular smooth muscle cells.

Page 162: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

147

Page 163: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

148

Figure 27: CaV3.1 protein expression in vascular smooth muscle cells. Cerebral arteries were

labeled with antibodies against smooth muscle actin (green) and CaV3.1 (red). In (A), CaV3.1

labeling (arrowheads) runs parallel to the long axis of smooth muscle cells (long white arrow). In

(B), smooth muscle actin (green) and CaV3.1 (red) are presented in individual panels. Secondary

antibody and peptide control experiments were negative for labeling (C). Secondary antibody

control experiments were done with both primary antibodies absent. Peptide experiment utilized

a peptide that blocks the specific epitope of the primary antibody, and was added to primary

antibodies overnight at a dilution 1 : 1, then added to tissue samples. Photomicrographs are

representative of 3 individual experiments (Performed by Rasha Abd El-Rahman).

Page 164: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

149

Page 165: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

150

Figure 28: BKCa protein expression in vascular smooth muscle cells. Cerebral arteries were

labeled with antibodies against smooth muscle actin (green) and BKCa (red). In (A), BKCa

labeling (arrowheads) runs parallel to the long axis of smooth muscle cells (long white arrow). In

(B), smooth muscle actin (green) and BKCa (red) are presented in individual panels. Secondary

and peptide control experiments were negative for labeling (C). Secondary antibody control

experiments were done with both primary antibodies absent. Peptide experiment utilized a

peptide that blocks the specific epitope of the primary antibody, and was added to primary

antibodies overnight at a dilution 1 : 1, then added to tissue samples. Photomicrographs are

representative of 3 individual experiments (Performed by Rasha Abd El-Rahman).

Page 166: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

151

Page 167: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

152

Figure 29: IP3R protein expression in vascular smooth muscle cells. Cerebral arteries were

labeled with antibodies against smooth muscle actin (green) and IP3R (red). In (A), IP3R labeling

(arrowheads) runs parallel to the long axis of smooth muscle cells (long white arrow). In (B),

smooth muscle actin (green) and IP3R (red) are presented in individual panels. Secondary and

peptide control experiments were negative for labeling (C). Secondary antibody control

experiments were done with both primary antibodies absent. Peptide experiment utilized a

peptide that blocks the specific epitope of the primary antibody, and was added to primary

antibodies overnight at a dilution 1 : 1, then added to tissue samples. Photomicrographs are

representative of 3 individual experiments (Performed by Rasha Abd El-Rahman).

Page 168: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

153

REFERENCES

Abd El-Rahman, R. R., Harraz, O. F., Brett, S. E., Anfinogenova, Y., Mufti, R. E., Goldman, D., & Welsh, D. G. (2013). Identification of L- and T-type Ca2+ channels in rat cerebral arteries: role in myogenic tone development. Am J Physiol Heart Circ Physiol, 304(1), H58-71.

Akaike, N., Kanaide, H., Kuga, T., Nakamura, M., Sadoshima, J., & Tomoike, H. (1989). Low-voltage-activated calcium current in rat aorta smooth muscle cells in primary culture. J Physiol, 416, 141-160.

Altier, C., & Zamponi, G. W. (2008). Signaling complexes of voltage-gated calcium channels and G protein-coupled receptors. J Recept Signal Transduct Res, 28(1-2), 71-81.

Amberg, G. C., Bonev, A. D., Rossow, C. F., Nelson, M. T., & Santana, L. F. (2003). Modulation of the molecular composition of large conductance, Ca(2+) activated K(+) channels in vascular smooth muscle during hypertension. J Clin Invest, 112(5), 717-724.

Aslanidi, O. V., Sleiman, R. N., Boyett, M. R., Hancox, J. C., & Zhang, H. (2010). Ionic mechanisms for electrical heterogeneity between rabbit Purkinje fiber and ventricular cells. Biophys J, 98(11), 2420-2431.

Bahring, R., Milligan, C. J., Vardanyan, V., Engeland, B., Young, B. A., Dannenberg, J., Waldschutz, R., Edwards, J. P., Wray, D., & Pongs, O. (2001). Coupling of voltage-dependent potassium channel inactivation and oxidoreductase active site of Kvbeta subunits. J Biol Chem, 276(25), 22923-22929.

Ball, C. J., Wilson, D. P., Turner, S. P., Saint, D. A., & Beltrame, J. F. (2009). Heterogeneity of L- and T-channels in the vasculature: rationale for the efficacy of combined L- and T-blockade. Hypertension, 53(4), 654-660.

Bannister, R. A., Pessah, I. N., & Beam, K. G. (2009). The skeletal L-type Ca(2+) current is a major contributor to excitation-coupled Ca(2+) entry. J Gen Physiol, 133(1), 79-91.

Berridge, M. J. (1995). Capacitative calcium entry. Biochem J, 312 ( Pt 1), 1-11. Berridge, M. J., Lipp, P., & Bootman, M. D. (2000). The versatility and universality of calcium

signalling. Nat Rev Mol Cell Biol, 1(1), 11-21. Bers, D. M. (2004). Macromolecular complexes regulating cardiac ryanodine receptor function. J

Mol Cell Cardiol, 37(2), 417-429. Bielefeldt, K. (1999). Molecular diversity of voltage-sensitive calcium channels in smooth

muscle cells. J Lab Clin Med, 133(5), 469-477. Bondarenko, V. E., Szigeti, G. P., Bett, G. C., Kim, S. J., & Rasmusson, R. L. (2004). Computer

model of action potential of mouse ventricular myocytes. Am J Physiol Heart Circ Physiol, 287(3), H1378-1403.

Bootman, M. D., Missiaen, L., Parys, J. B., De Smedt, H., & Casteels, R. (1995). Control of inositol 1,4,5-trisphosphate-induced Ca2+ release by cytosolic Ca2+. Biochem J, 306 ( Pt 2), 445-451.

Braunstein, T. H., Inoue, R., Cribbs, L., Oike, M., Ito, Y., Holstein-Rathlou, N. H., & Jensen, L. J. (2009). The role of L- and T-type calcium channels in local and remote calcium responses in rat mesenteric terminal arterioles. J Vasc Res, 46(2), 138-151.

Page 169: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

154

Brayden, J. E., & Bevan, J. A. (1985). Neurogenic muscarinic vasodilation in the cat. An example of endothelial cell-independent cholinergic relaxation. Circ Res, 56(2), 205-211.

Brayden, J. E., & Nelson, M. T. (1992). Regulation of arterial tone by activation of calcium-dependent potassium channels. Science, 256(5056), 532-535.

Cannell, M. B., & Soeller, C. (1997). Numerical analysis of ryanodine receptor activation by L-type channel activity in the cardiac muscle diad. Biophys J, 73(1), 112-122.

Carafoli, E., Santella, L., Branca, D., & Brini, M. (2001). Generation, control, and processing of cellular calcium signals. Crit Rev Biochem Mol Biol, 36(2), 107-260.

Carretero, O. A., & Oparil, S. (2000a). Essential hypertension : part II: treatment. Circulation, 101(4), 446-453.

Carretero, O. A., & Oparil, S. (2000b). Essential hypertension. Part I: definition and etiology. Circulation, 101(3), 329-335.

Catterall, W. A. (1995). Structure and function of voltage-gated ion channels. Annu Rev Biochem, 64, 493-531.

Catterall, W. A. (2000). Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol, 16, 521-555.

Catterall, W. A., Perez-Reyes, E., Snutch, T. P., & Striessnig, J. (2005). International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev, 57(4), 411-425.

Chalmers, S., Olson, M. L., MacMillan, D., Rainbow, R. D., & McCarron, J. G. (2007). Ion channels in smooth muscle: regulation by the sarcoplasmic reticulum and mitochondria. Cell Calcium, 42(4-5), 447-466.

Chen, C. C., Lamping, K. G., Nuno, D. W., Barresi, R., Prouty, S. J., Lavoie, J. L., Cribbs, L. L., England, S. K., Sigmund, C. D., Weiss, R. M., Williamson, R. A., Hill, J. A., & Campbell, K. P. (2003). Abnormal coronary function in mice deficient in alpha1H T-type Ca2+ channels. Science, 302(5649), 1416-1418.

Cheng, X., Pachuau, J., Blaskova, E., Asuncion-Chin, M., Liu, J., Dopico, A. M., & Jaggar, J. H. (2009). Alternative splicing of Cav1.2 channel exons in smooth muscle cells of resistance-size arteries generates currents with unique electrophysiological properties. Am J Physiol Heart Circ Physiol, 297(2), H680-688.

Cole, W. C., & Welsh, D. G. (2011). Role of myosin light chain kinase and myosin light chain phosphatase in the resistance arterial myogenic response to intravascular pressure. Arch Biochem Biophys, 510(2), 160-173.

Collier, M. L., Ji, G., Wang, Y., & Kotlikoff, M. I. (2000). Calcium-induced calcium release in smooth muscle: loose coupling between the action potential and calcium release. J Gen Physiol, 115(5), 653-662.

Coronado, R., Morrissette, J., Sukhareva, M., & Vaughan, D. M. (1994). Structure and function of ryanodine receptors. Am J Physiol, 266(6 Pt 1), C1485-1504.

Cox, R. H., & Petrou, S. (1999). Ca(2+) influx inhibits voltage-dependent and augments Ca(2+)-dependent K(+) currents in arterial myocytes. Am J Physiol, 277(1 Pt 1), C51-63.

Cribbs, L. L. (2001). Vascular smooth muscle calcium channels: could "T" be a target? Circ Res, 89(7), 560-562.

Davies, P. F. (2009). Hemodynamic shear stress and the endothelium in cardiovascular pathophysiology. Nat Clin Pract Cardiovasc Med, 6(1), 16-26.

Page 170: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

155

Davis, M. J., & Hill, M. A. (1999). Signaling mechanisms underlying the vascular myogenic response. Physiol Rev, 79(2), 387-423.

Davis, N. F., Callanan, A., McGuire, B. B., Flood, H. D., & McGloughlin, T. M. (2011). Evaluation of viability and proliferative activity of human urothelial cells cultured onto xenogenic tissue-engineered extracellular matrices. Urology, 77(4), 1007 e1001-1007.

Deng, J. T., Sutherland, C., Brautigan, D. L., Eto, M., & Walsh, M. P. (2002). Phosphorylation of the myosin phosphatase inhibitors, CPI-17 and PHI-1, by integrin-linked kinase. Biochem J, 367(Pt 2), 517-524.

Dimopoulos, G. J., Semba, S., Kitazawa, K., Eto, M., & Kitazawa, T. (2007). Ca2+-dependent rapid Ca2+ sensitization of contraction in arterial smooth muscle. Circ Res, 100(1), 121-129.

Dolmetsch, R. E., Pajvani, U., Fife, K., Spotts, J. M., & Greenberg, M. E. (2001). Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science, 294(5541), 333-339.

Dolphin, A. C. (2006). A short history of voltage-gated calcium channels. Br J Pharmacol, 147 Suppl 1, S56-62.

Earley, S. (2010). Vanilloid and melastatin transient receptor potential channels in vascular smooth muscle. Microcirculation, 17(4), 237-249.

Earley, S., Heppner, T. J., Nelson, M. T., & Brayden, J. E. (2005). TRPV4 forms a novel Ca2+ signaling complex with ryanodine receptors and BKCa channels. Circ Res, 97(12), 1270-1279.

Earley, S., Waldron, B. J., & Brayden, J. E. (2004). Critical role for transient receptor potential channel TRPM4 in myogenic constriction of cerebral arteries. Circ Res, 95(9), 922-929.

Faraci, F. M., Baumbach, G. L., & Heistad, D. D. (1990). Cerebral circulation: humoral regulation and effects of chronic hypertension. J Am Soc Nephrol, 1(1), 53-57.

Faraci, F. M., & Heistad, D. D. (1990). Regulation of large cerebral arteries and cerebral microvascular pressure. Circ Res, 66(1), 8-17.

Feng, M. G., Li, M., & Navar, L. G. (2004). T-type calcium channels in the regulation of afferent and efferent arterioles in rats. Am J Physiol Renal Physiol, 286(2), F331-337.

Fill, M., & Copello, J. A. (2002). Ryanodine receptor calcium release channels. Physiol Rev, 82(4), 893-922.

Filosa, J. A., Bonev, A. D., Straub, S. V., Meredith, A. L., Wilkerson, M. K., Aldrich, R. W., & Nelson, M. T. (2006). Local potassium signaling couples neuronal activity to vasodilation in the brain. Nat Neurosci, 9(11), 1397-1403.

Furuichi, T., Yoshikawa, S., Miyawaki, A., Wada, K., Maeda, N., & Mikoshiba, K. (1989). Primary structure and functional expression of the inositol 1,4,5-trisphosphate-binding protein P400. Nature, 342(6245), 32-38.

Furukawa, T., Miura, R., Honda, M., Kamiya, N., Mori, Y., Takeshita, S., Isshiki, T., & Nukada, T. (2004). Identification of R(-)-isomer of efonidipine as a selective blocker of T-type Ca2+ channels. Br J Pharmacol, 143(8), 1050-1057.

Goldman, D., & Popel, A. S. (2000). A computational study of the effect of capillary network anastomoses and tortuosity on oxygen transport. J Theor Biol, 206(2), 181-194.

Page 171: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

156

Gonzales, A. L., Amberg, G. C., & Earley, S. (2010a). Ca2+ release from the sarcoplasmic reticulum is required for sustained TRPM4 activity in cerebral artery smooth muscle cells. Am J Physiol Cell Physiol, 299(2), C279-288.

Gonzales, A. L., Garcia, Z. I., Amberg, G. C., & Earley, S. (2010b). Pharmacological inhibition of TRPM4 hyperpolarizes vascular smooth muscle. Am J Physiol Cell Physiol, 299(5), C1195-1202.

Gonzalez-Cobos, J. C., & Trebak, M. (2010). TRPC channels in smooth muscle cells. Front Biosci, 15, 1023-1039.

Guibert, C., Ducret, T., & Savineau, J. P. (2008). Voltage-independent calcium influx in smooth muscle. Prog Biophys Mol Biol, 98(1), 10-23.

Hadley, S. R., Blood, Q., Rubalcava, M., Waskel, E., Lumbard, B., Le, P., Longo, L. D., Buchholz, J. N., & Wilson, S. M. (2012). Maternal high-altitude hypoxia and suppression of ryanodine receptor-mediated Ca2+ sparks in fetal sheep pulmonary arterial myocytes. Am J Physiol Lung Cell Mol Physiol, 303(9), L799-813.

Hansen, P. B., Poulsen, C. B., Walter, S., Marcussen, N., Cribbs, L. L., Skott, O., & Jensen, B. L. (2011). Functional importance of L- and P/Q-type voltage-gated calcium channels in human renal vasculature. Hypertension, 58(3), 464-470.

Harder, D. R. (1984). Pressure-dependent membrane depolarization in cat middle cerebral artery. Circ Res, 55(2), 197-202.

Harraz, O. F., & Welsh, D. G. (2013). T-type Ca(2)(+) channels in cerebral arteries: approaches, hypotheses, and speculation. Microcirculation, 20(4), 299-306.

Hirota, S., Helli, P., & Janssen, L. J. (2007). Ionic mechanisms and Ca2+ handling in airway smooth muscle. Eur Respir J, 30(1), 114-133.

Hofmann, F., Lacinova, L., & Klugbauer, N. (1999). Voltage-dependent calcium channels: from structure to function. Rev Physiol Biochem Pharmacol, 139, 33-87.

Horiba, M., Muto, T., Ueda, N., Opthof, T., Miwa, K., Hojo, M., Lee, J. K., Kamiya, K., Kodama, I., & Yasui, K. (2008). T-type Ca2+ channel blockers prevent cardiac cell hypertrophy through an inhibition of calcineurin-NFAT3 activation as well as L-type Ca2+ channel blockers. Life Sci, 82(11-12), 554-560.

Iftinca, M. C., & Zamponi, G. W. (2009). Regulation of neuronal T-type calcium channels. Trends Pharmacol Sci, 30(1), 32-40.

Jaggar, J. H., Porter, V. A., Lederer, W. J., & Nelson, M. T. (2000). Calcium sparks in smooth muscle. Am J Physiol Cell Physiol, 278(2), C235-256.

Jaggar, J. H., Stevenson, A. S., & Nelson, M. T. (1998a). Voltage dependence of Ca2+ sparks in intact cerebral arteries. Am J Physiol, 274(6 Pt 1), C1755-1761.

Jaggar, J. H., Wellman, G. C., Heppner, T. J., Porter, V. A., Perez, G. J., Gollasch, M., Kleppisch, T., Rubart, M., Stevenson, A. S., Lederer, W. J., Knot, H. J., Bonev, A. D., & Nelson, M. T. (1998b). Ca2+ channels, ryanodine receptors and Ca(2+)-activated K+ channels: a functional unit for regulating arterial tone. Acta Physiol Scand, 164(4), 577-587.

Jan, L. Y., & Jan, Y. N. (1990). A superfamily of ion channels. Nature, 345(6277), 672. Janiak, R., Wilson, S. M., Montague, S., & Hume, J. R. (2001). Heterogeneity of calcium stores

and elementary release events in canine pulmonary arterial smooth muscle cells. Am J Physiol Cell Physiol, 280(1), C22-33.

Page 172: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

157

Jensen, L. J., & Holstein-Rathlou, N. H. (2009). Is there a role for T-type Ca2+ channels in regulation of vasomotor tone in mesenteric arterioles? Can J Physiol Pharmacol, 87(1), 8-20.

Jensen, L. J., Salomonsson, M., Jensen, B. L., & Holstein-Rathlou, N. H. (2004). Depolarization-induced calcium influx in rat mesenteric small arterioles is mediated exclusively via mibefradil-sensitive calcium channels. Br J Pharmacol, 142(4), 709-718.

Ji, G., Feldman, M., Doran, R., Zipfel, W., & Kotlikoff, M. I. (2006). Ca2+ -induced Ca2+ release through localized Ca2+ uncaging in smooth muscle. J Gen Physiol, 127(3), 225-235.

Jones, P. P., & Braun, A. P. (2009). Store operated Ca2+ entry (SOCE): from structure to function. Channels (Austin), 3(1), 1-2.

Keizer, J., & Levine, L. (1996). Ryanodine receptor adaptation and Ca2+(-)induced Ca2+ release-dependent Ca2+ oscillations. Biophys J, 71(6), 3477-3487.

Kisilevsky, A. E., & Zamponi, G. W. (2008). Presynaptic calcium channels: structure, regulators, and blockers. Handb Exp Pharmacol(184), 45-75.

Knot, H. J., Laher, I., Sobie, E. A., Guatimosim, S., Gomez-Viquez, L., Hartmann, H., Song, L. S., Lederer, W. J., Graier, W. F., Malli, R., Frieden, M., & Petersen, O. H. (2005). Twenty years of calcium imaging: cell physiology to dye for. Mol Interv, 5(2), 112-127.

Knot, H. J., Lounsbury, K. M., Brayden, J. E., & Nelson, M. T. (1999). Gender differences in coronary artery diameter reflect changes in both endothelial Ca2+ and ecNOS activity. Am J Physiol, 276(3 Pt 2), H961-969.

Knot, H. J., & Nelson, M. T. (1995). Regulation of membrane potential and diameter by voltage-dependent K+ channels in rabbit myogenic cerebral arteries. Am J Physiol, 269(1 Pt 2), H348-355.

Knot, H. J., & Nelson, M. T. (1998). Regulation of arterial diameter and wall [Ca2+] in cerebral arteries of rat by membrane potential and intravascular pressure. J Physiol, 508 ( Pt 1), 199-209.

Knot, H. J., Standen, N. B., & Nelson, M. T. (1998). Ryanodine receptors regulate arterial diameter and wall [Ca2+] in cerebral arteries of rat via Ca2+-dependent K+ channels. J Physiol, 508 ( Pt 1), 211-221.

Ko, E. A., Han, J., Jung, I. D., & Park, W. S. (2008). Physiological roles of K+ channels in vascular smooth muscle cells. J Smooth Muscle Res, 44(2), 65-81.

Koh, S. D., Monaghan, K., Ro, S., Mason, H. S., Kenyon, J. L., & Sanders, K. M. (2001). Novel voltage-dependent non-selective cation conductance in murine colonic myocytes. J Physiol, 533(Pt 2), 341-355.

Korovkina, V. P., & England, S. K. (2002). Molecular diversity of vascular potassium channel isoforms. Clin Exp Pharmacol Physiol, 29(4), 317-323.

Kotlikoff, M. I. (2003). Calcium-induced calcium release in smooth muscle: the case for loose coupling. Prog Biophys Mol Biol, 83(3), 171-191.

Kremer, J. R., Mastronarde, D. N., & McIntosh, J. R. (1996). Computer visualization of three-dimensional image data using IMOD. J Struct Biol, 116(1), 71-76.

Kuo, I. Y., Ellis, A., Seymour, V. A., Sandow, S. L., & Hill, C. E. (2010). Dihydropyridine-insensitive calcium currents contribute to function of small cerebral arteries. J Cereb Blood Flow Metab, 30(6), 1226-1239.

Page 173: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

158

Kuo, L., Chilian, W. M., & Davis, M. J. (1991). Interaction of pressure- and flow-induced responses in porcine coronary resistance vessels. Am J Physiol, 261(6 Pt 2), H1706-1715.

Lacinova, L. (2005). Voltage-dependent calcium channels. Gen Physiol Biophys, 24 Suppl 1, 1-78.

Lambert, R. C., Maulet, Y., Mouton, J., Beattie, R., Volsen, S., De Waard, M., & Feltz, A. (1997). T-type Ca2+ current properties are not modified by Ca2+ channel beta subunit depletion in nodosus ganglion neurons. J Neurosci, 17(17), 6621-6628.

Lee, C. H., Kuo, K. H., Dai, J., & van Breemen, C. (2005). Asynchronous calcium waves in smooth muscle cells. Can J Physiol Pharmacol, 83(8-9), 733-741.

Lee, J. H., Gomora, J. C., Cribbs, L. L., & Perez-Reyes, E. (1999). Nickel block of three cloned T-type calcium channels: low concentrations selectively block alpha1H. Biophys J, 77(6), 3034-3042.

Lee, T. J. (1982). Cholinergic mechanism in the large cat cerebral artery. Circ Res, 50(6), 870-879.

Lee, T. S., Kaku, T., Takebayashi, S., Uchino, T., Miyamoto, S., Hadama, T., Perez-Reyes, E., & Ono, K. (2006). Actions of mibefradil, efonidipine and nifedipine block of recombinant T- and L-type Ca channels with distinct inhibitory mechanisms. Pharmacology, 78(1), 11-20.

Leuranguer, V., Bourinet, E., Lory, P., & Nargeot, J. (1998). Antisense depletion of beta-subunits fails to affect T-type calcium channels properties in a neuroblastoma cell line. Neuropharmacology, 37(6), 701-708.

Li, Z., Wang, X., Gao, G., Qu, D., Yu, B., Huang, C., Elmslie, K. S., & Peterson, B. Z. (2010). A single amino acid change in Ca(v)1.2 channels eliminates the permeation and gating differences between Ca(2+) and Ba(2+). J Membr Biol, 233(1-3), 23-33.

Liao, P., Yong, T. F., Liang, M. C., Yue, D. T., & Soong, T. W. (2005). Splicing for alternative structures of Cav1.2 Ca2+ channels in cardiac and smooth muscles. Cardiovasc Res, 68(2), 197-203.

Liao, P., Yu, D., Li, G., Yong, T. F., Soon, J. L., Chua, Y. L., & Soong, T. W. (2007). A smooth muscle Cav1.2 calcium channel splice variant underlies hyperpolarized window current and enhanced state-dependent inhibition by nifedipine. J Biol Chem, 282(48), 35133-35142.

Liao, P., Yu, D., Lu, S., Tang, Z., Liang, M. C., Zeng, S., Lin, W., & Soong, T. W. (2004). Smooth muscle-selective alternatively spliced exon generates functional variation in Cav1.2 calcium channels. J Biol Chem, 279(48), 50329-50335.

Life technologies (2010) User Manual. Retrieved Oct 26, 2013, from http:// www. liftechnologies.com (http://tools.lifetechnologies.com/content/sfs/manuals/mp01240.pdf) Luykenaar, K. D., Brett, S. E., Wu, B. N., Wiehler, W. B., & Welsh, D. G. (2004). Pyrimidine

nucleotides suppress KDR currents and depolarize rat cerebral arteries by activating Rho kinase. Am J Physiol Heart Circ Physiol, 286(3), H1088-1100.

Marcantoni, A., Carabelli, V., Comunanza, V., Hoddah, H., & Carbone, E. (2008). Calcium channels in chromaffin cells: focus on L and T types. Acta Physiol (Oxf), 192(2), 233-246.

Martin, R. L., Lee, J. H., Cribbs, L. L., Perez-Reyes, E., & Hanck, D. A. (2000). Mibefradil block of cloned T-type calcium channels. J Pharmacol Exp Ther, 295(1), 302-308.

Page 174: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

159

Martin-Cano, F. E., Gomez-Pinilla, P. J., Pozo, M. J., & Camello, P. J. (2009). Spontaneous calcium oscillations in urinary bladder smooth muscle cells. J Physiol Pharmacol, 60(4), 93-99.

Martin-Moutot, N., Charvin, N., Leveque, C., Sato, K., Nishiki, T., Kozaki, S., Takahashi, M., & Seagar, M. (1996). Interaction of SNARE complexes with P/Q-type calcium channels in rat cerebellar synaptosomes. J Biol Chem, 271(12), 6567-6570.

Mastronarde, D. N. (1997). Dual-axis tomography: an approach with alignment methods that preserve resolution. J Struct Biol, 120(3), 343-352.

Mastronarde, D. N. (2005). Automated electron microscope tomography using robust prediction of specimen movements. J Struct Biol, 152(1), 36-51.

Mironneau, J., Arnaudeau, S., Macrez-Lepretre, N., & Boittin, F. X. (1996). Ca2+ sparks and Ca2+ waves activate different Ca(2+)-dependent ion channels in single myocytes from rat portal vein. Cell Calcium, 20(2), 153-160.

Mittman, S., Guo, J., Emerick, M. C., & Agnew, W. S. (1999). Structure and alternative splicing of the gene encoding alpha1I, a human brain T calcium channel alpha1 subunit. Neurosci Lett, 269(3), 121-124.

Moosmang, S., Schulla, V., Welling, A., Feil, R., Feil, S., Wegener, J. W., Hofmann, F., & Klugbauer, N. (2003). Dominant role of smooth muscle L-type calcium channel Cav1.2 for blood pressure regulation. EMBO J, 22(22), 6027-6034.

Morel, N., Buryi, V., Feron, O., Gomez, J. P., Christen, M. O., & Godfraind, T. (1998). The action of calcium channel blockers on recombinant L-type calcium channel alpha1-subunits. Br J Pharmacol, 125(5), 1005-1012.

Morita, H., Shi, J., Ito, Y., & Inoue, R. (2002). T-channel-like pharmacological properties of high voltage-activated, nifedipine-insensitive Ca2+ currents in the rat terminal mesenteric artery. Br J Pharmacol, 137(4), 467-476.

Mufti, R. E., Brett, S. E., Tran, C. H., Abd El-Rahman, R., Anfinogenova, Y., El-Yazbi, A., Cole, W. C., Jones, P. P., Chen, S. R., & Welsh, D. G. (2010). Intravascular pressure augments cerebral arterial constriction by inducing voltage-insensitive Ca2+ waves. J Physiol, 588(Pt 20), 3983-4005.

Muik, M., Fahrner, M., Schindl, R., Stathopulos, P., Frischauf, I., Derler, I., Plenk, P., Lackner, B., Groschner, K., Ikura, M., & Romanin, C. (2011). STIM1 couples to ORAI1 via an intramolecular transition into an extended conformation. EMBO J, 30(9), 1678-1689.

Muraki, K., Iwata, Y., Katanosaka, Y., Ito, T., Ohya, S., Shigekawa, M., & Imaizumi, Y. (2003). TRPV2 is a component of osmotically sensitive cation channels in murine aortic myocytes. Circ Res, 93(9), 829-838.

Navarro-Gonzalez, M. F., Grayson, T. H., Meaney, K. R., Cribbs, L. L., & Hill, C. E. (2009). Non-L-type voltage-dependent calcium channels control vascular tone of the rat basilar artery. Clin Exp Pharmacol Physiol, 36(1), 55-66.

Nelson, M. T., Cheng, H., Rubart, M., Santana, L. F., Bonev, A. D., Knot, H. J., & Lederer, W. J. (1995). Relaxation of arterial smooth muscle by calcium sparks. Science, 270(5236), 633-637.

Nelson, M. T., & Quayle, J. M. (1995). Physiological roles and properties of potassium channels in arterial smooth muscle. Am J Physiol, 268(4 Pt 1), C799-822.

Page 175: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

160

Nikitina, E., Zhang, Z. D., Kawashima, A., Jahromi, B. S., Bouryi, V. A., Takahashi, M., Xie, A., & Macdonald, R. L. (2007). Voltage-dependent calcium channels of dog basilar artery. J Physiol, 580(Pt. 2), 523-541.

Nilius, B., Prenen, J., Kamouchi, M., Viana, F., Voets, T., & Droogmans, G. (1997). Inhibition by mibefradil, a novel calcium channel antagonist, of Ca(2+)- and volume-activated Cl- channels in macrovascular endothelial cells. Br J Pharmacol, 121(3), 547-555.

Oguri, A., Tanaka, T., Iida, H., Meguro, K., Takano, H., Oonuma, H., Nishimura, S., Morita, T., Yamasoba, T., Nagai, R., & Nakajima, T. (2010). Involvement of CaV3.1 T-type calcium channels in cell proliferation in mouse preadipocytes. Am J Physiol Cell Physiol, 298(6), C1414-1423.

OLink-Bioscence (2012) Duolink In Situe Fluorescence,User Manual. Retrieved Oct 25, 2013 from http://www.olink.com (http://www.olink.com/sites/default/files/files/0650%20v3.1%20Duolink%20In%20Situ

%20Fluorescence%20User%20Manual_0.pdf). Owsianik, G., Talavera, K., Voets, T., & Nilius, B. (2006). Permeation and selectivity of TRP

channels. Annu Rev Physiol, 68, 685-717. Perchenet, L., & Clement-Chomienne, O. (2000). Characterization of mibefradil block of the

human heart delayed rectifier hKv1.5. J Pharmacol Exp Ther, 295(2), 771-778. Perez, G. J., Bonev, A. D., & Nelson, M. T. (2001). Micromolar Ca(2+) from sparks activates

Ca(2+)-sensitive K(+) channels in rat cerebral artery smooth muscle. Am J Physiol Cell Physiol, 281(6), C1769-1775.

Perez, G. J., Bonev, A. D., Patlak, J. B., & Nelson, M. T. (1999). Functional coupling of ryanodine receptors to KCa channels in smooth muscle cells from rat cerebral arteries. J Gen Physiol, 113(2), 229-238.

Perez-Reyes, E. (2003). Molecular physiology of low-voltage-activated t-type calcium channels. Physiol Rev, 83(1), 117-161.

Picht, E., Zima, A. V., Blatter, L. A., & Bers, D. M. (2007). SparkMaster: automated calcium spark analysis with ImageJ. Am J Physiol Cell Physiol, 293(3), C1073-1081.

Pluteanu, F., & Cribbs, L. L. (2011). Regulation and function of Cav3.1 T-type calcium channels in IGF-I-stimulated pulmonary artery smooth muscle cells. Am J Physiol Cell Physiol, 300(3), C517-525.

Poulsen, C. B., Al-Mashhadi, R. H., Cribbs, L. L., Skott, O., & Hansen, P. B. (2011). T-type voltage-gated calcium channels regulate the tone of mouse efferent arterioles. Kidney Int, 79(4), 443-451.

Pries, A. R., Secomb, T. W., Gaehtgens, P., & Gross, J. F. (1990). Blood flow in microvascular networks. Experiments and simulation. Circ Res, 67(4), 826-834.

Pries, A. R., Secomb, T. W., Gessner, T., Sperandio, M. B., Gross, J. F., & Gaehtgens, P. (1994). Resistance to blood flow in microvessels in vivo. Circ Res, 75(5), 904-915.

Ramsey, I. S., Delling, M., & Clapham, D. E. (2006). An introduction to TRP channels. Annu Rev Physiol, 68, 619-647.

Reynolds, E. S. (1963). The use of lead citrate at high pH as an electron-opaque stain in electron microscopy. J Cell Biol, 17, 208-212.

Rodman, D. M., Reese, K., Harral, J., Fouty, B., Wu, S., West, J., Hoedt-Miller, M., Tada, Y., Li, K. X., Cool, C., Fagan, K., & Cribbs, L. (2005). Low-voltage-activated (T-type)

Page 176: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

161

calcium channels control proliferation of human pulmonary artery myocytes. Circ Res, 96(8), 864-872.

Salemme, S., Rebolledo, A., Speroni, F., Petruccelli, S., & Milesi, V. (2007). L, P-/Q- and T-type Ca2+ channels in smooth muscle cells from human umbilical artery. Cell Physiol Biochem, 20(1-4), 55-64.

Sanders, K. M. (2001). Invited review: mechanisms of calcium handling in smooth muscles. J Appl Physiol, 91(3), 1438-1449.

Segal, S. S. (2000). Integration of blood flow control to skeletal muscle: key role of feed arteries. Acta Physiol Scand, 168(4), 511-518.

Segal, S. S., & Jacobs, T. L. (2001). Role for endothelial cell conduction in ascending vasodilatation and exercise hyperaemia in hamster skeletal muscle. J Physiol, 536(Pt 3), 937-946.

Shcheglovitov, A., Vitko, I., Bidaud, I., Baumgart, J. P., Navarro-Gonzalez, M. F., Grayson, T. H., Lory, P., Hill, C. E., & Perez-Reyes, E. (2008). Alternative splicing within the I-II loop controls surface expression of T-type Ca(v)3.1 calcium channels. FEBS Lett, 582(27), 3765-3770.

Si, M. L., & Lee, T. J. (2002). Alpha7-nicotinic acetylcholine receptors on cerebral perivascular sympathetic nerves mediate choline-induced nitrergic neurogenic vasodilation. Circ Res, 91(1), 62-69.

Smirnov, S. V., & Aaronson, P. I. (1992). Ca2+ currents in single myocytes from human mesenteric arteries: evidence for a physiological role of L-type channels. J Physiol, 457, 455-475.

Stamboulian, S., Kim, D., Shin, H. S., Ronjat, M., De Waard, M., & Arnoult, C. (2004). Biophysical and pharmacological characterization of spermatogenic T-type calcium current in mice lacking the CaV3.1 (alpha1G) calcium channel: CaV3.2 (alpha1H) is the main functional calcium channel in wild-type spermatogenic cells. J Cell Physiol, 200(1), 116-124.

Sutton, K. G., McRory, J. E., Guthrie, H., Murphy, T. H., & Snutch, T. P. (1999). P/Q-type calcium channels mediate the activity-dependent feedback of syntaxin-1A. Nature, 401(6755), 800-804.

Takeshima, H., Nishimura, S., Matsumoto, T., Ishida, H., Kangawa, K., Minamino, N., Matsuo, H., Ueda, M., Hanaoka, M., Hirose, T., & et al. (1989). Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature, 339(6224), 439-445.

Talavera, K., Staes, M., Janssens, A., Klugbauer, N., Droogmans, G., Hofmann, F., & Nilius, B. (2001). Aspartate residues of the Glu-Glu-Asp-Asp (EEDD) pore locus control selectivity and permeation of the T-type Ca(2+) channel alpha(1G). J Biol Chem, 276(49), 45628-45635.

Tanaka, Y., Koike, K., & Toro, L. (2004). MaxiK channel roles in blood vessel relaxations induced by endothelium-derived relaxing factors and their molecular mechanisms. J Smooth Muscle Res, 40(4-5), 125-153.

Tang, Z. Z., Liang, M. C., Lu, S., Yu, D., Yu, C. Y., Yue, D. T., & Soong, T. W. (2004). Transcript scanning reveals novel and extensive splice variations in human l-type

Page 177: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

162

voltage-gated calcium channel, Cav1.2 alpha1 subunit. J Biol Chem, 279(43), 44335-44343.

Teramoto, N. (2006). Physiological roles of ATP-sensitive K+ channels in smooth muscle. J Physiol, 572(Pt 3), 617-624.

Thorneloe, K. S., & Nelson, M. T. (2005). Ion channels in smooth muscle: regulators of intracellular calcium and contractility. Can J Physiol Pharmacol, 83(3), 215-242.

Townsley, M. I., King, J. A., & Alvarez, D. F. (2006). Ca2+ channels and pulmonary endothelial permeability: insights from study of intact lung and chronic pulmonary hypertension. Microcirculation, 13(8), 725-739.

van der Lee, R., Pfaffendorf, M., & van Zwieten, P. A. (1999). Comparative effects of mibefradil and other calcium antagonists on resistance arteries of different end organs. Fundam Clin Pharmacol, 13(2), 198-203.

Walsh, M. P., Kargacin, G. J., Kendrick-Jones, J., & Lincoln, T. M. (1995). Intracellular mechanisms involved in the regulation of vascular smooth muscle tone. Can J Physiol Pharmacol, 73(5), 565-573.

Wellman, G. C., Nathan, D. J., Saundry, C. M., Perez, G., Bonev, A. D., Penar, P. L., Tranmer, B. I., & Nelson, M. T. (2002). Ca2+ sparks and their function in human cerebral arteries. Stroke, 33(3), 802-808.

Wellman, G. C., & Nelson, M. T. (2003). Signaling between SR and plasmalemma in smooth muscle: sparks and the activation of Ca2+-sensitive ion channels. Cell Calcium, 34(3), 211-229.

Welsh, D. G., Morielli, A. D., Nelson, M. T., & Brayden, J. E. (2002). Transient receptor potential channels regulate myogenic tone of resistance arteries. Circ Res, 90(3), 248-250.

Welsh, D. G., Nelson, M. T., Eckman, D. M., & Brayden, J. E. (2000). Swelling-activated cation channels mediate depolarization of rat cerebrovascular smooth muscle by hyposmolarity and intravascular pressure. J Physiol, 527 Pt 1, 139-148.

Wray, S., & Burdyga, T. (2010). Sarcoplasmic reticulum function in smooth muscle. Physiol Rev, 90(1), 113-178.

Wray, S., Burdyga, T., & Noble, K. (2005). Calcium signalling in smooth muscle. Cell Calcium, 38(3-4), 397-407.

Wu, B. N., Luykenaar, K. D., Brayden, J. E., Giles, W. R., Corteling, R. L., Wiehler, W. B., & Welsh, D. G. (2007). Hyposmotic challenge inhibits inward rectifying K+ channels in cerebral arterial smooth muscle cells. Am J Physiol Heart Circ Physiol, 292(2), H1085-1094.

Yang, X. R., Lin, M. J., McIntosh, L. S., & Sham, J. S. (2006). Functional expression of transient receptor potential melastatin- and vanilloid-related channels in pulmonary arterial and aortic smooth muscle. Am J Physiol Lung Cell Mol Physiol, 290(6), L1267-1276.

Yin, J., & Kuebler, W. M. (2010). Mechanotransduction by TRP channels: general concepts and specific role in the vasculature. Cell Biochem Biophys, 56(1), 1-18.

Zhang, D. X., Gauthier, K. M., & Campbell, W. B. (2005). Mechanisms of histamine-induced relaxation in bovine small adrenal cortical arteries. Am J Physiol Endocrinol Metab, 289(6), E1058-1063.

Page 178: The Role of L- and T-type Ca2+ Channels in Rat Cerebral ...

163

Zhao, M., Li, P., Li, X., Zhang, L., Winkfein, R. J., & Chen, S. R. (1999). Molecular identification of the ryanodine receptor pore-forming segment. J Biol Chem, 274(37), 25971-25974.

Zhong, X., Liu, J. R., Kyle, J. W., Hanck, D. A., & Agnew, W. S. (2006). A profile of alternative RNA splicing and transcript variation of CACNA1H, a human T-channel gene candidate for idiopathic generalized epilepsies. Hum Mol Genet, 15(9), 1497-1512.

Zhou, C., Chen, H., King, J. A., Sellak, H., Kuebler, W. M., Yin, J., Townsley, M. I., Shin, H. S., & Wu, S. (2010). Alpha1G T-type calcium channel selectively regulates P-selectin surface expression in pulmonary capillary endothelium. Am J Physiol Lung Cell Mol Physiol, 299(1), L86-97.

Zhou, C., & Wu, S. (2006). T-type calcium channels in pulmonary vascular endothelium. Microcirculation, 13(8), 645-656.

Zhuang, H., Bhattacharjee, A., Hu, F., Zhang, M., Goswami, T., Wang, L., Wu, S., Berggren, P. O., & Li, M. (2000). Cloning of a T-type Ca2+ channel isoform in insulin-secreting cells. Diabetes, 49(1), 59-64.


Recommended