165
Wright State University Wright State University CORE Scholar CORE Scholar Browse all Theses and Dissertations Theses and Dissertations 2007 Functional Interplay Between Subthreshold Ion Channels in Functional Interplay Between Subthreshold Ion Channels in Preautonomic Neurons of the Hypothalamic Paraventricular Preautonomic Neurons of the Hypothalamic Paraventricular Nucleus in Health and Disease Conditions Nucleus in Health and Disease Conditions Patrick M. Sonner Wright State University Follow this and additional works at: https://corescholar.libraries.wright.edu/etd_all Part of the Biomedical Engineering and Bioengineering Commons Repository Citation Repository Citation Sonner, Patrick M., "Functional Interplay Between Subthreshold Ion Channels in Preautonomic Neurons of the Hypothalamic Paraventricular Nucleus in Health and Disease Conditions" (2007). Browse all Theses and Dissertations. 208. https://corescholar.libraries.wright.edu/etd_all/208 This Dissertation is brought to you for free and open access by the Theses and Dissertations at CORE Scholar. It has been accepted for inclusion in Browse all Theses and Dissertations by an authorized administrator of CORE Scholar. For more information, please contact [email protected].

Functional Interplay Between Subthreshold Ion Channels in

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Functional Interplay Between Subthreshold Ion Channels in

Wright State University Wright State University

CORE Scholar CORE Scholar

Browse all Theses and Dissertations Theses and Dissertations

2007

Functional Interplay Between Subthreshold Ion Channels in Functional Interplay Between Subthreshold Ion Channels in

Preautonomic Neurons of the Hypothalamic Paraventricular Preautonomic Neurons of the Hypothalamic Paraventricular

Nucleus in Health and Disease Conditions Nucleus in Health and Disease Conditions

Patrick M. Sonner Wright State University

Follow this and additional works at: https://corescholar.libraries.wright.edu/etd_all

Part of the Biomedical Engineering and Bioengineering Commons

Repository Citation Repository Citation Sonner, Patrick M., "Functional Interplay Between Subthreshold Ion Channels in Preautonomic Neurons of the Hypothalamic Paraventricular Nucleus in Health and Disease Conditions" (2007). Browse all Theses and Dissertations. 208. https://corescholar.libraries.wright.edu/etd_all/208

This Dissertation is brought to you for free and open access by the Theses and Dissertations at CORE Scholar. It has been accepted for inclusion in Browse all Theses and Dissertations by an authorized administrator of CORE Scholar. For more information, please contact [email protected].

Page 2: Functional Interplay Between Subthreshold Ion Channels in

FUNCTIONAL INTERPLAY BETWEEN SUBTHRESHOLD ION CHANNELS

IN PREAUTONOMIC NEURONS OF THE HYPOTHALAMIC

PARAVENTRICULAR NUCLEUS IN HEALTH AND DISEASE CONDITIONS

A dissertation submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

By

Patrick M. Sonner B.S., Ohio Northern University, 2000

2007 Wright State University

Page 3: Functional Interplay Between Subthreshold Ion Channels in

WRIGHT STATE UNIVERSITY SCHOOL OF GRADUATE STUDIES

12/10/07

I HEREBY RECOMMEND THAT THE DISSERTATION PREPARED UNDER MY SUPERVISION BY Patrick Michael Sonner ENTITLED Functional interplay between subthreshold ion channels in preautonomic neurons of the hypothalamic paraventricular nucleus in health and disease conditions BE ACCEPTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF Doctor of Philosophy. Javier E. Stern, M.D., Ph.D. Dissertation Director Gerald M. Alter, Ph.D. Director, Biomedical Sciences Ph.D. Program Joseph F. Thomas, Jr., Ph.D. Dean, School of Graduate Studies Committee on Final Examination Javier E. Stern, M.D., Ph.D. Robert Putnam, Ph.D. David Goldstein, Ph.D. James Olson, Ph.D. David Cool, Ph.D.

Page 4: Functional Interplay Between Subthreshold Ion Channels in

ABSTRACT

Sonner, Patrick M., Ph.D. Biomedical Sciences Ph.D. Program, Wright State University, 2007. Functional Interplay between Subthreshold Ion Channels in Preautonomic Neurons of the Hypothalamic Paraventricular Nucleus in Health and Disease Conditions.

Under normal conditions, blood pressure is tightly regulated through autonomic

tonic and reflex mechanisms. However, when the set-point for blood pressure is

chronically elevated, hypertension occurs. Hypertension if untreated can lead to further

complications including heart failure, stroke and kidney failure. Elevated sympathetic

outflow is known to contribute to the development and/or maintenance of hypertension,

and while the hypothalamic paraventricular nucleus (PVN), a preautonomic center, has

been implicated in the elevation of sympathetic activity during hypertension, the precise

pathophysiological mechanisms underlying sympathoexcitation remain unclear.

Subthreshold ion channels, including the A-type K+ (IA) and the T-type Ca2+ (IT), are

important mechanisms regulating the ability of neurons to generate firing activity, and

changes in IA activity have been reported during hypertension. Thus, the first aim of the

study focused on characterizing the basic biophysical and functional properties of IA in

presympathetic PVN neurons projecting to the rostral ventrolateral medulla (PVN-

RVLM). Our studies demonstrated the presence of a functionally relevant IA in PVN-

RVLM neurons, which actively modulated the action potential waveform and firing

activity. The second aim of the study was to determine whether alterations in the

iii

Page 5: Functional Interplay Between Subthreshold Ion Channels in

biophysical properties of IA contributed to enhanced neuronal excitability of PVN-RVLM

neurons during hypertension. Our studies indicated that diminished IA availability

constituted a contributing mechanism underlying hyperexcitability in these neurons

during hypertension. Previous studies have indicated an opposing balance between the

subthreshold ion channels, IA and IT. Thus, the final aim of the study assessed the

biophysical competition between IA and IT, and functionally addressed the influence of

such balance on the activity of PVN-RVLM neurons under normal and hypertensive

conditions. Our studies indicated that the balance between IA and IT was shifted during

hypertension in favor of IT activity, resulting in increased IT –dependent low threshold

spikes, elevated intracellular calcium levels, and enhanced basal spontaneous firing

activity during this condition. Taken together, this study confirms the importance of

intrinsic factors, in particular the balance between opposing subthreshold conductances,

in regulating the central control of cardiovascular output under normal and pathological

conditions.

iv

Page 6: Functional Interplay Between Subthreshold Ion Channels in

TABLE OF CONTENTS

I. INTRODUCTION…………………………………………………………………… 1

1.1 Introduction………………………………………………………………………….. 1

1.2 Neurogenic control of blood pressure………………………………………….......... 1

1.3 Preautonomic cardiovascular control centers………………………………...……... 2

1.4 PVN descending pathways involved in cardiovascular control…………………….. 4

1.5 Cellular organization of the PVN…………………………………………………… 5

1.6 Role of PVN preautonomic neurons (PVN-PA) in hypertension…………………… 6

1.7 Potential mechanisms contributing to altered neuronal function during hypertension…………………………………………………………………. 7

1.8 Major ionic mechanisms controlling neuronal activity of PVN neurons…….……... 8

1.9 Thesis focus and significance……………………………………………………... 10

II. MATERIALS AND METHODS………………………………………………… 12

III. RESULTS……………………………………………………………...…………. 30

3.1 Functional Role of A-type Potassium Currents in Rat Presympathetic PVN Neurons……………………………………………………………………………. 30

3.1.1 Intrinsic properties of PVN-RVLM neurons………………………………….. 30

3.1.2 Activation properties of IA in PVN-RVLM neurons………………………….. 33

3.1.3 Inactivation properties of IA in PVN-RVLM neurons………………………… 33

3.1.4 Recovery of IA from inactivation……………………………………………… 39

3.1.5 4-AP inhibits IA in PVN-RVLM neurons……………………………………... 39

v

Page 7: Functional Interplay Between Subthreshold Ion Channels in

3.1.6 IA shapes the Na+ action potential waveform in PVN-RVLM neurons………. 44

3.1.7 IA differentially regulates repetitive firing activity of PVN-RVLM neurons…. 47

3.1.8 IA restrains firing activity in a subset of PVN-RVLM neurons………………. 48

3.1.9 IA contributes to ongoing firing activity in a subset of PVN-RVLM neurons… 52

3.1.10 Effects of TEA on action potential waveform and firing activity in PVN-RVLM neurons……..………………………………………………….. 56

3.1.11 The effects of 4-AP on action potential waveform and firing activity in PVN-RVLM neurons are Ca2+ dependent………….……………..…………. 59

3.1.12 Possible potassium channel subunits underlying A-type potassium currents in PVN-RVLM neurons……………………………..…..…………………… 62

3.2 Altered A-type Potassium Current and Firing Properties in Rat Sympathetic Preautonomic PVN Neurons During Hypertension…………….……..................... 65

3.2.1 Intrinsic properties of sham and hypertensive PVN-RVLM neurons…………. 65

3.2.2 Changes in PVN-RVLM neuronal morphometry during hypertension………. 65

3.2.3 IA voltage-dependent activation properties……………………………………. 66

3.2.4 IA voltage-dependent inactivation properties………………………………….. 74

3.2.5 Recovery of IA from inactivation……………………………………………… 74

3.2.6 Non-stationary fluctuation analysis of IA in PVN-RVLM neurons: Changes during hypertension………………………………………………….. 79

3.2.7 The Na+ action potential waveform is affected in PVN-RVLM neurons during hypertension…………………………………………………... 82

3.2.8 Action potential-dependent increase in intracellular Ca2+ is enhanced in PVN-RVLM neurons during hypertension…………………………................. 85

3.2.9 Spike broadening during repetitive firing is enhanced in PVN-RVLM neurons during hypertension………………………………………….………………... 86

3.3 Contribution of Subthreshold Ion Channels, to Hypertension, in Preautonomic PVN Neurons……………………………………………………………………………. 91

vi

Page 8: Functional Interplay Between Subthreshold Ion Channels in

3.3.1 Intrinsic properties of sham and hypertensive PVN-RVLM neurons………… 91

3.3.2 IA and IT are expressed and compete within individual PVN-RVLM neurons... 91

3.3.3 IA/IT balance influences the expression and magnitude of LTS………………. 95

3.3.4 The balance of IA/IT is neurochemical-dependent…………………………... 100

3.3.5 LTS-dependent increase in intracellular calcium in PVN-RVLM neurons…. 100

3.3.6 Changes during hypertension………………………………………………... 105

IV. DISCUSSION………………………………………………………………….. 114

4.1 A-type Potassium Currents in PVN-RVLM Neurons…………………………… 115

4.1.1 Biophysical, pharmacological and molecular properties of IA in PVN-RVLM neurons………..………………………………………………. 115

4.1.2 IA influences excitability and repetitive firing properties of PVN-RVLM neurons………………………………………………………... 118

4.1.3 The effects of IA on PVN-RVLM firing activity are Ca2+-dependent………. 121

4.2 Balance between IA and IT in PVN-RVLM Neurons……………….…………… 121

4.2.1 Interplay between the subthreshold currents, IA and IT, in PVN-RVLM neurons………………………………………………………... 122

4.2.2 IA/IT balance influences neuronal activity…………..……………………….. 124

4.3 Altered Contribution of Subthreshold Ion Channels to Increased PVN-RVLM Neuronal Excitability During Hypertension…………………………………….. 125 4.3.1 Altered balance between IA and IT in PVN-RVLM neurons, during hypertension….....…………………………………………………… 126 4.3.2 Altered IA/IT balance during hypertension influences PVN-RVLM neuronal activity…………...………………………………………………… 127

4.4 Source of Disbalanced Interplay between IA and IT During Hypertension……… 129

4.5 Future Studies………………………………………………………………….... 130

4.6 General Conclusions…………………………………………………………..… 132

vii

Page 9: Functional Interplay Between Subthreshold Ion Channels in

Appendix A: Abbreviations…………………………………………………………. 133

Appendix B: Methodological Considerations……………….………………………. 135

V. REFERENCES………………………………………………………………….. 138

viii

Page 10: Functional Interplay Between Subthreshold Ion Channels in

ILLUSTRATIONS

Figure Page

1. PVN-RVLM projecting neurons are identified following microinjections of a fluorescent retrograde tracer in the RVLM……………………………..……...32

2. Isolation and voltage-dependent activation of IA in PVN-RVLM neurons……...…35

3. Voltage-dependent and kinetics of inactivation of IA……………………………....38

4. Time course of recovery from inactivation of IA……………………………….…..41

5. Inhibition of IA by 4-AP is both concentration and voltage-dependent………….…43

6. Effects of 4-AP on evoked Na+ action potential waveforms………………………. 46

7. 4-AP increases firing discharge in a subset of PVN-RVLM neurons……………....51

8. 4-AP diminished firing discharge in a subset of PVN-RVLM neurons…………….55

9. Effects of TEA on outward K+ currents and action potential waveform in PVN-RVLM neurons………………………………………………………………..58

10. 4-AP effects on PVN-RVLM firing discharge are Ca2+-dependent………………...61

11. Kv1.4, 4.2 and 4.3 immunoreactivity in retrogradely labeled PVN-RVLM neurons…………………………...………………………………………………….64

12. Changes in PVN-RVLM neuronal morphometry during hypertension……………..68

13. Changes in IA voltage-dependent activation properties during hypertension…….....70

14. Morphometric and state-dependent differences in IA voltage-dependent activation properties of PVN-DVC neurons……………………………..………….73

15. Changes in IA voltage-dependent inactivation properties in PVN-RVLM neurons during hypertension………………………………………………………..76

ix

Page 11: Functional Interplay Between Subthreshold Ion Channels in

16. Time course of recovery from inactivation of IA in PVN-RVLM neurons in sham and hypertensive rats……………………………….…………………………78

17. Non-stationary noise analysis of IA in PVN-RVLM neurons of sham and hypertensive rats…………………………………………………………….……....81

18. Changes in Na+ action potential waveform of PVN-RVLM neurons during hypertension…………………………………………………….…………………..84

19. Changes in action potential evoked [Ca2+]ic levels in PVN-RVLM neurons during hypertension…………………………………………………………………88

20. Changes in spike broadening in PVN-RVLM neurons during hypertension……….90

21. IA and IT are expressed within individual PVN-RVLM neurons……………..……..93

22. The relative expression of mRNA for A-type and T-type channel subunits correlates with the balance of these two subthreshold conductances…………………..……...97

23. The balance of IA and IT influences the expression and magnitude of LTS………...99

24. The neurochemical identity of PVN-RVLM neurons influences the balance of IA and IT……………………………………………………………..….102

25. LTS-evoked changes in intracellular calcium levels in PVN-RVLM neurons…….104

26. LTS-evoked changes in intracellular calcium in PVN-RVLM dendrites …………107

27. Shift in the balance of IA and IT during hypertension…………………………..….110

28. IT contributes to enhanced spontaneous firing activity in PVN-RVLM neurons during hypertension ………….………………………………………………...….113

x

Page 12: Functional Interplay Between Subthreshold Ion Channels in

TABLES

Table Page

1. Real Time PCR Primers……………………………………………………………25

2. Action Potential Waveform Parameters……………………………………………49

xi

Page 13: Functional Interplay Between Subthreshold Ion Channels in

ACKNOWLEDGMENTS

I would like to thank my advisor, Dr. Javier Stern. It has been such a rewarding

experience working with you, and I can not thank you enough for all of the experience

and guidance I have received over the years. I believe that you have truly helped me to

grow as a scientist. I would also like to thank all of the members of my committee for

their input and advice: Dr. Robert Putnam, Dr. David Goldstein, Dr. James Olson, and

Dr. David Cool.

I am grateful to Drs. Jessica Filosa and Victor Blanco for their assistance in

trouble-shooting, analyzing, and all things related to the calcium-imaging experiments. I

would also like to thank Dr. Vinicia Biancardi and Mrs. Keisha Kinney for teaching me

how to perform the renovascular hypertension and retrograde-injection surgeries. As

well, I thank Sora Lee for her assistance with the single-cell RT-PCR experiments.

Finally, I would like to thank the Biomedical Sciences Ph.D. program for funding

and support.

xii

Page 14: Functional Interplay Between Subthreshold Ion Channels in

DEDICATION

To Martha, your love, encouragement, and patience knows no bounds.

To my family, for support and laughter.

xiii

Page 15: Functional Interplay Between Subthreshold Ion Channels in

I. INTRODUCTION

1.1. Introduction

According to the American Heart Association, one in five Americans has

hypertension, however, a third are unaware of their condition. Hypertension if untreated

can lead to further complications including heart failure, stroke, and kidney failure.

While elevated sympathetic outflow is known to contribute to the development and/or

maintenance of hypertension, and is commonly observed in various forms of

hypertensive disorders, both in humans and in experimental animal models [59,81,95],

the precise pathophysiological mechanisms underlying sympathoexcitation in

hypertensive individuals remain largely unknown. However, accumulating evidence

supports altered central nervous system (CNS) control of autonomic function during

hypertension [81], and implicates the hypothalamic paraventricular nucleus (PVN) as a

major contributing neuronal substrate.

1.2. Neurogenic control of blood pressure

Neural regulation of blood pressure is controlled in both short and long-term time

scales. The main short-term controller of blood pressure is the arterial baroreceptor

reflex, in which peripheral baroreceptors sense changes in blood pressure and send

afferent projections to central regions, resulting in rapid changes in cardiac output and

1

Page 16: Functional Interplay Between Subthreshold Ion Channels in

vascular resistance. These responses can be associated with substantial changes in blood

pressure, and are physiologically adaptive, based upon behavioral and environmental

demands [81]. It has been demonstrated, however, that surgical denervation of centrally

projecting baroreceptors has no long-term effect on blood pressure [43,151], suggesting

that a central role in the long-term control of blood pressure remains intact. While the

sensor of neurally mediated long-term control of blood pressure remains unclear, it is

believed to involve compounds associated with body fluid homeostasis and

osmoreceptors, such as angiotensin II (AngII) and sodium [152]. Circumventricular

organs, central regions lacking the blood-brain barrier, including the subfornical organ

(SFO), are well known to be osmosensitive, as well as able to detect changes in

circulating levels of hormones involved in fluid volume regulation [140]. It has been

demonstrated that the SFO sends projections to a sympathetic preautonomic center

located within the hypothalamus, the PVN [9,60]. Thus, these regions have been

suggested to play an important role in the long-term control of blood pressure [47].

However, it has also been hypothesized that a central baroreceptor may exist within the

rostral ventrolateral medulla (RVLM) [152].

1.3. Preautonomic cardiovascular control centers

The central autonomic nervous system provides control over peripheral functions

necessary to maintain body homeostasis. This balance is achieved through the combined

and balanced activity of the sympathetic and parasympathetic preganglionic neurons

located in the intermediolateral column of the spinal cord (IML) as well as the nucleus

ambiguus (NA) and dorsal motor nucleus of the vagus (DMX), respectively. Relatively

2

Page 17: Functional Interplay Between Subthreshold Ion Channels in

basic aspects of tonic and reflex autonomic functions, including the baroreflex, are

mediated by spinal and bulbar regions. For example, when a drop in blood pressure

occurs, the change is monitored peripherally by baroreceptors located in the aortic arch

and carotid sinus providing afferent inputs to the nucleus of the solitary tract (NTS) via

vagal and glossopharyngeal nerves, respectively [182]. The NTS, which is a major

cardiovascular integrative center, then sends glutamatergic projections to the caudal

ventrolateral medulla (CVLM), which inhibits the IML-projecting rostral ventrolateral

medulla (RVLM) neurons through GABAergic projections, thus diminishing sympathetic

output [49,81]. The NTS also directly innervates the NA and the DMX, through

excitatory inputs [183] to regulate parasympathetic output. Thus, these preganglionic

centers are capable of correcting changes in blood pressure through combined

sympathetic and parasympathetic output. In addition to directly influencing sympathetic

and parasympathetic outflows, these spinal/bulbar centers also send ascending projections

to hierarchically higher centers including the major preautonomic centers. These are

located in the hypothalamus and other forebrain regions, and participate in the generation

of more complex and integrative autonomic responses, involving multiple territories and

systems. It is the preautonomic centers, through direct projections to sympathetic and

parasympathetic preganglionic neurons, which provide this complex integrative and

behavioral regulation of tonic and reflex autonomic function [3,52,79,162,171,208].

There are numerous preautonomic centers, including the suprachiasmatic nucleus (SCN),

ventrolateral medulla (VLM), A5, caudal raphe nucleus, NTS, RVLM and PVN

[17,24,69,90,165], which connect bidirectionally with lower central regions in order to

integrate a variety of behavioral and environmental signals. The NTS, as mentioned

3

Page 18: Functional Interplay Between Subthreshold Ion Channels in

above, acts as the major cardiovascular integrative center providing efferent projections

to a variety of autonomic centers. However, while the NTS acts as the major afferent

integrative center of the mechano- and chemoreceptors, it is the RVLM that primarily

regulates the barosensitive sympathetic efferents [47]. Interestingly, the PVN, located

within the forebrain, is a higher center which regulates both the RVLM and NTS

[4,55,195], and thus allows for integrative and behavioral control over the cardiovascular

system

1.4. PVN descending pathways involved in cardiovascular control

As mentioned above, the PVN is an important homeostatic integrative center, able

to regulate a variety of cardiovascular functions, such as blood volume [16,41,84,169],

osmolality [32,61,62,196], and blood pressure [61,99,157,211,212], via a combination of

neuroendocrine and preautonomic efferent projections [153,193]. The PVN is itself a

preautonomic center with direct projections to preganglionic neurons in the

intermediolateral column of the spinal cord [90,161,193], allowing for direct sympathetic

control of cardiovascular function [131]. The PVN also mediates preautonomic output

indirectly via the RVLM [34,48,83], a cardiovascular control center able to mediate tonic

and reflex sympathetic outflow [49,104,108,162,170,171,188] via direct projections to

preganglionic neurons in the IML [8,204]. RVLM-projecting PVN neurons (PVN-

RVLM) are capable of eliciting sympathoexcitatory and pressor responses upon

activation [4,42,110,195,205]. Interestingly, it has been shown that up to 30% of spinally

projecting PVN neurons send collaterals to the RVLM, likely reflecting a more complex

regulation of sympathetic nerve activity by the PVN [158,174].

4

Page 19: Functional Interplay Between Subthreshold Ion Channels in

The PVN also influences cardiovascular function through the regulation of

parasympathetic output. A major target innervated by the PVN is the dorsal vagal

complex (DVC) [55,72,98,178]. The DVC is an autonomic center comprising

parasympathetic neurons known to regulate tonic and reflex cardiac activity

[86,144,150,200]. Specifically, the DVC is comprised of multiple sub-nuclei that send

direct projections to vagal preganglionic neurons, including the nucleus ambiguus and the

dorsal motor nucleus [183], allowing for vagal motor control of the heart. The DVC also

houses the NTS, a major visceroreceptive autonomic integrating center and key

modulator of the baroreflex [69,111,138,154]. NTS-projecting PVN neurons (PVN-

NTS), upon stimulation, have been shown to suppress the baroreflex through inhibition of

the NTS [33,55,56,145], inducing an elevation in heart rate and cardiac output during the

defense reaction or exercise. This phenomenon is also commonly observed during

chronic hypertension [14,66,78].

Thus, through modulatory actions on the baroreceptor reflex, parasympathetic

and sympathetic nerve activity, the PVN plays a major role in the short- and long-term

control of the cardiovascular system [47].

1.5. Cellular organization of the PVN

The different modalities and pathways by which the PVN influences

cardiovascular control are represented in discrete neuronal groups, including

neuroendocrine and autonomic-related neurons. In general, neurons contained within the

PVN are classified as magnocellular or parvocellular. The magnocellular neurons are

contained within three distinct anatomical regions, and send neuroendocrine efferent

5

Page 20: Functional Interplay Between Subthreshold Ion Channels in

projections to the posterior pituitary. On the other hand, the parvocellular neurons are

distributed within five distinct anatomical regions. One group of parvocellular neurons

elicits neuroendocrine responses through projections to the median eminence resulting in

the release of neurohormones that control anterior pituitary function. A second major

group of parvocellular neurons mediate preautonomic functions via efferent projections

to various autonomic-related centers in the brain stem and spinal cord [193], including

the RVLM, NTS, intermediolateral column (IML) of the spinal cord, dorsal motor

nucleus of the vagus and the locus coeruleus.

1.6. Role of PVN preautonomic neurons (PVN-PA) in hypertension

Accumulating evidence supports an important role for the PVN in the

pathophysiology of hypertension. Elevated neuronal activation marker expression (e.g.,

hexokinase activity, FOS and Fra-like immunoreactivity) suggests increased neuronal

activity in the PVN in a variety of experimental models of hypertension, including high

Na+, renovascular, spontaneously hypertensive (SHR), and aortic depressor nerve

transection [23,45,50,96,114,198]. This corresponds with an increased excitability of

presympathetic PVN neurons, as recently shown [120]. In addition, lesions of the PVN

diminish the mean arterial pressure and reduce sympathetic nerve activity in renal wrap,

SHR, and high Na+ hypertensive rats [35,76,85,149]. More importantly, recent work

indicates that the PVN directly contributes to the elevated blood pressure and sympathetic

activity associated with spontaneous and renal wrap hypertension [3,4,85]. Furthermore,

both the RVLM and NTS have been shown to be altered during hypertension resulting in

sympathoexcitation [15,25,29,57,101,141,168]. Thus, altogether these data support that

6

Page 21: Functional Interplay Between Subthreshold Ion Channels in

enhanced PVN activity, and consequently altered RVLM and NTS function, contribute to

enhanced sympathetic output and suppression of the baroreflex, respectively, during

hypertension. Recent work in RVLM-projecting [4,120] and NTS-projecting PVN

neurons [137] supports this idea, showing enhanced firing activity of PVN-RVLM

neurons is mediated in part by a decrease and increase in GABAA and GABAB receptor

function, respectively. As well, PVN-NTS neurons show reduced oxytocin mRNA levels

and neurons in the NTS express less oxytocin receptor mRNA levels. Altogether, these

data suggest that hyperactivation of PVN neurons contributes to maintenance of elevated

blood pressure and sympathetic outflow during hypertension. Despite the pivotal role of

the PVN during hypertension, the precise cellular mechanisms contributing to elevated

PVN neuronal activity during hypertension remain largely unknown. This information is

critical, because detailed knowledge on pathophysiological mechanisms underlying

sympathoexcition is needed for the development of novel and/or more efficient

therapeutic approaches for the treatment of this prevalent disorder.

1.7. Potential mechanisms contributing to altered neuronal function during

hypertension

Similar to other CNS neuronal populations, preautonomic PVN neuronal

excitability and firing activity is fine-tuned by the combined actions of both intrinsic

properties and synaptic inputs [26,117,118,120-122,124,180,185]. In this sense, most of

the work in the field, so far, has focused on extrinsic, synaptic inputs. For example,

alterations in various PVN neurotransmitter systems have been reported during

hypertension, including diminished GABAergic activity [46,120], reduced nNOS levels

7

Page 22: Functional Interplay Between Subthreshold Ion Channels in

[54], and elevated glutamatergic [121] and angiotensin II activity [80,96]. Thus, these

studies indicate that a change in the balance of inhibitory and excitatory inputs, favoring

the latter, contributes to increased PVN neuronal activity and autonomic outflow during

hypertension. It is important to consider however, that some of these extrinsic

mechanisms may in fact be acting through modulation of intrinsic conductances. For

example, AngII has been shown to inhibit the transient K+ current, IA [125,148,202],

while simultaneously increasing the activity of the transient Ca2+ current, IT [181] within

the hypothalamus. This clearly emphasizes the importance of combined extrinsic/intrinsic

mechanisms in controlling the overall activity of PVN neurons. Remarkably, the

contribution of altered intrinsic membrane properties (i.e., number, distribution and

functional properties of ion channels, neuronal structure and geometry, etc) to PVN

neuronal hyperactivity during hypertension has not been thoroughly investigated. Thus,

the work of this thesis focuses on elucidating how altered intrinsic mechanisms contribute

to the maintenance of hypertension.

1.8. Major ionic mechanisms controlling neuronal activity of PVN neurons

In general, neuronal excitability is influenced by an extensive array of ion

channels, including, among others, voltage-dependent sodium, chloride, calcium and

potassium (both high-threshold and low-threshold, and calcium-dependent), many of

which have been reported to be expressed in the PVN. Importantly, subthreshold

currents, which are activated at membrane potentials below the threshold for activation of

Na+ channels, play a major role in controlling the ability of a neuron to generate action

potentials. Thus, the present work focuses on this particular group of voltage-gated ion

8

Page 23: Functional Interplay Between Subthreshold Ion Channels in

channels. Among subthreshold ion channels, both the A-type K+ current (IA) and the T-

type Ca2+ current (IT) have been found within non-identified PVN parvocellular neurons

[133]. IA has been identified as a rapidly activating and inactivating transient outward

current. Activation of IA decreases excitability, and is known to play an important role in

the modulation of repetitive neuronal firing activity, by restricting the duration of the

action potential waveform, and by increasing the interspike interval [164,166,172]. IT is

also a transient current, that when activated results conversely in an increased neuronal

excitability. IT has long been attributed to mediate bursting firing discharge [129,185].

However, IT also has been shown to regulate continuous firing frequency [159] and the

repolarization phase of the action potential waveform [112,139,147]. While the general

properties of IA and IT have been previously studied in magnocellular neuroendocrine and

non-identified parvocellular neurons [133], and have been shown to have overlapping

voltage-dependent and kinetic properties, there is no information available on the

biophysical and functional properties of IA or IT in identified populations of PVN-PA

neurons.

Based on their general biophysical properties and function, it is clear that these

two opposing conductances “compete” at subthreshold membrane potentials to influence

neuronal firing discharge, as shown in thalamic and cerebellar neurons [27,142,146,155].

Thus, it is expected that if the balance between these conductances is shifted, one will

predominate over the other, altering neuronal excitability and firing discharge. In this

sense, roles for altered potassium and calcium currents in the pathophysiology of

hypertension have been supported by recent work. Specifically, a diminished IA activity

has been observed in NTS neurons of renal wrap and spontaneously hypertensive rats

9

Page 24: Functional Interplay Between Subthreshold Ion Channels in

[13,191], suggesting that an attenuation of IA could contribute to increased neuronal

excitability during hypertension. As well, an enhanced repetitive firing frequency in

spontaneously hypertensive ganglionic sympathetic neurons has been attributed to an

enhanced calcium conductance [207]. These results suggest that an altered balance

between IA and IT contributes to neuronal hyper-excitability during hypertension.

Whether a balanced IA/IT activity influences neuronal function in PVN-PA neurons, and

whether a shift in this balance contributes to enhanced PVN neuronal activity during

hypertension is at present unknown, and is a major focus of the present work.

1.9. Thesis focus and significance

This research has focused on understanding the contribution of the subthreshold

ion channels, IA and IT, in regulating the activity of preautonomic PVN neurons, under

normal and pathological (i.e., hypertensive) conditions. To this end, a multidisciplinary

experimental approach, involving a variety of complementary techniques was employed,

including whole-cell patch-clamp electrophysiology, immunohistochemistry, real-time

PCR and confocal calcium imaging. The work is organized in three major parts. The first

part of the work focused on characterizing the basic biophysical and functional properties

of IA in PVN-RVLM neurons. These data demonstrate the presence of IA in PVN-RVLM

neurons, which actively modulates the action potential waveform and firing activity.

This is significant in supporting intrinsic conductances as an important mechanism in

controlling neuronal excitability in this presympathetic neuronal population. In the

second part of this work, I explored how alterations in the biophysical properties of IA

contributed to enhanced neuronal excitability of PVN-RVLM neurons during

10

Page 25: Functional Interplay Between Subthreshold Ion Channels in

hypertension. The data indicate that diminished IA availability constitutes a contributing

mechanism underlying aberrant neuronal function and autonomic control during

hypertension. Finally, the last part of the study specifically assessed the biophysical

interplay between subthreshold ion channels, IA and IT, and functionally addressed the

influence of such balance on PVN neuronal activity. Moreover, I determined whether a

shift in this balance contributed to enhanced PVN-RVLM neuronal excitability in

hypertensive conditions. The data indicate that the balance between IA and IT is shifted

during hypertension in favor of IT activity, resulting in increased IT-dependent influences,

elevated intracellular calcium levels, and enhanced basal spontaneous firing activity.

Taken together, the data presented here confirm the importance of intrinsic factors in

regulating the central control of cardiovascular output under normal and pathological

conditions.

11

Page 26: Functional Interplay Between Subthreshold Ion Channels in

II. MATERIALS AND METHODS

Male Wistar rats (120-140g and 200-300g) were purchased from Harlan

Laboratories (Indianapolis, IN, USA), and housed in a 12 h : 12 h light-dark cycle with

access to food and water ad libitum. All procedures were carried out in agreement with

both the University of Cincinnati and Wright State University Institutional Animal Care

and Use Committees’ guidelines, and in compliance with NIH guidelines.

Renovascular surgery

Rats weighing between 150-180 g (approximately 5-6 weeks old) were used to

induce the renovascular 2-kidney, 1-clip Goldblatt hypertension model, a well

characterized and widely used model [15,136]. Rats were anesthetized with isoflurane (3

%) throughout the surgery. Following an abdominal incision, the left kidney was

exposed, and a 0.2 mm clip was placed over the left renal artery, partially occluding it

[25]. Sham rats were subjected to the same surgical procedure, although the artery was

not occluded. Blood pressure was measured at the beginning of the sixth week post-

surgery, using a tail-cuff method. All rats were used for experiments during the sixth-

seventh week post-surgery.

Retrograde labeling of PVN-RVLM neurons

Preautonomic RVLM-projecting PVN neurons were identified by injecting

rhodamine beads unilaterally into the brainstem region containing the RVLM, as

12

Page 27: Functional Interplay Between Subthreshold Ion Channels in

previously described [124]. Rats were anesthetized intraperitoneally with a ketamine-

xylazine mixture (90 and 50 mg kg-1, respectively), the rat’s head was then placed in a

stereotaxic apparatus, and 200 nl of rhodamine-labelled microspheres (Lumaflor, Naples,

FL, USA) were pressure injected into the RVLM (starting from Bregma: 12 mm caudal

along the lamina, 2 mm medial lateral, and 8 mm ventral). In general, injection sites

were within the caudal pole of the facial nucleus to ~ 1 mm more caudal, and were

ventrally located with respect to the nucleus ambiguus (see example in Fig. 1A). In a

subset of experiments, injections were also performed in the area of the dorsal vagal

complex (DVC) (at the level of the obex: 1 mm lateral to the midline and 0.8 mm below

the dorsal surface) to label DVC-projecting PVN neurons [185] as a control neuronal

population. The location of the tracer was verified histologically as previously described

[124,185]. In a few instances, injections were located either rostrally or caudally to the

RVLM, in which cases no PVN retrograde labeling was observed. If the injection site

was not within the region of the RVLM, the experiment was discarded.

Hypothalamic slices

Two to seven days after the retrograde injection rats were anesthetized with

nembutol (50 mg kg-1) and perfused through the heart with a cold sucrose solution

(containing in mM: 200 sucrose, 2.5 KCl, 3 MgSO4, 26 NaHCO3, 1.25 NaH2PO4, 20 D-

glucose, 0.4 ascorbic acid, 1 CaCl2 and 2 pyruvic acid (290-310 mosmol l-1). This

method has previously been shown to improve cell viability in slices obtained from adult

rats [1]. Rats then were quickly decapitated, and brains dissected out. Slices were cut

coronally (300 μm thick) utilizing a vibroslicer (D.S.K. Microslicer, Ted Pella, Redding,

CA, USA). An oxygenated ice cold artificial cerebrospinal fluid (ACSF) was used

13

Page 28: Functional Interplay Between Subthreshold Ion Channels in

during slicing (containing in mM: 119 NaCl, 2.5 KCl, 1 MgSO4, 26 NaHCO3, 1.25

NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 2 CaCl2 and 2 pyruvic acid; pH 7.4; 290-310

mosmol l-1). After sectioning, slices were placed in a holding chamber containing ACSF

and then kept at room temperature until used.

Electrophysiological recordings

Slices were placed in a submersion style recording chamber, and bathed with

solutions (~ 3.0 ml min-1) that were bubbled continuously with a gas mix of 95% O2-5%

CO2, and maintained at room temperature (~ 22 °C). Thin-walled (1.5 mm o.d., 1.17 mm

i.d.) borosilicate glass (G150TF-3, Warner Instruments, Sarasota, FL, USA) was used to

pull patch pipettes (3-6 MΩ) on a horizontal Flaming/Brown micropipette puller (P-97,

Sutter Instruments, Novato, CA, USA). The internal solution contained (in mM): 140

potassium gluconate, 0.2 EGTA, 10 Hepes, 10 KCl, 0.9 MgCl2, 4 MgATP, 0.3 NaGTP

and 20 phosphocreatine (Na+); pH 7.2-7.3. Whole-cell recordings from PVN neurons

were visually made using a combination of fluorescence illumination and infrared

differential interference contrast (IR-DIC) video-microscopy. Recordings were obtained

with a Multiclamp 700A amplifier (Axon Instruments, Union City, CA, USA). The

voltage output was digitized at 16-bit resolution, 10 kHz (Digidata 1320A, Axon

Instruments), and saved on a computer to be analyzed offline. The series resistance was

monitored at the beginning and end of each experiment, and the experiment was

discarded if the series resistance was not stable throughout the recording. The liquid

junction potential (LJP, 6.5 mV) was experimentally determined using a 2 M KCl agar

bridge. Data shown were corrected for the LJP.

Voltage-clamp recordings of isolated voltage-gated K+ and Ca2+ currents

14

Page 29: Functional Interplay Between Subthreshold Ion Channels in

For isolated A-type K+ currents (IA), slices were bathed in an ACSF with nominal

Ca2+ (0 mM) (containing in mM: 102 NaCl, 2.5 KCl, 3 MgSO4, 26 NaHCO3, 1.25

NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 2 pyruvic acid, 3 EGTA, 200 μM CdCl2, 30

TEA and 0.5 μM TTX; pH 7.4; 290-310 mosmol l-1). The T-type Ca2+ currents (IT) were

isolated by using an ACSF that enhances the signal to noise ratio of calcium currents (5

mM CaCl2), blocks sodium channels, delayed rectifier potassium channels (IKDR), as well

as the transient potassium channels, IA (containing in mM: 92 NaCl, 2.5 KCl, 1 MgSO4,

26 NaHCO3, 1.25 NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 5 CaCl2, 2 pyruvic acid, 0.5

μM TTX, 30 TEA and 5 4-AP; pH 7.4; 290-310 mosmol l-1). Series resistance was

electronically compensated for at least 60% throughout the recordings. The voltage error

due to uncompensated series resistance at the half-activation and half-inactivation

potentials for IA were calculated, however results were not corrected for them.

The quality of the space clamp was assessed as previously described [133].

Briefly, IA 10-90% rise time and time constant (τ) of inactivation were measured

following activation of the current by a test command (-10 mV), preceded by

conditioning steps of varying amplitudes (-120 mV through -30 mV). Plots of 10-90%

rise time and inactivation time constant (τ) as a function of conditioning steps were then

generated. Varying the conditioning step should affect only the amplitude of the current,

without affecting its kinetic properties. Thus, only neurons showing unchanging 10-90%

rise time and inactivation τ as a function of the conditioning pulse, as well as a lack of

relationship between current amplitude and kinetics were included for analysis.

All protocols were run with an output gain of 2 and a Bessel filter of 2 kHz, and

were leak subtracted using a P/4 protocol.

15

Page 30: Functional Interplay Between Subthreshold Ion Channels in

Voltage-dependence of activation of IA and IT

In order to isolate IA, a combination of electronic and pharmacological methods

were used. Calcium channels were blocked using a 0 Ca2+ ACSF containing EGTA and

CdCl2 (see above). TTX and tetraethyl ammonium (TEA) were also used to block

voltage-dependent Na+ channels and delayed rectifier K+ channels (IKDR), respectively.

Since in many instances some TEA insensitive IKDR remained, two separate

electrophysiological protocols were run in order to further isolate IA electronically. The

first utilized a hyperpolarized conditioning pulse (-90 mV), which removed inactivation

from IA. This pulse was followed by depolarizing command pulses (-70 to +25 mV),

which resulted in the activation of both IA and IKDR. A second protocol was then run, in

which a more depolarized (-40 mV) conditioning pulse was used to completely inactivate

IA. Thus, when the same command pulses as above were applied, only IKDR was

activated. Currents recorded under these two protocols were then electronically

subtracted offline using Clampfit 8.2 (Axon Instruments).

In order to isolate IT, a combination of electronic and pharmacological methods

also were used. TTX, TEA and 4-aminopyridine (4-AP) were used to block voltage-

dependent Na+ channels, IKDR, and transient subthreshold potassium channels (IA). To

isolate IT from high-threshold calcium currents (HVA), the same protocol as above was

used, however, the command pulses were only depolarized to -20 mV, limiting activation

to that of IT only. The maximum peak amplitude and half-activation of IT were,

therefore, unable to be measured. In a subset of recordings IA and IT were both activated

in a single protocol using a modified ACSF (containing in mM: 97 NaCl, 2.5 KCl, 1

MgSO4, 26 NaHCO3, 1.25 NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 5 CaCl2, 2 pyruvic

16

Page 31: Functional Interplay Between Subthreshold Ion Channels in

acid, 0.5 μM TTX and 30 TEA; pH 7.4; 290-310 mosmol l-1). The same protocol for the

isolation of IT (see above) was used.

For IA, the chord conductance was calculated by measuring the peak amplitude of

the evoked current at each command potential, divided then by the difference of the

command potential and the reversal potential (calculated to be -104.2 mV from the

Nernst equation). The chord conductance was then normalized to the maximum chord

conductance obtained at +25 mV, and plotted as a function of the command potential.

The plots were then fit with a Boltzmann function, and the half-activation potential (the

Vm at which 50% of IA currents are activated) was obtained.

The current densities for IA and IT were determined by dividing the current

amplitude at each command potential by the cell capacitance, obtained by integrating the

area under the transient capacitive phase of a 5 mV depolarizing step pulse, in the

voltage-clamp mode. The rate of activation of IA was determined by measuring the 10-

90% rise time from the baseline to the peak of the current (command potential = -10

mV), while that of IT was measured at a command potential of -40 mV.

Voltage dependence of inactivation of IA

In order to determine the voltage-dependence of inactivation, neurons were

voltage-clamped at -70 mV, and the membrane was subjected to conditioning pulses of

varying amplitude (-120 to -35 mV, 50 ms), which removed varying amounts of

inactivation from IA. A command pulse to -10 mV was then used to activate IA. In

separate sets of experiments, the duration of the prepulses were extended to 115, 120 and

150 ms, as indicated. The mean normalized IA peak amplitude was plotted as a function

of the conditioning step potentials, and the I-V plots were fitted with a Boltzmann

17

Page 32: Functional Interplay Between Subthreshold Ion Channels in

function, to determine the half-inactivation potential (the Vm at which 50% of IA is

inactivated). The voltage-dependence of IT was not determined in this study. The tau

(τ) of inactivation of IA and IT was determined by fitting a single exponential function to

the decay phase of the current activated at -10 mV and -40 mV, respectively, following a

conditioning step to -90 mV.

Time-dependence of inactivation of IA

To determine the time-dependence of IA inactivation, neurons were voltage-

clamped at -70 mV, and conditioning pulses to -45 mV or -50 mV of varying durations

(10 to 200 ms, 10 ms increments) were followed by a command pulse to -10 mV (300

ms). Plots of the IA peak amplitude as a function of the duration of the conditioning pulse

were then generated. Plots were fit by a monoexponential function, and the time constant

of the decay was used for quantitative purposes.

Kinetics of recovery from inactivation of IA

Once the A-type K+ channel is inactivated following membrane depolarization, a

sufficient amount of time, in which the membrane is hyperpolarized, must elapse before

the channel can recover and be fully activated again. In order to determine the kinetics of

recovery from the inactivated channel state, neurons were voltage-clamped at -50 mV,

and a hyperpolarizing conditioning pulse (-100 mV) of increasing duration (Δ10 ms) was

applied, followed by a depolarizing command pulse (-10 mV). The mean normalized

peak amplitude was plotted against the conditioning pulse duration. A single exponential

function was fit to the plot and the time constant (τ) of recovery from inactivation was

then calculated.

Non-stationary fluctuation analysis of IA

18

Page 33: Functional Interplay Between Subthreshold Ion Channels in

For these studies, we followed procedures originally described by Sigworth

[176,177] and subsequently used by others [5,53,163,197]. Traces of IA (n= 100-130

traces) evoked with a command pulse to +40 mV were obtained and analyzed with Mini

Analysis software (Synaptosoft, Fort Lee, NJ, USA). The mean evoked current was

scaled to individual IA waveforms. The difference between the scaled mean and the single

current resulted in difference currents with a variance that was higher than background

levels. The variance of the individual IA current around the scaled average was then

computed and variance-amplitude relationships were plotted and fit with a parabolic

function. Values of unitary current, open probability and number of channels were

provided by Mini Analysis algorithms.

Current-clamp recordings of action potential waveform and firing activity

For most current-clamp experiments, the ACSF used contained (in mM): 119

NaCl, 2.5 KCl, 1 MgSO4, 26 NaHCO3, 1.25 NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 2

CaCl2, 2 pyruvic acid; pH 7.4; 290-310 mosmol l-1. In addition, the AMPA and NMDA

glutamate receptor antagonists DNQX (10 μM) and AP-5 (100 μM), respectively, or the

non-selective glutamate receptor antagonist kynurenic acid (2 mM), as well as the

GABAA receptor antagonist bicuculline (40 μM), or picrotoxinin (0.3 mM) were added to

this solution (see below). All protocols run used an output gain of 10 and a Bessel filter

of 10kHz.

To test for the influence of IA on PVN-RVLM firing properties, the K+ channel

blocker 4-aminopyridine (4-AP) was used. Since 4-AP facilitates pre-synaptic release of

neurotransmitter [67], direct effects of 4-AP on intrinsic properties could be masked by

this presynaptic effect. Thus, all current-clamp experiments were performed in the

19

Page 34: Functional Interplay Between Subthreshold Ion Channels in

presence of receptor blockers of the main excitatory and inhibitory neurotransmitters in

this system: GABA and glutamate (see above). Supporting the efficacy of this approach,

we found that in the absence of these receptor blockers, 4-AP induced a significant

decrease in PVN-RVLM input resistance (control = 1414 ± 462.7 MΩ; 4-AP = 829.4 ±

244.3 MΩ; n=7; P= 0.05), an effect likely due to overall increased neuronal conductance

following robust release of neurotransmitters. Conversely, an increased input resistance

was observed when 4-AP was applied in the presence of the receptor blockers listed

above (control = 886.1 ± 127.0 MΩ; 4-AP = 1167 ± 147.5 MΩ; n= 23; P< 0.05). When

needed, 4-AP was bath applied using a peristaltic pump (Gilson, Middleton, WI, USA;

flow ~ 2 ml/min) for a period of 5 minutes before recording its effects. In most cases, in

our hands, a complete washout of 4-AP and its effects were not accomplished within the

period that we were able to maintain a good quality recording (see, however, Fig. 6A).

Thus, values corresponding to the washout period are not reported.

Low-threshold spikes

Evoked low-threshold spikes (LTS) were simultaneously recorded with changes

in intracellular calcium levels, before and after bath application of 5 mM 4-AP. The

ACSF used, contained (in mM): 110 NaCl, 2.5 KCl, 1 MgSO4, 26 NaHCO3, 1.25

NaH2PO4, 20 D-glucose, 0.4 ascorbic acid, 5 CaCl2, 2 pyruvic acid, 0.5 μM TTX, 300

μM picrotoxinin and 2 mM kynurenic acid; pH 7.4; 290-310 mosmol l-1). Thus, isolated

LTS (no Na+ spikes) were evoked, from a Vm held at ~ -90 mV, by injecting

depolarizing pulses (20-100 pA, 220ms). The LTS threshold was obtained by fitting a

single monoexponential function to the trace and determining at what Vm the function no

longer fits the trace. From threshold the LTS trace was then baselined to zero. The LTS

20

Page 35: Functional Interplay Between Subthreshold Ion Channels in

area (under the curve) and peak amplitude were then measured. For LTS experiments

combined with calcium imaging, recordings were obtained with an Axopatch 200B

amplifier (Axon Instruments). Those without calcium imaging were obtained as

mentioned above.

Evoked Action Potentials

To elicit individual action potentials, PVN-RVLM neurons were current-clamped

at either -80 mV or -50 mV, and subjected to depolarizing pulses (5 ms; 0.5-1.0 nA). At

least ten sweeps of evoked action potentials were averaged together, and the mean action

potential half-width and 90-10% decay time were analyzed and compared before and

after addition of the A-type K+ channel blocker 4-AP, using algorithms provided by Mini

Analysis software (Synaptosoft, Fort Lee, NJ, USA).

Repetitive firing activity

Spontaneous or evoked (DC current injection) firing discharge was recorded from

PVN-RVLM neurons in continuous mode. The mean firing frequency obtained before

and after addition of 4-AP or the voltage-dependent calcium channel inhibitor, NiCl2

(100 μM) (2 min period) were calculated and compared using Mini Analysis software

(Synaptosoft). Neurons were arbitrarily considered responsive to 4-AP or NiCl2 if a

change in firing rate >5% was observed.

In addition, all action potentials from the control and treated group were averaged

into a single spike waveform, and the half-width and 90-10% decay time of the action

potential were compared. In addition, parameters related to the hyperpolarizing

afterpotential (HAP), including peak amplitude, area and kinetics, were calculated and

compared before and after 4-AP addition. Action potential threshold was measured at the

21

Page 36: Functional Interplay Between Subthreshold Ion Channels in

abrupt transition from the pre-spike depolarizing ramp to the up-stroke of the action

potential, as determined by algorithms provided by Mini Analysis software (Synaptosoft).

HAP properties (e.g., amplitude, area and decay time course) were also determined by

algorithms provided by Mini Analysis software (Synaptosoft).

To study spike broadening during repetitive firing, neurons were held at a Vm

~ -90 mV and depolarizing pulses (110-130 pA, 180 ms) were used to evoke firing

discharge. The half-widths of the evoked action potentials were calculated. The degree of

spike broadening was quantified as the ratio of the half-width of the third spike to that of

the first spike (Shao et al., 1999; Stern, 2001).

Confocal calcium imaging

Dye-loading into identified retrogradely-labeled PVN-RVLM neurons was

achieved by addition of fluo-5F pentapotassium salt (100 μM; Molecular Probes,

Carlsbad, CA, USA) into the internal pipette solution. Once whole cell was established,

the dye was allowed to diffuse into the cell for at least 20 minutes before the initiation of

the recordings. Calcium imaging was conducted using the Yokogawa real time live cell

laser confocal system (CSU-10) combined with a highly-sensitive EMCCD camera

(iXon+885, Andor Technology, South Windsor, CT, USA). Fluorescence images were

obtained using diode-pumped solid-state laser (Melles Griot, Carlsbad, CA, USA), and

fluorescence emission was collected at >495 nm. Images were acquired at a rate of 40 -

50 Hz. The fractional fluorescence (F/F0) was determined by dividing the fluorescence

intensity (F) within a region of interest (ROI; 6 x 6 pixels ≈ 4.9 x 4.9 μm) by a baseline

fluorescence value (F0) determined from 30 images before single action potentials were

evoked (a period showing no change in intracellular calcium levels) [63]. Between 8-10

22

Page 37: Functional Interplay Between Subthreshold Ion Channels in

traces of calcium transients were averaged together in order to increase the signal to noise

ratio. Data was analyzed using Andor IQ software (Andor Technology).

Single cell real time, reverse transcription-polymerase chain reaction

The cytoplasm of a single neuron was gently pulled into a pipette with negative

pressure, taking care not to contain the nucleus. The cytoplasm in the pipette was

dissipated into a prepared tube containing (in μl): 18 of nuclease free water, 3 of 7x

genomic DNA Wipeout Buffer and stored at -70 °C. After finishing all recordings, the

tubes were heated to 42 °C for 2 min and then incubated on ice for at least 1 min. The

mixture of (in μl) 6 of 5x Quantiscript RT Buffer, 1 of RT Primer Mix, 1 of Quantiscript

Reverse Transcriptase was subsequently added and incubated at 42 °C for 15 min. The

reaction was terminated by heating at 95 °C for 3 min and stored at -20 °C. All reagents

for reverse transcription were purchased from Qiagen (Valencia, CA, USA). Real time

PCR amplification was induced by using a fraction of the single cell cDNA as a template.

The cDNA of the single neuron was split by 4 μl for three primer sets of GAPDH, Kv4.3

and Cav3.1. The reaction mixture (20 μl total volume) contained (in μl): 1 of 25 μM each

forward and reverse primer, 4 of nuclease free water and 4 of the cDNA template, 10 of

2x SYBR Green Master Mix (Applied Biosystems, Foster City, CA, USA). The

annealing temperature in the thermal cycler was 60 °C and 50 cycles were performed

using an ABI Prism 7700 sequence detector (Applied Biosystems). Quantification was

conducted by determining the relative changes in gene expression, to a chosen reference

gene, using the 2-ΔΔCT method [128]. Statistical significance was determined by

comparing raw ΔCt values. The standard deviations were used to calculate range values

23

Page 38: Functional Interplay Between Subthreshold Ion Channels in

(2-(ΔΔCT ± S.D. of ΔCT)) for graphs. We compared the efficiency between the target gene

(Kv4.3 or Cav3.1) and GAPDH by plotting the delta Ct versus the log of the dilution

ratio. The primers were considered efficient if the absolute value of the slope was less

than 1. The slopes of Kv4.3 and Cav3.1 were determined to be 0.07 and 0.005,

respectively. All primers except GAPDH used in this study were designed by Primer 3

(Whitehead Institute, Cambridge, MA, USA) and synthesized by Bioneer (Alameda, CA,

USA). The primer set used in this study is presented in Table #1. The product size was

adjusted around 150 bp that is suitable for the real time PCR reaction. The primer for

GADPH was purchased from Qiagen.

Neuronal morphometry

In a subset of recordings, cells were intracellularly filled with biocytin (0.2%) and then

stained with the avidin-biotin complex (ABC)-diaminobenzidine tetrahydrochloride

(DAB) as previously described [185]. Briefly, after recordings were completed, slices

were placed in a 4% paraformaldehyde and 0.2% picric acid solution, dissolved in 0.3 M

PBS (pH~ 7.3) overnight and then thoroughly rinsed with 0.01 M phosphate buffered

saline (PBS). Slices then were incubated at 4 °C for 1 hour in 10% normal horse serum

with 0.01 M PBS and 0.5% Triton X-100. Slices were again thoroughly rinsed with 0.01

M PBS and incubated overnight in ABC (Vector Laboratories, Burlingame, CA, USA)

diluted 1:100 in 0.01 M PBS containing 0.5% Triton X-100. Slices were then reacted

with DAB (60 mg/100 ml) in 0.01 M PBS containing 0.5% Triton X-100, 0.05% nickel

sulfate, and 0.006% H2O2, for approximately 2-3 minutes. Sections then were rinsed in

0.01 M PBS, mounted, and dried for 24 hours [185]. For morphometric analysis, the

24

Page 39: Functional Interplay Between Subthreshold Ion Channels in

Real Time PCR Primers

Gene (Accession No)

Sequence (forward/reverse)

Tm ()

Product size (bp)

Kv4.3 (U42975)

5’-AGGCTTCTTCATTGCGGTCT-3’ 5’-GTGTCCAGGCAGAAGAAAGC-3’

60 126

Cav3.1 (AF027984)

5’-CAGTGTGTGGGAGATTGTGG-3’ 5’-TGTCCATGGTCTTCATGAGC-3’

60 140

Table 1. List of primers for A-type and T-type channel subunits

used in the single cell real time PCR reactions.

25

Page 40: Functional Interplay Between Subthreshold Ion Channels in

entire somatic and dendritic compartments were reconstructed three-dimensionally using

a tracing system (Neurolucida, Microbrightfield). Reconstructions were performed using

a x60 oil objective, and the course of each dendrite was traced by digitizing the X, Y, and

Z coordinates as well as its width. Several parameters, including dendritic length, surface

area and volume were calculated by algorithms provided by Neurolucida. Axons were

identified by their thinner diameter and beaded appearance, and were cut at the slice

surface [185]. Damaged (e.g., somatic swelling, missing or cut dendritic trees) or weakly

labeled neurons were not included in the analysis [185]. Data were not corrected for

tissue shrinkage.

Immunohistochemistry

In order to determine which A-type channel subunits are expressed in PVN-

RVLM neurons, these neurons were retrogradely labeled using cholera toxin B (CTB;

1%, List Biological Laboratories), using the same stereotaxic procedure as described

above. Two to seven days after surgery, rats were anesthetized with nembutol (50 mg kg-

1) and perfused transcardially in 4% paraformaldehyde in 0.01 M phosphate buffer saline

(PBS). Brains were then removed, postfixed for 2-4 hours, cryoprotected in 30%

sucrose in 0.01 M PBS (4 °C, 3 days), and then stored at -80 °C until further use.

Coronal slices (30 μm) containing the PVN were cut and collected in 0.01 M

PBS. Slices then were incubated in 0.01 M PBS with 0.1% Triton X-100, 0.04% NaN3

(PBSTXNaN3), and 5% normal horse serum for 1 hour at room temperature. Slices were

then rinsed thoroughly with 0.01 M PBS, followed by incubation with one of the

following primary antibodies for A-type K+ channel subunits (Kv1.4 1:100, Kv4.2 1:500,

and Kv4.3 1:10000; Alomone Labs, Jerusalem, Israel), along with an anti-CTB antibody

26

Page 41: Functional Interplay Between Subthreshold Ion Channels in

(goat anti-CTB 1:2500; List Biological Laboratories) for 2 days at 4 °C in PBSTXNaN3.

Slices were again thoroughly rinsed. Secondary antibodies were then applied for 4 hours

at 4 °C in PBSTXNaN3 (donkey anti-goat Cy-5: 1:50 and donkey anti-rabbit FITC: 1:250;

Jackson ImmunoResearch Laboratories, West Grove, Pennsylvania, USA). Slices were

then rinsed thoroughly, mounted, and visualized using confocal microscopy (Carl Zeiss

MicroImaging, Inc., Thornwood, NY, USA); 63x oil immersion, zoomed x2; single

optical plane = 0.5 μm thick).

Pre-adsorption controls were run in order to test the specificity of the Kv primary

antibodies. The antigens were used at concentrations 5x those of the primary antibodies

(Kv1.4 15 μg/ml, Kv4.2 8 μg/ml, and Kv4.3 0.4 μg/ml; Alomone Labs) and incubated in

PBSTXNaN3 for 2 hours at room temperature, with or without the primary antibodies.

The solutions were then centrifuged for 5 minutes at 10000xg, and the supernatants

applied to the tissue, following the same procedure described above.

In a subpopulation of recordings, in order to determine whether PVN-RVLM

neurons expressed vasopressin and/or oxytocin, neurons were intracellularly filled with

biocytin (0.2%). Slices were briefly fixed overnight in a 4% paraformaldehyde-0.2%

picric acid solution, dissolved in 0.3 M phosphate buffered saline (PBS; pH~7.3) and

then thoroughly rinsed with 0.01 M PBS. Slices were then taken through a dehydration

procedure, in which slices were incubated in increasing concentrations of ethyl alcohol

(60-100%; 10% increments; 10 min. each step except 100% for 20 min.). Slices were

then incubated in xylene for 10 min. followed by the reverse ethyl alcohol procedure

(100-60%). Slices were then rinsed in 0.01 M PBS with 0.5% Triton X-100 (TX) for 10

minutes. Slices were then incubated for 45 minutes in 10% normal horse serum with

27

Page 42: Functional Interplay Between Subthreshold Ion Channels in

0.01 M PBS, 0.5% TX and 0.04% NaN3. Slices were then thoroughly rinsed with 0.01 M

PBS, 0.5% TX and 0.04% NaN3, followed by incubation with the primary antibodies for

2 days (1 day at room temperature and 1 day at 4°C): OTgp and VPgp (both 1:50,000;

Bachem, Torrance, CA, USA) in 0.01 M PBS, 0.5% TX and 0.04% NaN3. Slices were

then rinsed in 0.01 M PBS, 0.5% TX and 0.04% NaN3 for 30 minutes. Slices were then

incubated with the secondary antibodies for 1 day (room temperature): CY5-streptavidin

(1:10,000) and FITCgp (1:400; both from Jackson ImmunoResearch Laboratories) in

0.01 M PBS, 0.5% TX and 0.04% NaN3. Slices were then thoroughly rinsed in 0.01 M

PBS for 20 minutes, mounted, and visualized using fluorescence microscopy (20x;

Olympus America Inc., Melville, NY, USA).

Chemicals

All chemicals were obtained from Sigma-Aldrich (St Louis, MO), with the

exceptions of pyruvic acid (MP Biomedicals, Aurora, OH) and tetrodotoxin (Alomone

Labs, Jerusalem, Israel).

Statistical analysis

All values are expressed as means ± S.E.M. Student’s paired t test was used to

compare differences in various physiological parameters, as indicated in the text.

Unpaired t test was used to compare between group differences, as indicated. A one- or

two-way ANOVA with Bonferroni’s post hoc test were used when appropriate. Fisher’s

exact test was used to determine differences in the incidence of a particular effect, and

Pearson’s correlation test was used to determine if correlations existed between two

parameters. Differences were considered statistically significant at a P< 0.05. All

28

Page 43: Functional Interplay Between Subthreshold Ion Channels in

statistical analyses were conducted using the same statistical software, GraphPad Prism

(GraphPad Software, San Diego, CA, USA).

29

Page 44: Functional Interplay Between Subthreshold Ion Channels in

III. RESULTS

3.1. Functional Role of A-type Potassium Currents in Rat Presympathetic PVN

Neurons

3.1.1. Intrinsic properties of PVN-RVLM neurons

Whole-cell patch clamp recordings were obtained from 119 retrogradely-labeled

PVN-RVLM neurons, located in the three main PVN subnuclei known to contain long-

descending preautonomic neurons: the ventromedial, dorsal cap and posterior subnuclei

[185,192,194]. In many instances, neurons were efficiently loaded intracellularly with

biocytin, and identified following an ABC-DAB procedure (see Fig 1B). Overall, 40, 16

and 24 neurons were found to be located within the ventromedial, dorsal cap and

posterior subnuclei, respectively. Mean input resistance and cell capacitance of recorded

neurons were 992.7 ± 49.3 MΩ, and 35.2 ± 1.4 pF, respectively. In current-clamp mode,

most PVN-RVLM neurons displayed low threshold spikes (LTS) in response to positive

current injection from hyperpolarized membrane potentials (~ -90 mV). Similarly to our

previous study in PVN neurons innervating the dorsal vagal complex [185], LTSs in

PVN-RVLM neurons varied greatly in shape and amplitude, including fast spikes, small

humps, and long lasting plateaus (Fig 1C).

In voltage-clamp mode, using solutions that allowed isolation of K+ currents (see

Methods), membrane depolarization evoked both transient (A-type, IA) and sustained

outward (IKDR) components. To pharmacologically isolate the transient component, 30

30

Page 45: Functional Interplay Between Subthreshold Ion Channels in

Figure 1. PVN-RVLM projecting neurons are identified following microinjections

of a fluorescent retrograde tracer in the RVLM. A- Representative example of a

retrograde tracer injection site in the RVLM at three rostrocaudal brainstem levels (A1:

Bregma – 11.30, A2: Bregma – 11.96; A3, Bregma – 14.08). Sections in the right panels

show the rhodamine beads injection site (arrowheads). Bright light and fluorescent

images were superimposed to better depict the injection site. Images on the left panels

were obtained from similar rostrocaudal levels, and were counterstained to better depict

the anatomy of the region. Arrows in A1, A2 and A3 point to the facial nucleus, the

nucleus ambiguus and the area postrema, respectively. CC, central canal. B- A

representative example of an intracellularly labeled PVN neuron located in the posterior

subnucleus is shown in B1. The same neuron is shown at an expanded scale in B2, along

with a following an ABC-DAB staining. C- PVN-RVLM neurons displayed low

threshold spikes (LTS) (arrows) of varying shapes and magnitudes.

31

Page 46: Functional Interplay Between Subthreshold Ion Channels in

32

Page 47: Functional Interplay Between Subthreshold Ion Channels in

mM TEA was used [130,166]. However, since a remaining, TEA-insensitive sustained

component was still present at relatively depolarized membrane potentials, IA was further

isolated electronically (see Methods and Fig. 2). Since the focus of the present study was

on IA, the properties of IKDR were not further studied herein.

3.1.2. Activation properties of IA in PVN-RVLM neurons

As previously described in other neuronal types [12,94], A-type K+ currents (IA)

in PVN-RVLM neurons were characterized by strong voltage-dependency and rapid

activation and inactivation kinetics, resulting in a transient outward current.

The voltage-dependent activation properties of IA were studied in 26 PVN-RVLM

neurons. Depolarizing steps of increasing amplitudes (-70 mV to +25 mV, in 5 mV

increments) were used to activate IA (Fig. 2A). The mean IA peak amplitude, current

density, and chord conductance at +25 mV were 933.8 ± 110.8 pA, 31.7 ± 4.0 pA•pF-1,

and 8.3 ± 1.4 nS, respectively (see Methods). The mean IA 10-90% rise time (see

Methods) was determined to be 6.1 ± 0.4 ms at a command potential of -10 mV (Fig.

2B). Plots of chord conductance vs. command potential were generated (see Methods),

and the voltage-dependent properties of activation of IA were calculated using a

Boltzmann fit (Fig. 2C). The mean IA activation threshold was -48.9 ± 1.4 mV, and the

half-activation voltage was -24.5 ± 1.6 mV, with a slope factor of 11.7 ± 0.6 mV.

3.1.3. Inactivation properties of IA in PVN-RVLM neurons

The voltage and time-dependent inactivation properties of IA were studied in 35

PVN-RVLM neurons. The time-dependence of inactivation of IA was studied using a

command pulse to -10 mV (300 ms) from a conditioning step of either -50 mV or -45 mV

of successively longer durations (10-200 ms, 10 ms increments). Plots of the evoked IA

33

Page 48: Functional Interplay Between Subthreshold Ion Channels in

Figure 2. Isolation and voltage-dependent activation of IA in PVN-RVLM neurons.

A- Outward currents were activated in the presence of 30 mM TEA using depolarizing

steps (from -70 to +25 mV in 5 mV increments, 400 ms) following a conditioning step to

-90 mV (340 ms). Responses included a transient (IA, arrow) and a TEA-insensitive

sustained component (IKDR, arrowheads) (A1). Using a conditioning step to -40 mV, the

same depolarizing steps resulted only in the activation of IKDR (A2). IA was then

electronically isolated by digitally substracting traces in A1 and A2 (A3). B-

Representative example of the rapid rate of activation of IA following a command

potential to -10 mV. C- Plot of the normalized chord conductance versus the command

potential was created, and a Boltzmann function was fit to the I-V plot. The mean half-

activation potential was -24.5 ± 1.6 mV (n= 23), with a slope factor of 11.7 ± 0.6 mV.

Data were corrected for LJP.

34

Page 49: Functional Interplay Between Subthreshold Ion Channels in

35

Page 50: Functional Interplay Between Subthreshold Ion Channels in

amplitude as a function of the conditioning step duration were generated (see Fig. 3A).

As shown, the amplitude of the evoked IA current rapidly decreased as a function of

duration of the -45 mV conditioning step, reaching steady-state within a range of 80-110

ms (n= 6). Plots were fit with a monoexponential function, and a mean time constant of

28.3 ± 3.8 ms was obtained. Similar results were observed with conditioning steps of -50

mV (results not shown).

The voltage-dependence of inactivation was then studied. To remove variable

amounts of IA inactivation, neurons were depolarized using a command pulse to -10 mV,

from a range of conditioning steps (-120 to -35 mV, 5 mV increments, 50 ms duration),

preceded by a fixed pulse to -40 mV (65 ms) (Fig. 3B). I-V plots were generated and fit

with a Boltzmann function, and the voltage-dependent inactivation properties were then

calculated (see Methods) (Fig. 3C). The mean IA half-inactivation potential was -87.4 ±

3.1 mV, with a slope factor of 13.1 ± 0.5 mV. Similar values were obtained when

conditioning steps were prolonged in duration from 50 ms to 115, 120 and 150 ms

(results not shown).

The mean inactivation τ of IA using a command potential to -10 mV was 33.9 ±

3.0 ms (Fig. 3C inset), and was found to be independent of the command potential (F=

0.41; P= 0.8; one-way ANOVA, data not shown).

When the voltage-dependent activation and inactivation curves of IA were plotted

together, a small region of overlap between these two curves (i.e., “window current”) at

potentials between -55 mV to -40 mV, was observed (Fig 3D). Despite its small

amplitude (1.25% of maximal IA = ~12 pA at a membrane potential of -51.5 mV this

window current may contribute substantially to subthreshold changes in membrane

36

Page 51: Functional Interplay Between Subthreshold Ion Channels in

Figure 3. Voltage-dependent and kinetics of inactivation of IA. A- Time-dependence

of inactivation of IA (n= 6). A command pulse to -10 mV (300 ms) was applied from a

conditioning step of -45 mV of successively longer durations (10-200 ms, 10 ms

increments). Note that the amplitude of the evoked IA rapidly decreased as a function of

the duration of the conditioning step. The inset shows a plot of the evoked IA peak

amplitude as a function of the conditioning step duration. The plot was fit with a

monoexponential function, and a time constant (τ) of 27.0 ms was obtained. B-

Representative example of IA currents evoked by voltage steps to -10 mV from

conditioning step to between -120 and -35 mV in 5 mV increments (50 ms). C- The

mean normalized current amplitude was plotted against the conditioning potential and a

Boltzmann function was fit to the I-V plot. The mean half-inactivation potential was -

87.4 ± 3.1 mV (n=21), with a slope factor of 13.1 ± 0.5 mV. Data were corrected for

LJP. Inset, a single exponential function (thick line) was fit to the IA decay phase. The

mean time constant of inactivation (t) was 33.9 ± 3.0 ms at a command potential of -10

mV. D- When the mean normalized IA amplitude obtained with the voltage-dependent

activation (triangles) and inactivation (squares) protocols are plotted together as a

function of the command potential, a region of overlap between -55 mV and -40 mV is

observed (gray area). Note that the plots were expanded to better depict the overlapping

region. Data were corrected for LJP. Inset, summary data showing that blockade of IA

with 5 mM 4-AP induced a significant membrane depolarization (*P<0.005, n=4).

Neurons were current-clamped at ~ -50 mV.

37

Page 52: Functional Interplay Between Subthreshold Ion Channels in

38

Page 53: Functional Interplay Between Subthreshold Ion Channels in

potential, due to the relatively high input resistance of PVN-RVLM neurons. This is in

fact supported by our results showing that pharmacological blockade of IA with 5 mM 4-

AP (see below) induced a significant membrane depolarization (in 4/5 cells tested), when

neurons were current-clamped at ~ -50 mV in the presence of TTX (Δ Vm: 5.5 ± 0.5 mV,

P< 0.005, paired t-test, n= 4, see Fig 3D inset).

3.1.4. Recovery of IA from inactivation

The time-course of recovery of IA from inactivation was studied in 16 PVN-

RVLM neurons. To vary the amount of IA available for activation, neurons were

hyperpolarized to -100 mV using conditioning steps of increasing duration (10 to 250 ms,

in 10 ms increments). IA was then activated using a depolarizing command potential to -

10 mV (Fig. 4A). The normalized IA peak amplitude at each command was plotted as a

function of the duration of the respective conditioning step (Fig. 4A). The plots were best

fit by a single exponential function. As depicted in the example of Fig. 4B, IA recovery

from inactivation in PVN-RVLM neurons was strongly time-dependent, with a mean

recovery time constant (τ) of 65.7 ± 4.5 ms.

3.1.5. 4-AP inhibits IA in PVN-RVLM neurons

The sensitivity of IA to the K+ channel blocker 4-AP was studied in 17 PVN-

RVLM neurons. Similar protocols as those used to study activation of IA were used here.

Currents were recorded before and after bath application of 1 or 5 mM 4-AP (Fig. 5A).

Using a command step to -10 mV, we found IA to be significantly inhibited by both

concentrations used (1 mM 4-AP: 20.3 ± 2.8% inhibition, n= 5; 5 mM 4-AP: 44.3 ± 2.3%

inhibition, n= 12; P< 0.01 and P< 0.0001, compared to control, respectively). The larger

inhibition observed with 5 mM 4-AP (P< 0.0001, when compared to 1 mM 4-AP),

39

Page 54: Functional Interplay Between Subthreshold Ion Channels in

Figure 4. Time course of recovery from inactivation of IA. A- Representative

example of IA currents evoked with command steps to -10 mV (200 ms), from

conditioning steps to -100 mV of increasing duration (10 ms increments). Note that

longer conditioning steps removed increasing amounts of inactivation of IA, allowing for

larger IA amplitudes at the command step. B- The mean normalized current amplitude

evoked at the command test was plotted against the conditioning step duration. A single

exponential function was fit to the plot, and the mean time course of recovery from

inactivation (τ) of IA in PVN-RVLM neurons was calculated to be 65.7 ± 4.5 ms (n= 16).

40

Page 55: Functional Interplay Between Subthreshold Ion Channels in

41

Page 56: Functional Interplay Between Subthreshold Ion Channels in

Figure 5. Inhibition of IA by 4-AP is both concentration and voltage-dependent. A-

Representative trace of isolated IA before and after 5 mM 4-AP, at a command potential

of -10 mV. B- Summary data showing partial block of IA with 1 and 5 mM 4-AP (n=

17). Note the larger inhibition induced by the larger 4-AP concentration. C- The mean

current amplitude of IA in control ACSF and in the presence of 5 mM 4-AP was plotted

versus the command potential (n= 3). Data was corrected for LJP. #P<0.01 •P<0.05 and

*P<0.0001vs. control ACSF.

42

Page 57: Functional Interplay Between Subthreshold Ion Channels in

43

Page 58: Functional Interplay Between Subthreshold Ion Channels in

supports a concentration-dependent sensitivity of IA to 4-AP in PVN-RVLM neurons, as

previously described in other neuronal types [18,179].

In a subset of neurons (n= 3), the effect of 5 mM 4-AP was tested at a wide range

of command steps (-70 mV to +25 mV) (Fig. 5C). As shown in the mean I-V plot

generated from these recordings, IA sensitivity to 4-AP was found to be voltage-

dependent (F= 63.3; n= 3; P< 0.0001, 2-way ANOVA), with larger inhibition observed at

more depolarized membrane potentials.

3.1.6. IA shapes the Na+ action potential waveform in PVN-RVLM neurons

To determine whether IA modulates action potential waveform in PVN-RVLM

neurons, recordings were obtained in the current-clamp mode (n= 15). Action potential

amplitudes in all recorded neurons were ≥ +50 mV. Individual action potentials were

evoked using short (5 ms) depolarizing pulses, while clamping the neurons at two

different membrane potentials (-80 mV and -50 mV, see Methods), in order to obtain

different degrees of IA inactivation. The effects of 4-AP (5 mM) on various action

potential parameters were then determined. Results are summarized in Fig. 6.

At a holding potential of -80 mV, evoked action potential half width, and 90-10%

decay time were 2.0 ± 0.1 ms and 1.9 ± 0.2 ms, respectively. Bath application of 4-AP at

this holding membrane potential prolonged spike duration by 94.2 ± 16.7% (P< 0.0001 vs

control ACSF), and slowed down its decaying phase by 166.7 ± 27.5% (P< 0.0001, vs

control ACSF) (see Fig. 6).

Interestingly, similar changes in action potential waveform to those induced by 4-

AP were observed when neurons were clamped at a more depolarized Vm. Thus, at a

44

Page 59: Functional Interplay Between Subthreshold Ion Channels in

Figure 6. Effects of 4-AP on evoked Na+ action potential waveforms. A-

Representative examples of evoked (500 pA, 5 ms pulse) single action potentials before

and after 5 mM 4-AP, at holding potentials of -80 (left) or -50 mV (right). B- Summary

data showing the effects of 4-AP on action potential width and decay time at these two

holding potentials (n= 15). Note that 4-AP significantly prolonged spike width and decay

times, effects that were significantly larger when neurons were held at -80 mV.

***P<0.0001 vs. control within same holding potential group. #P< 0.05, ##P< 0.01 and

###P< 0.0005 compared to same treatment at -80 mV.

45

Page 60: Functional Interplay Between Subthreshold Ion Channels in

46

Page 61: Functional Interplay Between Subthreshold Ion Channels in

holding potential of -50 mV, spikes were broader (half width: 3.2 ± 0.2 ms; P< 0.0005,

vs, -80 mV), and displayed slower 90-10% decay times (2.6 ± 0.3 ms; P< 0.01, vs -80

mV). These changes were likely due to a higher degree of IA inactivation at the more

depolarized Vm.

At this depolarized Vm (-50 mV), 4-AP still induced similar changes in action

potential waveform as those observed when neurons were clamped at -80 mV. However,

these effects were significantly reduced. Thus, at a holding potential of -50 mV, 4-AP

prolonged spike duration by 54.5 ± 12.4% (P< 0.05, vs. % change in 4-AP at -80 mV)

and slowed down the decaying phase of the action potential by 95.1 ± 16.6% (P< 0.05,

vs. % change in 4-AP at -80 mV). Results are summarized in Fig. 6. Altogether, these

results suggest that the Na+ action potential waveform in PVN-RVLM neurons is

regulated by IA, an effect found to be likely dependent on its voltage-dependent

availability.

3.1.7. IA differentially regulates repetitive firing activity of PVN-RVLM neurons

The effects of 4-AP on firing activity were studied in 21 PVN-RVLM neurons.

About 62% (13 out of 21) of recorded neurons were spontaneously active. In the

remainder, firing activity was induced by injecting depolarizing DC current (+0.6 pA -

+36.7 pA). In the majority of recorded cells (~62%, 13/21), 5 mM 4-AP resulted in an

increased firing discharge. In the remainder (~38%, 8/21), a diminished firing discharge

was observed.

Interestingly, PVN-RVLM neurons that were differentially affected by 4-AP also

differed in some basic intrinsic properties. Neurons whose firing activity was enhanced

by 4-AP had a more hyperpolarized resting Vm (-51.7 ± 1.7 mV vs. -42.8 ± 2.2 mV, P<

47

Page 62: Functional Interplay Between Subthreshold Ion Channels in

0.005) and displayed a lower incidence of spontaneous activity than those inhibited by 4-

AP (38.5% vs. 100%, respectively, P< 0.01, Fisher’s exact test). No differences in input

resistance between the two groups were observed (842.6 ± 154.2 MΩ vs. 896.8 ± 157.4

MΩ in 13 and 8 neurons, respectively, P> 0.5). Furthermore, as summarized in Table 2,

significant differences in the Na+ action potential waveform were observed between the

two groups of PVN-RVLM neurons. For example, neurons whose firing activity was

increased by 4-AP displayed narrower (32% P< 0.0005) and faster (44% P< 0.001)

decaying action potentials than those inhibited by 4-AP.

For simplistic purposes, further results obtained from these two differently

responsive PVN-RVLM neurons are presented below in separate sections.

3.1.8. IA restrains firing activity in a subset of PVN-RVLM neurons

Among the neurons whose firing activity were enhanced by 4-AP, ~69% (9/13)

and ~31% (4/13) displayed continuous or bursting firing patterns, respectively. The

relative low incidence of the latter group precluded us from obtaining a detailed analysis

of the effects of 4-AP on bursting properties. Thus, all these neurons were pooled and the

effect of 4-AP on their mean firing discharge was analyzed (see Methods). Results are

summarized in Fig. 7. Within this group, 4-AP resulted in ~ 90% increase in firing rate

(control: 1.0 ± 0.2 Hz; 4-AP: 1.9 ± 0.4 Hz; n= 13; Fig. 7A).

To determine whether IA also modulates action potential waveform during

repetitive firing activity, action potentials recorded in periods before and during 4-AP

application were analyzed (Methods). Similarly to the effects observed on single evoked

spikes (see above), bath application of 4-AP prolonged action potential duration (~70%)

and 90-10% decay time (~93%), and slightly increased action potential amplitude (~9%).

48

Page 63: Functional Interplay Between Subthreshold Ion Channels in

Action Potential Waveform Parameters

4-AP Stimulated Group 4-AP Inhibited Group

ACSF 5mM 4-AP ACSF 5mM 4-AP

Peak Amplitude

(mV)

64.6 ± 1.4 70.3 ± 1.3 * 61.1 ± 2.7 67.8 ± 3.1 *

Half-width (ms) 3.0 ± 0.09 5.1 ± 0.3 * 4.1 ± 0.2 # 7.5 ± 0.6 *, #

90-10% Decay

Time (ms)

1.5 ± 0.06 2.9 ± 0.2 * 2.0 ± 0.08 # 5.2 ± 0.6 *, #

Threshold (mV) -31.2 ± 0.8 -32.8 ± 0.8 * -31.0 ± 0.9 -31.8 ± 1.7

Table 2. The effects of 4-AP upon action potential waveform parameters. * P< 0.05

within same group; #P< 0.05 between different groups (same drug treatment).

49

Page 64: Functional Interplay Between Subthreshold Ion Channels in

Figure 7. 4-AP increases firing discharge in a subset of PVN-RVLM neurons. A-

Representative examples of firing discharge in a continuously (upper trace) and bursting

(lower trace) firing neurons before (left), during (middle) and after (right) bath

application of 5 mM 4-AP. As summarized in the bar graphs, the firing frequency was

significantly increased by 4-AP (n= 13). B- Representative examples of averaged

spontaneous action potential waveform before (thin line) and after (thick line) application

of 5 mM 4-AP, obtained from a neuron whose firing activity was increased by 4-AP.

Inset, the HAP peaks were normalized and overlapped in order to more clearly depict the

effects of 4-AP on the HAP slope. The mean effects of 4-AP on HAP properties are

summarized in the bar graphs. *P<0.05 vs control. C- Another example of a 4-AP-

induced increment in firing discharge in a PVN-RVLM neuron. In this case, negative

current was injected in the presence of 4-AP, in order to diminish firing rate back to

control levels. A plot of firing frequency as a function of time (10 s binning) is shown in

the middle panel. In the right panel, a plot of action potential half-width as a function of

time (10 s binning) is shown. Note that in the presence of 4-AP, the width of the action

potentials were still prolonged, even when firing rate was similar to control levels.

50

Page 65: Functional Interplay Between Subthreshold Ion Channels in

51

Page 66: Functional Interplay Between Subthreshold Ion Channels in

Importantly, action potential threshold was shifted (~ 2 mV) to a more hyperpolarized

membrane potential in the presence of 4-AP. Results are summarized in Table 2.

In order to rule out that changes in action potential waveform induced by 4-AP

were secondary to membrane depolarization and increased firing rate per se, we

performed a subset of recordings (n= 3) in which the membrane potential in the presence

of 4-AP was hyperpolarized with DC current injection, in order to bring the firing

frequency of the recorded cell back to control levels. As shown in Fig. 7C, the action

potential half-width was still prolonged in the presence of 4-AP (92.8 ± 7.1% increase in

half width, compared to control , P< 0.05, paired t-test), even when the firing rate was

decreased near to control levels.

In addition, we analyzed the effects of 4-AP on hyperpolarizing afterpotentials

(HAPs) during repetitive firing (Fig. 7B). Neither the HAP peak amplitude (measured as

the difference from threshold to peak) nor its area were affected by 4-AP (P> 0.05 in both

cases). Nonetheless, the absolute HAP peak potential reached a significantly more

hyperpolarized membrane potential in the presence of 4-AP (control: -55.5 ± 0.9 mV; 4-

AP: -58.2 ± 1.2 mV; n= 13, P< 0.05, paired t-test), likely due to the hyperpolarizing shift

in spike threshold. Finally, the HAP decay time course was found to be steeper (~ 25%,

P< 0.05) in the presence of 4-AP.

Multiple correlation analysis within this subset of PVN-RVLM neurons failed to

identify any significant correlation between the degree of 4-AP-induced changes in firing

activity, intrinsic membrane properties and/or action potential properties (R2 values:

0.009 – 0.1).

3.1.9. IA contributes to ongoing firing activity in a subset of PVN-RVLM neurons

52

Page 67: Functional Interplay Between Subthreshold Ion Channels in

Among the neurons whose firing activity were inhibited by 4-AP, 75% (6/8) fired

in continuous mode, while the rest (2/8) displayed a bursting pattern. In 1 case, 4-AP

switched the firing pattern from continuous to bursting mode (Fig. 8A). On average, 4-

AP decreased the firing rate of this subgroup of PVN-RVLM neurons by ~ 40% (control

ACSF: 2.3 ± 0.5 Hz; 4-AP: 1.4 ± 0.3 Hz; Fig. 8A).

Similarly to the other subset of PVN-RVLM neurons, bath application of 4-AP in

the subset of PVN-RVLM neurons inhibited by 4-AP, increased the amplitude of the

action potential (~ 10%), and prolonged its duration (~ 80%) and 90-10% decay time

(160%). On the other hand, action potential threshold was not affected by 4-AP (results

are summarized in Table 2).

In a subset of recordings (n= 4) the membrane potential in the presence of 4-AP

was depolarized with DC current injection, in order to bring the firing frequency of the

recorded neurons back to control levels. In these cases, action potential width was still

prolonged by 4-AP (78.7 ± 25.8% increase in half width, compared to control, P< 0.05

paired t-test).

Interestingly, opposing effects on various HAP parameters within the subset of 4-

AP inhibited PVN-RVLM neurons were observed when compared to the 4-AP-enhanced

group. For example, 4-AP slowed down the HAP decay slope by ~ 24% (P< 0.05), and

despite a slight decrease in HAP peak amplitude (~ 14%, P< 0.05), the overall HAP area

was increased by ~ 27% (P< 0.01), likely due to the slower HAP decay time course. No

differences in the absolute HAP peak potential were observed between control ACSF and

5 mM 4-AP (control: -53.4 ± 1.3 mV; 4-AP: -51.0 ± 3.0 mV; n= 8, P= 0.2). Results are

summarized in Fig. 8B.

53

Page 68: Functional Interplay Between Subthreshold Ion Channels in

Figure 8. 4-AP dimished firing discharge in a subset of PVN-RVLM neurons. A-

Representative examples of firing discharge in a continuously (upper trace) and bursting

(lower trace) firing neurons before (left) and during (right) bath application of 5 mM 4-

AP. As summarized in the bar graphs, the firing frequency was significantly diminished

by 4-AP (n= 8). B- Representative examples of averaged spontaneous action potential

waveforms before (thin line) and after (thick line) application of 5 mM 4-AP, obtained

from a neuron whose firing discharge was diminished by 4-AP. Inset, the HAP peaks

were normalized and overlapped in order to more clearly depict the effects of 4-AP on

the HAP slope. The mean effects of 4-AP on HAP properties are summarized in the bar

graphs. *P< 0.05 and **P< 0.01 vs control.

54

Page 69: Functional Interplay Between Subthreshold Ion Channels in

55

Page 70: Functional Interplay Between Subthreshold Ion Channels in

In this subset of neurons, a significant correlation between percent changes

induced by 4-AP on firing activity and HAP decay slope was observed (R2: 0.6, P< 0.02).

No other significant correlations between 4-AP-induced changes in firing activity,

intrinsic membrane properties and/or action potential properties were observed (R2

values: 0.009 – 0.3).

3.1.10. Effects of TEA on action potential waveform and firing activity in PVN-RVLM

neurons

In addition to preferentially blocking IA, 4-AP at low millimolar concentrations

may also partially block IKDR [133,166]. In an attempt to further explore this possibility,

we tested the effects of TEA, which preferentially blocks IKDR over IA [126,143], on

action potential waveform and firing activity in PVN-RVLM neurons. Bath application of

30 mM TEA robustly diminished IKDR (94.9 ± 2.2%; n= 8) while only slightly inhibiting

IA (17.6 ± 3.5%; n= 8, P< 0.0001, unpaired t-test; Fig. 9A). Differently from 4-AP, long

lasting plateau potentials were observed in the presence of TEA, resulting in all cases in a

robust inhibition of firing discharge (control: 6.0 ± 1.2 Hz; TEA: 1.0 ± 0.2 Hz; n= 4, P<

0.05 (paired t-test; Fig. 9C). Since TEA per se induced robust changes in action potential

waveform and firing activity, testing the effects of 4-AP in the presence of TEA in

current clamp recordings was not feasible. Thus, while we cannot completely rule out

that 4-AP effects on firing activity are in part due to blockade of IKDR, the disparity in the

effects observed between 4-AP and TEA treatments, along with the voltage-dependency

of 4-AP effects on some of the measured parameters (e.g., duration of action potential)

would suggest that these compounds act through different mechanisms.

56

Page 71: Functional Interplay Between Subthreshold Ion Channels in

Figure 9. Effects of TEA on outward K+ currents and action potential waveform in

PVN-RVLM neurons. A- Representative traces of outward currents evoked by a

depolarizing step to -10 mV, from a conditioning step to -100 mV (500 ms), before and

during bath application of 30 mM TEA. Note that TEA completely abolished the

sustained outward component, while having very little effect on the transient A-type K+

component (n= 8). B- Representative examples of spontaneous action potentials before

and during bath application of 30 mM TEA. Note that TEA induced a long lasting

plateau potential on the spontaneous spike. C- Summary data showing a significant

decrease in the firing frequency of PVN-RVLM neurons in the presence of 30 mM TEA

(n= 4). *P<0.05 vs control.

57

Page 72: Functional Interplay Between Subthreshold Ion Channels in

58

Page 73: Functional Interplay Between Subthreshold Ion Channels in

3.1.11. The effects of 4-AP on action potential waveform and firing activity in PVN-

RVLM neurons are Ca2+-dependent

By prolonging action potentials, blockade of IA could result in an enhanced Ca2+

entry per spike [31,89], leading in turn to activation of various Ca2+ dependent

mechanisms, such as Ca2+-dependent K+ channels [167]. Depending on the complement

of voltage-gated Ca2+, and Ca2+-dependent conductances available within each particular

neuronal type, 4-AP-induced changes in Ca2+ entry could either increase or decrease

membrane excitability. Thus, to determine to what extent 4-AP-induced changes in firing

discharge in PVN-RVLM neurons were Ca2+-dependent, experiments were repeated in

the presence of the broad spectrum Ca2+ channel blocker Cd2+ (200 μM; n= 12). Results

are summarized in Fig. 10. In the presence of Cd2+, after-hyperpolarizing potentials

(AHPs) following trains of spikes, a well-characterized Ca2+-dependent property

[20,91,187], were blocked, supporting the efficacy of Cd2+ to block Ca2+-dependent

membrane properties in these neurons (Fig. 10A).

In the presence of Cd2+, 4-AP was still able to prolong the duration of the action

potential. However, the magnitude of this effect was smaller (though the difference did

not reach statistical significance) as compared to control conditions (4-AP-ACSF= 77.2 ±

8.7%, 4-AP-Cd2+= 55.5 ± 14.1%, P= 0.09). These results indicate that 4-AP-induced

spike broadening likely results from both the slower action potential decaying phase, as

well as increased Ca2+ influx during the depolarizing phase of the action potential

prolonging its duration, as previously shown in magnocellular neurosecretory neurons

[20]. In fact, the 90-10% decay time course of the action potential was still significantly

slowed down by 4-AP (P< 0.05) in the presence of Cd2+, to a similar extent as that

59

Page 74: Functional Interplay Between Subthreshold Ion Channels in

Figure 10. 4-AP effects on PVN-RVLM firing discharge are Ca2+-dependent. A-

The afterhyperpolarizing potential (AHP), a Ca2+-dependent process, is blocked by bath

application of 200 µM Cd2+. B- In the presence of 200 µM Cd2+, 4-AP failed to affect

the firing discharge of PVN-RVLM neurons. Representative traces and the summary data

are shown in the upper and lower panels, respectively (n= 12).

60

Page 75: Functional Interplay Between Subthreshold Ion Channels in

61

Page 76: Functional Interplay Between Subthreshold Ion Channels in

observed in control ACSF (4-AP-ACSF= 115.5 ± 15.2%, 4-AP-Cd2+= 97.7 ± 26.7%, P=

0.3).

Importantly, 4-AP failed to affect the firing activity of all recorded neurons in

slices pre-incubated (~ 5 min.) with Cd2+ (control: 1.2 ± 0.2 Hz; 4-AP: 1.4 ± 0.2 Hz; n=

12, P= 0.3, paired t-test; Fig. 10B). Furthermore, 4-AP failed to affect the action

potential peak amplitude (p= 0.3), threshold (p= 0.4), HAP area (p= 0.3), or HAP slope

(p= 0.1) in the presence of Cd2+.

3.1.12. Possible potassium channel subunits underlying A-type potassium currents in

PVN-RVLM neurons

To gain insights into the possible Kv subunits underlying IA in PVN-RVLM

neurons, we combined immunohistochemical identification of Kv1.4, 4.2 and 4.3

subunits, (known to underlie A-type K+ currents in other CNS neurons, [37]) with

neuronal tracing techniques (see Methods). Representative confocal photomicrographs of

Kv1.4, 4.2 and 4.3 immunoreactivities in retrogradely-labeled PVN-RVLM neurons are

shown in Fig. 11. Kv1.4 and 4.3 subunits were found to be widely expressed and

distributed within the PVN (Fig. 11A). On the other hand, Kv4.2 immunoreactivity was

very weak and/or undetectable in the PVN. Conversely, a robust Kv4.2 immunoreactivity

was observed in ependymal cells lining the third ventricle, and in glial-like processes in

the median eminence (Fig. 11A2), supporting the efficacy of our approach to detect this

immunoreactivity. All Kv immunoreactive reactions were blocked following pre-

adsorption of antibodies with their respective peptides (results not shown).

Immunoreactivities for these Kv subunits were sampled in a total of 68

retrogradely-labeled PVN-RVLM neurons at higher magnification. Strong

62

Page 77: Functional Interplay Between Subthreshold Ion Channels in

Figure 11. Kv1.4, 4.2 and 4.3 immunoreactivity in retrogradely labeled PVN-

RVLM neurons. A- Low magnification confocal images of the PVN (10X) showing

Kv1.4 (A1), Kv4.2 (A2) and Kv4.3 (A3) immunoreactivities. The inset in A2 shows

Kv4.2 immunoreactivity in the median eminence. B- High magnification confocal

images of retrogradely labeled PVN-RVLM neurons (B1) and Kv 1.4 immunoreactivity

(B2). Both images were superimposed in B3 to better depict colocalization of the two

signals. Arrowheads in B3 point to representative Kv1.4 immunoreactive clusters located

at the surface of the retrogradely labeled neuron. A magnified image is shown in the

inset. C- High magnification confocal images of retrogradely labeled PVN-RVLM

neurons (C1) and Kv 4.2 immunoreactivity (C2). Both images were superimposed in C3.

Note the lack of Kv4.2 immunoreactivity in the neurons displayed. D- High

magnification confocal images of retrogradely labeled PVN-RVLM neurons (D1) and Kv

4.3 immunoreactivity (D2). Both images were superimposed in C3 to better depict

colocalization of the two signals. Arrows in B-D point to the location of PVN-RVLM

retrogradely labeled neurones (n= 68).

63

Page 78: Functional Interplay Between Subthreshold Ion Channels in

64

Page 79: Functional Interplay Between Subthreshold Ion Channels in

immunoreactivities for Kv1.4 and Kv4.3 subunits were clearly present in PVN-RVLM

neurons (32/33 and 15/16, respectively), as well as in nearby, unlabeled neurons.

Representative examples are shown in Fig. 11B1-B3 (for Kv1.4) and Fig. 11D1-D3 (for

Kv4.3). Immunoreactivities for these subunits were highly punctate in nature, with

immunoreactive clusters located within the cytoplasm as well as near the surface

membrane, both in somatic and dendritic compartments. In the case of Kv4.3 subunits,

nuclear immunoreactivity was also evident. On the other hand, weak or undetectable

Kv4.2 immunoreactivity was observed in PVN-RVLM, or other PVN neurons (Fig.

11C1-C3).

3.2. Altered A-type Potassium Current and Firing Properties in Rat Sympathetic

Preautonomic PVN Neurons During Hypertension

3.2.1. Intrinsic properties of sham and hypertensive PVN-RVLM neurons

Whole-cell patch clamp recordings were obtained from retrogradely labeled PVN-

RVLM neurons in sham (n= 41) and hypertensive (n= 43) rats. The mean input resistance

was 1226.0 ± 170.8 MΩ and 1178.0 ± 109.4 MΩ for sham and hypertensive PVN-RVLM

neurons, respectively (P> 0.8). Neuronal cell capacitance was significantly reduced in

PVN-RVLM neurons during hypertension (sham: 27.92 ± 1.37 pF; hypertensive: 23.05 ±

1.31 pF, P= 0.01).

3.2.2. Changes in PVN-RVLM neuronal morphometry during hypertension

Cell capacitance is a general indicator of neuronal membrane surface area [82,127].

Thus, changes in cell capacitance during hypertension may be indicative of

somatodendritic structural changes during this condition. To determine if this was the

65

Page 80: Functional Interplay Between Subthreshold Ion Channels in

case, a subset of recorded neurons that were properly filled with biocytin (n= 24 and 27

in sham and hypertensive) were reconstructed in three-dimension to compare for

differences and/or changes in somatic/dendritic surface area among groups (Methods).

Results are summarized in Fig.12. Overall neuronal surface area was significantly

reduced in PVN-RVLM neurons in hypertensive rats (P< 0.05). While a tendency for

reduced somatic area was observed (P= 0.07), our results suggest that the reduced

neuronal size was mostly due to a significant reduction in dendritic surface area (P=

0.02).

Importantly, significant correlations were observed between cell capacitance and

somatic area, as well as cell capacitance and dendritic surface area (r= 0.6, P< 0.0001 in

both cases, Pearson’s correlation test), supporting cell capacitance measurements in our

studies as reliable indicators of neuronal surface area (see below).

3.2.3. IA voltage-dependent activation properties

In a first set of studies (voltage-clamp mode), a combination of pharmacological

and digital methods were used to isolate and compare the magnitude and voltage-

dependent properties of the A-type K+ current (IA) between the two experimental groups

(n= 27 and 20 in sham and hypertensive rats). Depolarizing steps of increasing

amplitudes (-70 mV to +25 mV, in 5 mV increments) were used to activate IA (Figure

13A), and I/V plots were generated and fit with a Boltzmann function (Fig. 13B-C). IA

amplitude/density, activation threshold, half-activation potential, and 10-90% rise time

were analyzed and compared between groups.

We first compared the IA activation curves. As shown in Fig. 13B, IA activation

was significantly reduced in PVN-RVLM neurons from hypertensive rats (F= 37.2, P<

66

Page 81: Functional Interplay Between Subthreshold Ion Channels in

Figure 12. Changes in PVN-RVLM neuronal morphometry during hypertension.

A- Representative examples of reconstructed PVN-RVLM neurons in sham and

hypertensive rats. Axons are identified with arrows. B- Summary data of overall

neuronal, somatic and dendritic surface areas (n= 24 and 27 sham and hypertensive

neurons, respectively). Note the reduction in the overall neuronal and dendritic surface

area in PVN-RVLM neurons during hypertension. *P< 0.05.

67

Page 82: Functional Interplay Between Subthreshold Ion Channels in

68

Page 83: Functional Interplay Between Subthreshold Ion Channels in

Figure 13. Changes in IA voltage-dependent activation properties during

hypertension. A- Representative example of IA voltage-dependent activation in PVN-

RVLM neurons obtained from a sham and a hypertensive rat. Note the smaller maximum

current amplitude in the latter. B- Mean plots of current amplitude versus the command

potential fit with Boltzmann functions. Squares and triangles represent sham and

hypertensive rats, respectively (n= 27 and 20, respectively). Note the significant

reduction of IA magnitude during hypertension. C- Mean plots of IA current density

versus the command potential fit with Boltzmann functions. Note the lack in reduction of

IA current density in PVN-RVLM neurons during hypertension. D- Summary data of IA

activation threshold and half-activation potential. No significant differences were

observed among the experimental groups. *P< 0.05, **P< 0.01.

69

Page 84: Functional Interplay Between Subthreshold Ion Channels in

70

Page 85: Functional Interplay Between Subthreshold Ion Channels in

0.0001, 2 way ANOVA). Based on the differences found in cell size/morphometry

between groups (see above), we addressed whether changes in current amplitude during

hypertension persisted after being normalized by cell size (i.e., current density, Fig. 13C).

Our results indicate no differences in PVN-RVLM IA current density between sham and

hypertensive rats (F= 1.4, P> 0.2, 2 way ANOVA). No differences in IA activation

threshold or in the half-activation potential were observed between sham and

hypertensive rats (P> 0.5 in both cases) (Fig. 13D). Similarly, no differences in the slope

factor (k) of the voltage-dependent activation curve, which describes the steepness of the

activation curve (voltage sensitivity), were observed between groups (P> 0.5). Finally, no

differences in the rate of activation of IA, calculated as the 10-90% rise time following a

command potential to -10 mV, were observed between the two groups (P= 0.9).

To determine whether changes in IA properties during hypertension were specific

to PVN-RVLM neurons, or alternatively, whether they affected other preautonomic cell

populations, we obtained recordings from identified PVN neurons that innervate the

dorsal vagal complex (DVC), as a control group. PVN-DVC neurons comprise a

heterogeneous population that innervates both the nucleus of the solitarii tract (NTS), as

well as the dorsal motor nucleus of the vagus [55,72,98,178]. Results are summarized in

Fig. 14. Differently from PVN-RVLM neurons, no structural morphometric changes

were observed during hypertension in PVN-DVC neurons (Fig. 14B). Nonetheless, and

similar to PVN-RVLM neurons, IA activation was significantly reduced in PVN-DVC

neurons from hypertensive rats (F= 87.6, P< 0.0001, 2 way ANOVA). Since no

differences in cell size were observed between sham and hypertensive rats, IA current

71

Page 86: Functional Interplay Between Subthreshold Ion Channels in

Figure 14. Morphometric and state-dependent differences in IA voltage-dependent

activation properties of PVN-DVC neurons. A- Representative examples of

reconstructed PVN-DVC neurons in sham and hypertensive rats. Axons are identified

with arrows. B- Summary data of somatic and dendritic surface area. Note the lack of

changes during hypertension. C- Mean plots of the current amplitude and current density

versus the command potential fit with Boltzmann functions. Squares and triangles

represent sham and hypertensive rats, respectively (n= 18 and 14, respectively). Note the

significant reduction of IA amplitude and current density during hypertension. *P< 0.05,

**P< 0.01, ***P<0.0001.

72

Page 87: Functional Interplay Between Subthreshold Ion Channels in

73

Page 88: Functional Interplay Between Subthreshold Ion Channels in

density was also diminished in PVN-DVC neurons during hypertension (F= 61.0, P<

0.0001, 2 way ANOVA). Due to the more homogenous nature, and more established

relevance in cardiovascular control, the rest of our studies focused on PVN-RVLM

neurons.

3.2.4. IA voltage-dependent inactivation properties

The voltage-dependent and kinetic properties of inactivation of IA were studied in

25 sham and 18 hypertensive PVN-RVLM neurons. Results are summarized in Fig.15.

Overall, IA steady state inactivation was found to be significantly increased in PVN-

RVLM neurons from hypertensive rats (F= 37.8, P< 0.0001, 2 way ANOVA). Moreover,

the mean half-inactivation potentials (V1/2inact, i.e., the Vm at which 50% of the maximum

current is inactivated) were significantly shifted to a more hyperpolarized membrane

potential (~ 6 mV) in PVN-RVLM neurons from hypertensive rats (P< 0.05). The slope

factor (k) of the inactivation curve was slightly though significantly increased in PVN-

RVLM neurons during hypertension (sham: 12.1 ± 0.4 mV; hypertensive: 14.5 ± 0.6 mV

(P< 0.01). Finally, no differences in IA rate of inactivation (τ) were observed between

groups (P>0.5).

3.2.5. Recovery of IA from inactivation

The rate of IA recovery from inactivation was determined in 20 sham and 12

hypertensive PVN-RVLM neurons. A progressively larger amount of IA activation was

achieved following hyperpolarizing pulses of increasing durations (-100 mV, 10-250 ms,

in 10 ms increments), followed by a command potential to -10 mV (Fig. 16A), supporting

hyperpolarization-dependent removal of IA inactivation. Plots of the normalized IA peak

74

Page 89: Functional Interplay Between Subthreshold Ion Channels in

Figure 15. Changes in IA voltage-dependent inactivation properties in PVN-RVLM

neurons during hypertension. A- Representative traces of the voltage dependence of

inactivation of IA in sham and hypertensive rats. The gray traces (arrow) represent the

current evoked at the half-inactivation potential (values listed under the traces). B- Plots

of mean normalized current amplitude versus the conditioning potential fit by Boltzmann

functions. Note the hyperpolarizing shift in PVN-RVLM neurons, during hypertension.

Squares and triangles represent sham and hypertensive rats, respectively (n= 25 and 18,

respectively). C- Summary data of IA half-inactivation potential. *P< 0.05.

75

Page 90: Functional Interplay Between Subthreshold Ion Channels in

76

Page 91: Functional Interplay Between Subthreshold Ion Channels in

Figure 16. Time course of recovery from inactivation of IA in PVN-RVLM neurons

in sham and hypertensive rats. A- Representative traces of IA currents evoked from a

command pulse of -10 mV with a conditioning pulse (-100 mV) of increasing duration

(10-250 ms, Δ10 ms increments). Lower panels depict the normalized current amplitude

evoked at the command test, plotted against the conditioning step duration. A single

exponential function was fit to the plot, and the time course of recovery from inactivation

(τ) of IA was calculated. The time constant (τ) of recovery from inactivation is listed

below each trace. B- Summary data of mean τ of recovery from inactivation (n= 20 sham

and 12 hypertensive neurons). No significant differences were observed between groups.

77

Page 92: Functional Interplay Between Subthreshold Ion Channels in

78

Page 93: Functional Interplay Between Subthreshold Ion Channels in

amplitude as a function of the duration of the hyperpolarizing conditioning steps were

built, and fit with a monoexponential function (Fig. 16B), to determine the τ of recovery

from inactivation. No significant differences in the mean τ of recovery from inactivation

were observed between PVN-RVLM neurons in sham and hypertensive rats (P> 0.5).

3.2.6. Non-stationary fluctuation analysis of IA in PVN-RVLM neurons: Changes during

hypertension

To determine whether changes in single channel properties contributed to the

diminished IA current observed during hypertension (see above), we used non-stationary

fluctuation analysis, a method originally described by Sigworth [176,177]. This approach

enables the extraction of single-channel information from macroscopic currents,

including unitary current, open probability and number of channels. IA evoked (n= 100-

130 traces) with a command pulse to +40 mV were obtained. The mean evoked current

was scaled to individual IA waveforms, and the variance of the individual IA current

around the scaled average was computed. Variance-amplitude relationships were plotted

and fit with a parabolic function (Methods). A representative example is shown in Fig.

17. In most cases, experimental points over the right side of the parabola were missing,

indicative that channel open probability never reached a value of 1.0, likely due to

channel inactivation [5,176,177]. Under this condition, we were limited to estimate the

unitary current (i), determined by the slope of the parabola, shown to be highly accurate

even at limited levels of open probability. On the other hand, the number of channels can

only be accurately estimated at an open channel probability near 1 [5,176,177]. As shown

in Figure 17C, our results indicate a significantly diminished IA single channel

conductance in PVN-RVLM neurons during hypertension (P< 0.05).

79

Page 94: Functional Interplay Between Subthreshold Ion Channels in

Figure 17. Non-stationary noise analysis of IA in PVN-RVLM neurons of sham and

hypertensive rats. A- Representative traces of IA evoked with a command potential to

+40 mV. The grey trace is the mean evoked current. B- Plot of the variance from the

mean trace as a function of the current amplitude, fit with a parabolic function. C-

Summary data of mean single channel conductance in sham and hypertensive rats (n= 12

and 9, respectively). Note the significantly reduced single channel conductance in the

hypertensive neurons. *P<0.05.

80

Page 95: Functional Interplay Between Subthreshold Ion Channels in

81

Page 96: Functional Interplay Between Subthreshold Ion Channels in

3.2.7. The Na+ action potential waveform is affected in PVN-RVLM neurons during

hypertension

Abundant evidence supports a major role for IA in the regulation of single and

repetitive neuronal firing behavior [39,105,166,173]. Thus, we evaluated whether action

potential properties known to be regulated by IA are altered in PVN-RVLM neurons

during hypertension. In a first series of studies, we evaluated differences in the single

action potential waveform. Individual action potentials were evoked using short (5 ms)

depolarizing pulses, while injecting current to maintain Vm at approximately two

different membrane potentials (~ -80 mV and ~ -50 mV, see Methods), in order to obtain

different degrees of IA inactivation. In PVN-RVLM neurons from sham rats, action

potentials evoked from a more depolarized Vm (~ -50 mV) were wider (by ~ 177.5 ±

16.3%, P< 0.001) and displayed a slower decay time course when compared to spikes

evoked at a Vm of ~ -80 mV (by 160.0 ± 23.3%, P< 0.05 for width and decay time,

respectively). These differences were likely due to a higher degree of IA inactivation at

the more depolarized Vm, as we recently reported [180]. At a membrane potential of ~ -

80 mV, action potentials in PVN-RVLM neurons from hypertensive rats were wider and

decayed slower than those evoked in sham rats at a similar Vm (P< 0.05 in both cases).

Spike width in hypertensive rats was further increased at a Vm ~ -50 mV (151.3 ± 11.0%,

P< 0.05), though to a lesser extent than that observed in sham rats (see above). Moreover,

the decay time course at ~ -50 mV was not significantly different than that observed at ~ -

80 mV (P> 0.5). Results are summarized in Fig. 18A.

As we previously reported [180], 4-AP, a K+ channel blocker that more

selectively blocks IA over other voltage-dependent K+ channels (e.g., IKDR) prolonged the

82

Page 97: Functional Interplay Between Subthreshold Ion Channels in

Figure 18. Changes in Na+ action potential waveform of PVN-RVLM neurons

during hypertension. A1- Representative traces of evoked action potentials at a

membrane potential ~ -80 mV (black trace) and ~ -50 mV (grey trace) in a PVN-RVLM

neuron from a sham rat. Action potentials were aligned to better depict differences in

their waveforms. A2- Summary data of evoked action potential half-widths and decay-

times at membrane potentials of ~ -80 mV and ~ -50 mV. *P<0.05, **P< 0.01. B1-

Representative traces of spontaneous firing activity in PVN-RVLM neurons from a sham

and a hypertensive rat. The averaged spontaneous action potential from the sham (black

trace) and the hypertensive (grey) rat were superimposed and shown in the right panel.

B2- Summary data of basal spontaneous firing activity, half-width, decay-time, and HAP

decay slope from the averaged spontaneous single spikes (n= 9 sham and 12 hypertensive

neurons). *P<0.05, **P<0.01.

83

Page 98: Functional Interplay Between Subthreshold Ion Channels in

84

Page 99: Functional Interplay Between Subthreshold Ion Channels in

action potential width and slowed down the decay time course in PVN-RVLM neurons,

effects that were significantly diminished in hypertensive rats (spike width: sham: 243.4

± 17.2%; hypertensive: 176.4 ± 16.9%; decay time: sham: 285.0 ± 27.4%; hypertensive:

190.9 ± 24.7 % , P< 0.05 in both cases, paired-t test).

To determine whether the action potential waveform during repetitive firing

activity in PVN-RVLM neurons also was altered in hypertensive rats, action potentials

from spontaneously active neurons (n=9 and 12 in sham and hypertensive rats,

respectively) were grouped and compared between experimental groups (Methods).

Representative examples are shown in Fig. 18B. Spontaneous firing activity was

significantly higher in PVN-RVLM neurons from hypertensive rats (P< 0.01). Similarly

to the differences observed on single evoked spikes, spontaneous action potentials in

hypertensive rats were wider (P< 0.05) and displayed a slower decay time course (P<

0.005) than those recorded from sham rats. Moreover, the slope of the hyperpolarizing

afterpotential (HAP) following each spike was significantly steeper in PVN-RVLM

neurons from hypertensive rats (P< 0.01). Altogether, these data support a diminished IA

influence in the regulation of the action potential waveform in PVN-RVLM neurons

during hypertension.

3.2.8. Action potential-dependent increase in intracellular Ca2+ is enhanced in PVN-

RVLM neurons during hypertension

Since an increase in spike duration is known to be associated with an increase in

intracellular calcium levels [106], we combined simultaneous patch clamp recordings and

confocal Ca2+ imaging in order to determine if intracellular calcium levels ([Ca2+]ic) were

altered during hypertension. Retrogradely-labeled PVN-RVLM neurons were

85

Page 100: Functional Interplay Between Subthreshold Ion Channels in

intracellularly filled with Fluo-5F (100 μM), and relative changes in [Ca2+]ic in response

to single evoked action potentials were measured (Methods, Fig. 19A,B), and between 8-

10 traces of calcium transients were averaged together in order to increase the signal to

noise ratio. The peak of the [Ca2+]ic transient evoked from a single spike was measured

before and after 5 mM 4-AP application in PVN-RVLM neurons from sham (n=9) and

hypertensive (n=11) rats. Results are summarized in Fig. 19C. Action potential-evoked

[Ca2+]ic transients in PVN-RVLM neurons from hypertensive rats were significantly

larger to those evoked in sham rats (P< 0.05). Moreover, while 4-AP significantly

enhanced the Ca2+ transient in the sham group (78.1 ± 19.3%, P<0.05), it failed to do so

in the hypertensive group (P>0.05) (Fig 19C).

3.2.9. Spike broadening during repetitive firing is enhanced in PVN-RVLM neurons

during hypertension

Repetitive firing activity of PVN neurons, including identified preautonomic

neurons, is characterized by a progressive increment in action potential duration (spike

broadening) [10,185], a phenomenon shown to be dependent on progressive increment in

[Ca2+]ic and steady-state inactivation of IA [88]. Thus, we explored whether spike

broadening in PVN-RVLM neurons was also affected during hypertension. Repetitive

firing was evoked with depolarizing pulses (110-130 pA, 180 ms), and the spike width of

the evoked action potentials was measured and compared between groups. The degree of

spike broadening was quantified as the ratio of the duration of the third spike to that of

the first spike during the train [175,185]. Representative examples and summary data are

shown in Fig. 20. While spike broadening was observed in both groups, the degree of

86

Page 101: Functional Interplay Between Subthreshold Ion Channels in

Figure 19. Changes in action potential-evoked [Ca2+]ic levels in PVN-RVLM

neurons during hypertension. A- Representative traces of an evoked action potential

(grey) and the resultant change in [Ca2+]ic levels (black) in a sham rat. B- Representative

confocal images showing a Fluo-5 loaded PVN-RVLM neuron (left panel), as well as

basal and peak action potential-evoked [Ca2+]ic levels (pseudocolor images) in the

absence and presence of 5 mM 4-AP (scale units are F/F0). Note the increase in the

amplitude of [Ca2+]ic levels upon 4-AP application. The lower traces show the time

course of changes in [Ca2+]ic in control (black trace) and in the presence of 4-AP (grey

trace). C- Summary data of the peak of the [Ca2+]ic transient, before and after 4-AP, in

sham and hypertensive neurons (n= 9 and 11, respectively). *P<0.05.

87

Page 102: Functional Interplay Between Subthreshold Ion Channels in

88

Page 103: Functional Interplay Between Subthreshold Ion Channels in

Figure 20. Changes in spike broadening in PVN-RVLM neurons during

hypertension. A- Representative trace of an evoked burst of spikes (110 pA, 180 ms) in

a PVN-RVLM neuron from a sham rat (left panel). In the right panel, the evoked spikes

were scaled and superimposed to better depict the progressive increase in width. B-

Mean ratio of the half-width of the third spike to that of the first spike. Note the

significantly increased ratio in the hypertensive group (n= 22 and 15 in sham and

hypertensive rats, respectively). *P<0.05.

89

Page 104: Functional Interplay Between Subthreshold Ion Channels in

90

Page 105: Functional Interplay Between Subthreshold Ion Channels in

broadening was significantly larger in PVN-RVLM neurons from hypertensive rats (P<

0.05).

3.3. Contribution of Subthreshold Ion Channels, to Hypertension, in Preautonomic

PVN Neurons

3.3.1. Intrinsic Properties of sham and hypertensive PVN-RVLM neurons

Whole-cell patch clamp recordings were obtained from retrogradely labeled PVN-

RVLM neurons in sham (n= 80) and hypertensive (n= 90) rats. The mean input

resistance was 1003 ± 84.0 MΩ and 1034 ± 79.2 MΩ for sham and hypertensive PVN-

RVLM neurons, respectively (P= 0.8). Neuronal cell capacitance was significantly

reduced in PVN-RVLM neurons during hypertension (sham: 20.6 ± 1.0 pF; hypertensive:

17.6 ± 0.9 pF, P< 0.05), as previously reported (see 3.2.1.).

3.3.2. IA and IT are expressed and compete within individual PVN-RVLM neurons

In response to depolarizing steps from a hyperpolarized Vm, and under conditions

where K+ and Ca2+ currents were pharmacologically isolated (Methods), both a transient

outward K+ current (IA) and a transient inward Ca2+ current (IT) were observed,

respectively, in PVN-RVLM neurons (Fig. 21A). As we previously reported, IA and IT

were sensitive to block by 4-AP (2-5 mM) and NiCl2 (100 μM), respectively (not shown,

[180,185]. IT rise-time (10-90%) and τinactivation ranged from 3.2-25.4 ms (mean rise

time= 14.6 ± 3.1 ms) and from 8.4-50.6 ms (mean τinactivation= 31.9 ± 5.5 ms),

respectively. These values were similar to those we recently reported for IA in these

neurons [180]. While no differences in current density between IA and IT were observed

over the voltage range tested (P= 0.12, 2-way ANOVA, Fig. 21B), IT activated at a

91

Page 106: Functional Interplay Between Subthreshold Ion Channels in

Figure 21. IA and IT are expressed within individual PVN-RVLM neurons. A-

Representative examples of isolated IA and IT in different PVN-RVLM neurons. B-

Mean plots of IA (squares) and IT (triangles) current densities versus the command

potential. C- Representative example of a PVN-RVLM neuron in which both IA and IT

were recorded at the same command potential. Note that when the IA inhibitor 4-AP (5

mM) was applied (right), IA was blocked and IT became larger in amplitude. D- Plot of

current amplitude versus command potential of example trace in C (IA control= squares,

IT control= triangle, IT in 4-AP= upside down triangle). E- Summary data of the mean IT

current amplitude measured at a command potential of -20 mV before and after 4-AP

application (n= 9). Note the more robust IT amplitude upon IA inhibition. *P< 0.05.

92

Page 107: Functional Interplay Between Subthreshold Ion Channels in

93

Page 108: Functional Interplay Between Subthreshold Ion Channels in

significantly more hyperpolarized Vm than IA (IA= -46.3 ± 1.2 mV and IT= -54.6 ± 1.6

mV, P= 0.001).

Based on their similar magnitude and voltage-dependent properties, we

investigated whether both opposing currents “competed” at similar membrane potentials

[27,142,146,155]. To this end, IA and IT were simultaneously recorded in the same PVN-

RVLM neurons (Methods) (Fig. 21C). In all cases (n= 16, both sham and hypertensive),

both currents were observed within individual neurons. In most cases (13/16 neurons),

and as shown in the representative example in Fig. 21D, IT became apparent at a

significantly more hyperpolarized Vm than IA (IT= -51.8 ± 1.9 mV and IA= -36.2 ± 2.3

mV; P< 0.0005). Under this condition, the activation threshold of IA was more

depolarized than that observed when IA was pharmacologically isolated (see above). As

the Vm became more depolarized (-45 through -20 mV), IA grew in amplitude while

opposing changes were observed in IT (Fig. 21D). These results suggest an active

competition/balance between the two opposing currents, with IT and IA predominating at

relatively hyperpolarized and depolarized Vms, respectively. This was further supported

by our results showing that pharmacological inhibition of IA with 5 mM 4-AP

“unmasked” a more robust IT (P< 0.05, n= 9; Fig. 21D,E).

In recent studies, we demonstrated that the Kv4.3 and Cav3.1 K+ and Ca2+

subunits are predominantly expressed in PVN-RVLM neurons [115,180], likely

mediating IA and IT, respectively. To determine whether the relative expression of these

subunits influenced the balance of these two subthreshold conductances, we recorded IA

and IT in a subset of neurons (n= 15), in which the cytoplasmic content was subsequently

aspirated, and quantitative single-cell RT-PCR was performed (Methods). The ratio of

94

Page 109: Functional Interplay Between Subthreshold Ion Channels in

the amplitude of the evoked IT and IA at a Vm of -20 mV was plotted as a function of the

Cav3.1/Kv4.3 subunit expression for that particular neuron (Fig. 22). A significant

correlation between the two ratios was observed (r2= 0.61; P< 0.001).

3.3.3. IA /IT balance influences the expression and magnitude of LTS

It is well-established from studies in numerous neuronal types that while IT

mediates low threshold spikes (LTS) [129,185], IA on the other hand, underlies a transient

outward rectification (TOR) [19,65,133]. Thus, it is unclear within a given cell that

contains both IA and IT, such as PVN neurons, what dictates which membrane property is

observed. Therefore, we simultaneously recorded IA and IT under voltage-clamp

conditions, and determined whether an LTS or a TOR was observed in current-clamp

mode (Methods). We observed that in all cases where IT was evident at a command

potential more hyperpolarized than IA, an LTS would be observed (n= 40/49).

Conversely, in the few cases where IA was evident at a command potential more

hyperpolarized than IT, a TOR was observed (P< 0.0001, Fisher’s exact test) (Fig. 23A).

We then investigated whether pharmacologically altering the IT/IA balance

towards an inward current predominance affected the magnitude of the LTS. LTS

properties were analyzed before and after application of 5 mM 4-AP (n= 22). While the

LTS threshold was not significantly affected by IA blockade (∆= 0.62 ± 0.65 mV, P= 0.9),

both the LTS peak amplitude and area were significantly enhanced during IA blockade

(100.1 ± 18.1%, P< 0.0005 and 44.9 ± 16.2%, P= 0.01, respectively) (Fig. 23B).

95

Page 110: Functional Interplay Between Subthreshold Ion Channels in

Figure 22. The relative expression of mRNA for A-type and T-type channel

subunits correlates with the balance of these two subthreshold conductances. A-

Representative example of a single-cell real-time RT-PCR amplification plot obtained

from an identified PVN-RVLM neuron, in which the A-type channel subunit Kv4.3

(dashed line) and the T-type channel subunit Cav3.1 (dotted line) was tested against the

reference gene, GAPDH (whole line). The cycle numbers that the respective lines

crossed the x-axis are their threshold values (Ct). Note Cav3.1 mRNA crossed the x-axis

at a later cycle number than did Kv4.3, indicating less Cav3.1 mRNA expression

compared to Kv4.3 within the sampled neuron. B- Plot of the IT/IA amplitude ratio

measured at a command potential of -20 mV vs. the relative expression of Cav3.1/Kv4.3

mRNA (r2= 0.61, P< 0.001, n= 15). Note that the lower the ΔCt ratio value is, the larger

the expression of Cav3.1 relative to that of Kv4.3 is, and the larger the IT amplitude is

relative to IA.

96

Page 111: Functional Interplay Between Subthreshold Ion Channels in

97

Page 112: Functional Interplay Between Subthreshold Ion Channels in

Figure 23. The balance of IA and IT influences the expression and magnitude of

LTS. A- PVN-RVLM neurons were categorized in subgroups, depending on whether IT

or IA activated at a more hyperpolarized membrane potential, and whether an LTS or

TOR was observed upon membrane depolarization (n= 49). Note that when IT activated

first, an LTS was observed. Conversely, when IA activated first, a TOR was observed.

Right- Representative traces in which IA or IT activated at a more hyperpolarized potential

and the resultant TOR or LTS, respectively. B- Representative example of an LTS before

(left, black) and after (center, gray) application of the IA inhibitor, 4-AP (5 mM). Note

the increased peak and area of the LTS in 4-AP, when the two traces are overlapped

(right). C- Summary data of LTS area in sham and hypertensive PVN-RVLM neurons,

before (white) and after application of 4-AP (black). Note the LTS area was significantly

increased in the sham group upon 4-AP application. Also, note the more robust basal

LTS area in the hypertensive group (n= 24) compared to sham (n= 22), and the lack of a

further effect of 4-AP in this group of PVN-RVLM neurons. *P< 0.05

98

Page 113: Functional Interplay Between Subthreshold Ion Channels in

99

Page 114: Functional Interplay Between Subthreshold Ion Channels in

3.3.4. The balance of IA/IT is neurochemical-dependent

We wanted to investigate more on potential factors influencing the expression of

IT-LTS vs IA-TOR in PVN-RVLM neurons. Long descending PVN neurons, including

PVN-RVLM, are neurochemically heterogeneous, a proportion of which express the

neurohormones vasopressin (VP) and/or oxytocin (OT) [134,194,204]. Since

magnocellular OT and VP neuroendocrine neurons of the PVN and SON, express large IA

and TOR [65,133], we hypothesized that PVN-RVLM neurons in which IA overcomes IT

at more hyperpolarized Vms and expresses TOR, may also express VP and/or OT.

Therefore, in a subset of recordings (n= 12), we tested for the presence of OT/VP

immunoreactivity in recorded PVN-RVLM neurons (Methods). As shown in Fig. 24, the

majority of neurons (n= 6/7) in which IT was evident at a more hyperpolarized potential

than IA were VP/OT immunonegative, whereas all of those in which IA activated at a

more hyperpolarized potential than IT were VP/OT immunoreactive (n= 5/5) (P= 0.01,

Fisher’s exact test).

3.3.5. LTS-dependent increase in intracellular calcium in PVN-RVLM neurons

In addition to an increase in membrane conductance and induction of LTS,

activation of T-type Ca2+ channels could also lead to changes in intracellular Ca2+ levels

([Ca2+]ic), and activation of downstream Ca2+-dependent signaling, as previously shown

[58,74,156]. Thus, in order to determine whether LTS in PVN-RVLM neurons evoked

an increase in [Ca2+]ic and whether this was dependent on the IT/IA balance, we performed

simultaneous patch clamp recordings and confocal Ca2+ imaging. Our results indicate that

LTS in PVN-RVLM neurons consistently evoked transient increases in somatic [Ca2+]ic

(n= 13, Fig. 25). The peak of the Ca2+ transient was delayed with respect to the LTS peak

100

Page 115: Functional Interplay Between Subthreshold Ion Channels in

Figure 24. The neurochemical identity of PVN-RVLM neurons influences the

balance of IA and IT. A- Representative low magnification image (20x) of a PVN-

RVLM neuron displaying vasopressin and/or oxytocin (VP/OT) immunoreactivity.

Right- Summary data of numbers of PVN-RVLM neurons expressing VP/OT, segregated

into those in which either IA or IT activated at a more hyperpolarized command potential

(P= 0.01, Fisher’s exact test, n= 12). These cells also expressed LTS or TOR, based upon

which current activated at a more hyperpolarized Vm (as explained in Figure 23A). Note

the larger proportion of neurons expressing VP/OT were those in which IA activated at a

more hyperpolarized potential than IT.

101

Page 116: Functional Interplay Between Subthreshold Ion Channels in

102

Page 117: Functional Interplay Between Subthreshold Ion Channels in

Figure 25. LTS-evoked changes in intracellular calcium levels in PVN-RVLM

neurons. A- Representative confocal images of LTS-evoked changes in intracellular

calcium levels (pseudocolor) in a PVN-RVLM neuron, before and after 4-AP application.

Images of the basal intracellular calcium levels (black/white and #1), maximum LTS-

evoked peaks of intracellular calcium (#2), and subsequent return to basal levels (#3), are

displayed (scale units are F/F0). B- Traces depicting the time course of the LTS-evoked

relative changes in intracellular calcium levels obtained from the pseudocolor images in

A. Note the increase in the amplitude of intacellular calcium levels upon 4-AP addition

(gray) compared to control (black). C- Traces of the evoked LTS (grey) and resultant

change in intracellular calcium levels (black) are overlapped, to better depict the time-

course relationship between them. The upper and lower traces correspond to before and

after 4-AP application, respectively. Right- Expanded view of the LTS and peak of the

calcium transient. Note that the peak of the LTS occurs before the peak of the calcium

transient. D- Summary data of the peak of the LTS-evoked calcium transient in sham (n=

13) and hypertensive (n= 11) PVN-RVLM neurons, before and after 4-AP application.

Note the LTS-evoked calcium transient was significantly increased in the sham group

upon 4-AP application. Also, note the more robust LTS-evoked calcium response in the

hypertensive group compared to sham, and the lack of a further effect of 4-AP in this

group of PVN-RVLM neurons. E- Plot of the LTS-evoked calcium transient as a

function of the area of the LTS, depicting a positive correlation between the two

parameters (r2= 0.66, P< 0.005). Sham and hypertensive control groups were plotted

together. *P< 0.05, **P< 0.01

103

Page 118: Functional Interplay Between Subthreshold Ion Channels in

104

Page 119: Functional Interplay Between Subthreshold Ion Channels in

(time to peak LTS: 184.3 ± 24.3 ms; time to peak [Ca2+]ic: 305.0 ± 49.4 ms, P= 0.01), P<

0.0001), and in all cases, the evoked change in [Ca2+]ic persisted beyond the duration of

the LTS. The LTS-evoked change in [Ca2+]ic was blocked in nominal 0 Ca2+ ACSF, and

subthreshold depolarizations failed to evoke a change in [Ca2+]ic (not shown).

When the IT/IA balance was shifted in the presence of 4-AP, the LTS-evoked

change in [Ca2+]ic was significantly enhanced (P<0.01, Fig. 25D). In general, a strong

correlation between the LTS area and the peak of the Ca2+ transient was observed (r=

0.66, P< 0.005, Pearson’s correlation test, Fig. 25E). On the other hand, the kinetics of

the changes in [Ca2+]ic (i.e., rise and decay time courses) were not affected by 4-AP (not

shown).

In a few instances, relatively long dendritic processes contained within the same

focal plane were efficiently loaded with Fluo-5 (n= 3). In those cases, we found the

somatic-induced LTS to efficiently increase [Ca2+]ic along the dendritic process (Fig. 26).

As shown in the representative example in Fig. 26, the magnitude (i.e. area) of the Ca2+

transient increased as a function the distance from the soma (P< 0.0001, 2-way ANOVA),

and was significantly enhanced in the presence of 4-AP (P< 0.001, 2-way ANOVA) (Fig.

26C).

3.3.6. Changes during hypertension

Our studies indicate that the subthreshold balance between IT/IA influences both

membrane properties and activity-dependent Ca2+ dynamics in PVN-RVLM neurons.

Moreover, a pharmacological shift in this balance towards a predominance of IT resulted

in enhanced LTS and somatodendritic evoked Ca2+ transients. Finally, we wanted to

determine whether similar changes were observed during a pathophysiological condition

105

Page 120: Functional Interplay Between Subthreshold Ion Channels in

Figure 26. LTS-evoked changes in intracellular calcium in PVN-RVLM dendrites.

A- Representative confocal images (pseudocolor) of LTS-evoked changes in intracellular

calcium, in a PVN-RVLM neuron in which the dendrites were efficiently filled with

Fluo-5. Images represent periods corresponding to basal levels (left), and the peak of the

LTS-evoked calcium transient before (middle) and after 4-AP application. Note that the

evoked changes in dendritic intracellular calcium levels were enhanced upon 4-AP

addition (scale is a colorimetric range of the relative levels of intracellular calcium

compared to background; dark- no change from background, red- maximal change from

background). B- Representative traces of calcium transients from the cell in A at various

dendritic distances from the soma, under control conditions. Note that the farther

dendritic distance from the soma, the larger the relative change in intracellular calcium

was (dark blue- smallest change, light blue- largest change). C- Summary plot of the

mean area of the calcium transient along the dendritic process, before and after addition

of 4-AP, obtained from 3 neurons. Note the overall increase in the area of the

intracellular calcium levels after 4-AP application, and how the evoked calcium transient

linearly increased as a function of the dendritic distance (control r2= 0.67, 4-AP r2= 0.66,

binning= 10 µm).

106

Page 121: Functional Interplay Between Subthreshold Ion Channels in

107

Page 122: Functional Interplay Between Subthreshold Ion Channels in

in which the IT/IA balance is intrinsically changed. We have recently reported a

diminished IA availability in PVN-RVLM neurons during chronic hypertension (see 3.2).

Further supporting a change in the IT/IA balance during hypertension, we found, in this

study, an enhanced IT current amplitude and density in PVN-RVLM neurons during

hypertension (P< 0.0005 in both cases, 2-way ANOVA, Fig. 27A). No differences in

activation threshold and kinetic properties (rise time and τinactivation) were observed

between sham and hypertensive rats (not shown).

In agreement with a diminished IA and enhanced IT, results from single-cell real-

time PCR studies in identified PVN-RVLM neurons from control and hypertensive rats

(n= 20 and 31 neurons, respectively) indicated a 57% reduction and a 294% increase in

the relative expression levels of Kv4.3 mRNA Cav3.1 mRNA, respectively, during

hypertension (P< 0.05 in both cases) (Fig. 27C). In fact, a nearly 10 fold decrease in the

relative expression of Kv4.3 to Cav3.1 mRNA within individual PVN-RVLM neurons

was observed during hypertension (P< 0.005).

The magnitude of the evoked LTS was significantly larger in PVN-RVLM

neurons from hypertensive (n= 24) when compared to control rats (P< 0.05, Fig. 23C).

Moreover, blockade of IA in hypertensive rats, unlike what we observed in controls,

failed to significantly increase LTS magnitude (3.3 ± 8.1%, P= 0.4) (Fig. 23C). Likewise,

the magnitude of the LTS-evoked change in [Ca2+]ic was significantly larger in PVN-

RVLM neurons from hypertensive (n= 11) when compared to control rats (P< 0.05, Fig.

25D), and blockade of IA in hypertensive rats failed to significantly increase the

magnitude of the evoked Ca2+ transient (P> 0.1, not shown).

108

Page 123: Functional Interplay Between Subthreshold Ion Channels in

Figure 27. Shift in the balance of IA and IT during hypertension. A- Mean plots of IT

current densities obtained in PVN-RVLM neurons from sham (squares, n= 9) and

hypertensive (triangles, n= 7) rats. Note the significantly larger IT current density in the

hypertensive group. B- Representative examples of single-cell RT-PCR amplification

plots for Kv4.3 (dashed line), Cav3.1 (dotted line) and GAPDH (whole line) obtained

from PVN-RVLM neurons in a sham (black) and a hypertensive (gray) rat. C- Summary

data of the mean normalized expression ± range of Kv4.3 mRNA (left), Cav3.1 mRNA

(middle), and the Kv4.3 mRNA/Cav3.1 mRNA ratio (right) (n= 20 sham and 31

hypertensive neurons). Note the expression of Kv4.3 mRNA was reduced during

hypertension, while Cav3.1 mRNA expression was enhanced. Also, note the significant

reduction in the relative expression of Kv4.3 to Cav3.1, during hypertension. *P< 0.05,

**P< 0.01.

109

Page 124: Functional Interplay Between Subthreshold Ion Channels in

110

Page 125: Functional Interplay Between Subthreshold Ion Channels in

Altogether, these results support an imbalanced IT/IA relationship during

hypertension, resulting in an enhanced LTS and Ca2+ entry. Since IT has been shown to

regulate spontaneous activity in a variety of CNS neurons [94,159], we addressed

whether the subthreshold IT/IA imbalance resulted and/or contributed to increased

ongoing firing activity in PVN-RVLM neurons during hypertension. PVN-RVLM basal

spontaneous firing frequency was significantly higher in hypertensive when compared to

control rats (n= 12 and 8, respectively, P< 0.05). Bath application of NiCl2 (100 μM),

which at this concentration more efficiently blocks IT over HVAs [68,77] failed to affect

firing rate in control rats (P= 0.4). On the other hand, the firing rate in hypertensive

neurons was significantly reduced in the presence of NiCl2 (P< 0.005) (Fig. 28),

supporting an enhanced IT contribution to spontaneous firing activity during

hypertension.

111

Page 126: Functional Interplay Between Subthreshold Ion Channels in

Figure 28. IT contributes to enhanced spontaneous firing activity in PVN-RVLM

neurons during hypertension. A- Representative traces of spontaneous firing activity in

a hypertensive PVN-RVLM neuron, before and after addition of 100 μM NiCl2. B-

Summary data of the mean spontaneous firing activity in sham (n= 8) and hypertensive

(n= 12) neurons before and after NiCl2 application. Note the significantly elevated basal

firing rate in the hypertensive group compared to sham. Also note that while NiCl2

significantly reduced the firing activity in the hypertensive group, it failed to do so in the

sham group. *P< 0.05.

112

Page 127: Functional Interplay Between Subthreshold Ion Channels in

113

Page 128: Functional Interplay Between Subthreshold Ion Channels in

IV. DISCUSSION

The PVN is a highly heterogeneous region containing both magnocellular

neurosecretory (type I) and parvocellular (types II-III) neurons, the latter including both

neurosecretory and preautonomic neurons [193,194]. Previous studies indicate that these

two major PVN populations can be distinguished based on their electrophysiological

properties. Thus, while magnocellular neurosecretory neurons are characterized by the

expression of a robust transient outward rectification (TOR) [21,133,186], parvocellular

neurons, including preautonomic ones, are characterized by their ability to generate a

Ca2+-dependent low-threshold spike (LTS) [133,185]. Elegant work by Tasker and

colleagues showed that while both IA and IT are in fact present in both types of PVN

neurons, differences in their relative expression and voltage-dependent properties

determine their relative contribution to the general membrane properties characteristic of

these two neuronal populations (i.e., TOR and LTS, respectively) [133]. However,

whether and how these two opposing conductances compete with each other, and what

the impact of such competition is on the membrane excitability in PVN neurons, has not

been established. In this study, we focused on a specific group of parvocellular

preautonomic neurons, those that innervate the RVLM and are known to play an

important role in sympathetic control [4,42,110,195,205]. While most previous

electrophysiological studies on parvocellular neurons have focused on IT and the role of

114

Page 129: Functional Interplay Between Subthreshold Ion Channels in

the LTS [133,185], little is known about the characteristics and functional role of IA in

this neuronal population. Thus, major goals of this study included: a) to characterize the

basic biophysical and functional properties of IA in PVN-RVLM neurons, b) to determine

whether interactions between IA and IT influence membrane excitability in these neurons,

and c) to evaluate whether changes in the properties of IA and its interaction with IT

contribute to enhanced excitability of PVN-RVLM neurons during hypertension. Our

major hypothesis is that a balanced interaction between the subthreshold currents, IA and

IT, plays an important role in maintaining physiological neuronal excitability/activity.

Moreover, we hypothesize that during hypertension, a disbalanced interaction favoring IT

results in enhanced neuronal excitability, constituting thus an important underlying

mechanism contributing to the characteristic sympathoexcitation observed in this disease.

This hypothesis is supported by results from our work. Thus, our data indicate that in

addition to IT [185], preautonomic PVN neurons also express a functionally relevant IA,

which shapes the Na+ action potential waveform and modulates their firing activity. In

addition, our results support an important balance between IA and IT, which regulates

action potential-evoked changes in intracellular calcium levels, the magnitude of LTS and

overall firing activity. Finally, our results support a disbalance between IA and IT during

hypertension, favoring IT, resulting in more robust LTS, enhanced intracellular calcium

entry during LTS and single action potentials, and increased neuronal excitability.

4.1. A-type Potassium Currents in PVN-RVLM Neurons

4.1.1. Biophysical, pharmacological and molecular properties of IA in PVN-RVLM

neurons

115

Page 130: Functional Interplay Between Subthreshold Ion Channels in

While the biophysical and pharmacological properties of IA in PVN-RVLM

neurons reported herein are generally consistent with those previously described in other

CNS neuronal types [22,201], some interesting differences were observed. For example,

the activation threshold, half-activation and half-inactivation Vm of IA in PVN-RVLM

neurons were all found to be more hyperpolarized than those previously reported in non-

identified parvocellular PVN neurons, and very similar in fact to those reported in Type 1

PVN magnocellular neurons [125,133].

In agreement with previous reports on other CNS regions, including the

hypothalamus [64,87,125,133], we found IA in PVN-RVLM neurons to consistently show

high sensitivity to the traditional A-type blocker 4-AP, and low sensitivity to the

delayed-rectifier K+ channel blocker TEA (Fig. 9A) [166]. The sensitivity to 5 mM 4-AP

block in PVN-RVLM neurons was relatively consistent across neurons (32-54%

inhibition). This differs from the previously reported large variability in type II

parvocellular PVN neurons (10-69% inhibition) [133], likely due to the diverse neuronal

types (e.g, neurosecretory and/or alternative preautonomic ones) included in the type II

category. Thus, while IA seems to be ubiquitously present in the hypothalamus, some of

their major properties appear to be regulated in a cell type-dependent manner.

In this sense, both the biophysical and pharmacological properties of IA have been shown

to be dependent upon specific A-type K+ channels subunit composition, as well as the

auxiliary subunits and Kv channel-interacting proteins (KChIPs) associated with them

[6,11,37]. Thus, differences in A-type properties between PVN-RVLM and other PVN

neuronal populations could rely on the differential expression of molecularly diverse

channels. A-type K+ channels have been shown to consist of Kv1.4 or Kv4.1-3 subunits,

116

Page 131: Functional Interplay Between Subthreshold Ion Channels in

either homomerically, or heteromerically assembled with other subunits from the same

subfamily [189]. Despite the fact that the presence and functional relevance of A-type K+

currents in various hypothalamic neuronal types has been long recognized, the subunit

composition of the underlying channels remains at present unknown. Using tract-tracing

techniques in combination with immunohistochemistry, we found here that PVN-RVLM

neurons consistently expressed high immunoreactive levels for the Kv1.4 and 4.3

subunits, with almost no immunoreactivity for the Kv4.2 subunit. These results suggest

that A-type K+ channels in these neurons likely comprise one or both of the former two

subunits. As a caveat, the lack of commercially available Kv4.1 antibody prevented us

from assessing its expression in PVN-RVLM neurons. Moreover, the presence of

immunoreactivity for these subunits does not necessarily imply their incorporation into

functional channels. In this sense, previous studies indicate that Kv1- and Kv4-

containing channels recover from inactivation with time constants of several seconds or

milliseconds, respectively [37,173]. Thus, our findings, showing a recovery from

inactivation time constant of ~70 ms, suggest that functional channels in PVN-RVLM

neurons, at least those more likely activated during our recordings (i.e., those located in

somatic and proximal dendritic compartments), are probably comprised of Kv4.3

subunits. Supporting this, our single-cell RT-PCR studies in identified PVN-RVLM

neurons confirmed the presence of Kv4.3 mRNA. Further studies using subunit-selective

pharmacological tools for these Kv subunits would be needed to further clarify this issue.

In summary, an A-type K+ current is present within PVN-RVLM preautonomic neurons;

it is likely mediated by channels containing the Kv4.3 subunit, and has biophysical

117

Page 132: Functional Interplay Between Subthreshold Ion Channels in

properties similar to those previously reported centrally, yet unique to those reported in

the PVN.

4.1.2. IA influences excitability and repetitive firing properties of PVN-RVLM neurons

In agreement with previous reports in other neuronal types [107,201], our data

support the presence of a tonically active IA (i.e., “window current) in PVN-RVLM

neurons. This “window” current was found to be available between a narrow, though

apparently physiologically relevant range of membrane potentials. However, it is

important to take into account that due to differences in our recording conditions between

voltage- and current-clamp studies, and our limitation to examine the divalent cation

sensitivity of IA in these neurons (see Appendix B), the voltage-dependent availability of

the “window” current may not represent the physiological condition. Nonetheless, our

results showing membrane depolarization and increased input resistance following

pharmacological blockade of IA, suggest that IA may play a role in setting resting

membrane potential and/or in the regulation of subthreshold variations of membrane

potential in PVN-RVLM neurons.

By regulating action potential waveform and interspike intervals [135,164,166],

IA has been shown to play important roles in shaping temporal firing patterns of various

neuronal types. Our present studies indicate this to be the case in PVN presympathetic

neurons as well: pharmacological blockade of IA, and/or its voltage-dependent

inactivation, prolonged the duration of the action potential waveform. Importantly, the

firing activity of these neurons was differentially affected by IA blockade: while the

majority of the neurons (60%) responded to 4-AP with an increased firing rate, opposite

effects were observed in the remainder, suggesting a differential role of IA within PVN-

118

Page 133: Functional Interplay Between Subthreshold Ion Channels in

RVLM neurons. The precise mechanisms underlying such disparate roles are at present

unknown. However, some insights could be drawn from the present results. Worth

noting is the fact that some basic intrinsic properties differed between the two

subpopulations of 4-AP sensitive PVN-RVLM neurons (e.g., resting Vm, degree of

spontaneous activity, action potential duration), suggesting the presence of distinct

subsets of PVN-RVLM neurons. This is in agreement with previous observations

indicating a high degree of functional and neurochemical heterogeneity among

preautonomic PVN neurons, even within those innervating a common target

[185,192,194]. Alternatively, differences in 4-AP sensitivity could represent variability

within a single population of PVN-RVLM neurons. Future studies comparing for instance

the neurochemical identity of these subsets of PVN-RVLM neurons will provide more

insights into this important phenomenon. Nonetheless, the overall lack of significant

correlations between 4-AP induced changes in firing activity and any of the parameters

that differed between the two subpopulations of PVN-RVLM neurons, indicates that the

opposing effects of 4-AP on firing discharge were not due to differences in these

parameters per se. Similarly, both 4-AP sensitive subgroups of PVN-RVLM neurons

displayed either continuous or bursting firing patterns, indicating that initial firing pattern

was not a key factor influencing the type of response induced by IA blockade. Moreover,

the general properties of individual action potential waveforms in the two subsets of

PVN-RVLM neurons were similarly affected by 4-AP (i.e., spikes were increased in

amplitude and prolonged in duration) suggesting that changes in these action potential

parameters per se also did not contribute to the differential effect of 4-AP on the firing

discharge of these neurons.

119

Page 134: Functional Interplay Between Subthreshold Ion Channels in

Conversely, other neuronal properties were differentially affected by 4-AP in

these two subsets of PVN-RVLM neurons. As previously shown in SON and PVN

magnocellular neurosecretory neurons [7,20], the repolarizing phase of the action

potential in PVN-RVLM neurons was followed by a prominent HAP. In those neurons in

which an increase in firing rate was observed, the HAP peak reached a more

hyperpolarized membrane potential in 4-AP, an effect expected to result in a more

efficient removal of Na+ channel inactivation following each action potential. This in

turn may contribute to the more hyperpolarized Na+ spike threshold observed in this

subset of PVN-RVLM neurons. In fact, previous studies have shown that modulation of

the degree of Na+ channel inactivation is an effective mechanism by which IA modulates

firing properties [38,89]. Moreover, a steeper HAP decay slope, with concomitant

shorter interspike intervals, was also observed in this group in the presence of 4-AP.

Thus, these combined actions of 4-AP (i.e., hyperplolarized HAP peak, faster HAP decay

time course, shorter interspike intervals, and hyperpolarized Na+ spike threshold) likely

contribute to the enhanced firing discharge found in this subset of PVN-RVLM neurons.

On the other hand, in those neurons in which 4-AP decreased their firing

discharge, these parameters were either unchanged, or affected in opposite ways by 4-AP.

For example, a slower HAP decay time course, leading in turn to an overall enhanced

HAP area, was induced by 4-AP, effects that likely contributed to the decreased

interspike interval found in this subgroup of PVN-RVLM neurons.

In summary, our results indicate that IA differentially modulates firing discharge

in PVN-RVLM neurons, actions that seem to be in part mediated by differential and/or

120

Page 135: Functional Interplay Between Subthreshold Ion Channels in

opposing modulatory effects on HAP parameters. These results raise the intriguing

question of how such disparate actions can be mediated by IA.

4.1.3. The effects of IA on PVN-RVLM firing activity are Ca2+-dependent

An important finding of the present work is that 4-AP effects on the firing activity

of PVN-RVLM neurons were blocked by the broad spectrum Ca2+ channel blocker Cd2+.

Several mechanisms could be proposed to underlie this Ca2+-dependent effect. For

example, it could be argued that these effects are due to Ca2+-dependent inhibition of IA,

as previously shown in magnocellular supraoptic neurons [19]. However, the fact that a

large IA component was always present in a 0 Ca2+/EGTA/Cd2+ containing ACSF (see

Figs 2-5), along with lack of evidence for a diminished IA amplitude before and after

addition of Cd2+ to control ACSF (data not shown), argue against this possibility.

Alternatively, and due to the mostly overlapping voltage-dependent properties between IA

and IT, it is possible that blockade of IA results in amplification and/or prolongation of

the Ca2+-dependent LTS typically found in preautonomic PVN neurons (Fig. 1, [185]),

and as previously shown in other neuronal populations [27,156]. Thus, we hypothesized

that IA and IT act in a concerted manner to regulate subthreshold membrane excitability in

PVN-RVLM neurons, and that a diminished availability of IA (i.e., pharmacological

blockade or during a pathological condition, see below) could lead to increased influence

of IT and a Ca2+-dependent effect on neuronal firing discharge.

4.2. Balance between IA and IT in PVN-RVLM Neurons

Based on the results described above, we aimed in our study to characterize the

opposing competition between subthreshold ion channels, IA and IT, and to determine

121

Page 136: Functional Interplay Between Subthreshold Ion Channels in

whether a balance between these two conductances influenced neuronal excitability in

PVN-RVLM neurons.

4.2.1. Interplay between the subthreshold currents, IA and IT, in PVN-RVLM neurons

An opposing competition between the subthreshold ion currents, IA and IT, has

been well documented centrally [27,142,146,155], and our data, for the first time,

supports that this competition also plays an important role in modulating neuronal

excitability in PVN-RVLM neurons. Previously, it was widely believed that IA and IT

were functionally segregated among PVN neurons, with IA more abundantly expressed

and predominantly influencing type 1 magnocellular neurons, whereas IT was more

abundantly expressed and predominantly influencing type 2 parvocellular neurons [133].

However, our data supports the presence of overlapping IA and IT conductances within

individual PVN-RVLM neurons, at similar membrane potentials.

As mentioned above, differences in the relative expression and voltage-dependent

properties of IA and IT have been shown to determine their relative contribution to general

membrane properties (i.e., TOR and LTS, respectively) [133]. While the TOR is a

membrane property that delays excitation in response to membrane depolarization,

decreasing overall membrane excitability, the LTS facilitates and boosts membrane

depolarization, increasing overall membrane excitability. In the majority of PVN-RVLM

neurons, IT activated at a more hyperpolarized membrane potential than IA, corresponding

with the expression of an LTS in this group of neurons. Conversely, in a subpopulation of

PVN-RVLM neurons, IA activated at a more hyperpolarized membrane potential than IT,

corresponding with the expression of a TOR in this group of neurons.

122

Page 137: Functional Interplay Between Subthreshold Ion Channels in

Thus, the relative predominance of IT over IA seems to determine whether an LTS or

TOR will be expressed in PVN-RVLM neurons. Numerous factors could influence the

balance and predominance of one conductance over the other one. For example,

it has been suggested that differences in activation threshold can shift the balance

between IA and IT [155]. In this sense, our data shows that in the majority of cells, IT

activates at a more hyperpolarized potential than IA. This may account for the larger

proportion of cells expressing LTS that we observed, and has previously been shown in

preautonomic PVN neurons [185]. In addition, our data also demonstrated a positive

correlation between the relative expression of mRNA for Cav3.1/Kv4.3 subunits and

IT/IA amplitude within single cells. Thus, it is also possible that differences in the relative

expression of IT and IA channels may influence the ability of PVN-RVLM neurons to

express an excitatory (LTS) or inhibitory (TOR) membrane behavior. This will need to be

determined in future studies.

Interestingly, the balance between IA and IT, and the membrane properties

observed in PVN-RVLM neurones, seemed to be in part dependent on their

neurochemical phenotype. As has previously been demonstrated, OT and VP

neuroendocrine neurons of the PVN and SON express large IA and TOR [65,133].

However, a proportion of preautonomic PVN-RVLM neurons also express VP [204].

Surprisingly, our present studies indicate that the majority of PVN-RVLM neurons in

which IA activated at a more hyperpolarized Vm than IT, and expressed TOR, were

immunoreactive for VP/OT, while the majority of neurons in which IT activated at a more

hyperpolarized Vm were not.

123

Page 138: Functional Interplay Between Subthreshold Ion Channels in

In summary, IA and IT are both expressed within individual preautonomic PVN-

RVLM neurons, and opposingly interact at similar membrane potentials. Their relative

expression, will dictate the predominance of excitatory ( LTS) or inhibitory (TOR)

membrane behavior, a phenomenon that seems to be in part dependent on the

neurochemical phenotype of PVN-RVLM neurons (IA= VP/OT+, IT= VP/OT-),

4.2.2. IA/IT balance influences neuronal activity

As we have mentioned above, the balance of IA and IT dictates whether an LTS or

TOR is expressed. However, it has also been demonstrated that subtle changes in this

balance may affect the magnitude of these membrane properties. For example,

pharmacological blockade of IA has been shown to shift the balance towards a

predominance of IT, resulting in a more robust LTS [155]. Our data support this

phenomenon in PVN-RVLM neurons as well, as inhibition of IA with 4-AP resulted in a

significant increase in the peak and area of the LTS. Concomitantly, as the LTS

magnitude was enhanced, so too was the LTS-mediated change in intracellular calcium

levels, in both somatic and dendritic compartments. Thus, a decrease in IA magnitude can

lead to secondary increases in intracellular calcium levels. Altogether, this data indicate

that the Ca2+-dependent effect of IA on firing activity (see above) could be due to an

enhanced LTS-dependent Ca2+ conductance. As calcium is increased intracellularly, a

cascade of downstream Ca2+-dependent mechanisms could be activated, resulting in a

differential influence on firing activity, depending on the complement of Ca2+-dependent

mechanisms present in specific sets of neurons. It is currently unknown what calcium

dependent mechanisms may be activated within subpopulations of PVN-RVLM neurons,

and whether these are differentially distributed within subpopulations of PVN-RVLM

124

Page 139: Functional Interplay Between Subthreshold Ion Channels in

neurons. However, it is likely that a variety of voltage-gated Ca2+ channels, Ca2+-

dependent conductances, and Ca2+-dependent second messenger systems are present in

PVN-RVLM neurons. One possible mechanism that has been suggested to differentially

regulate neuronal excitability is the presence of two second-messenger systems within a

given cell, calcium/calmodulin-dependent protein kinase II (CamKII) and protein kinase

A (PKA) [206,213]. Phosphorylation is well known to regulate ion channel activity

[116], and CAMKII and PKA have been shown to increase and decrease IA, [199,206],

resulting in decreased and increased firing activity, respectively [206]. PKA has also

recently been shown to increase IT [30], the opposite effect as that of IA. Thus, the

relative presence and activity of these second messenger systems may contribute to the

differential Ca2+-dependent effect of IA blockade in subsets of PVN-RVLM neurons. In

summary, the balance of IA and IT influence PVN-RVLM neuronal activity, by

modulating the magnitude of LTS, intracellular calcium levels and likely activation of

downstream Ca2+-dependent mechanisms.

4.3. Altered Contribution of Subthreshold Ion Channels to Increased PVN-RVLM

Neuronal Excitability During Hypertension

Sympathoexcitation of a central origin is characteristic of hypertension [81,95],

and mounting evidence supports an important role for the PVN in the pathophysiology of

this disease state [4,35,76,85,149]. Increased neuronal excitability, associated with

elevated blood pressure and sympathetic activity, has recently been demonstrated in

PVN-RVLM neurons [4,120]. However, while alterations in various PVN

neurotransmitter systems have been reported during hypertension [46,54,80,96,120,121],

the contribution of altered intrinsic membrane properties to PVN neuronal hyperactivity

125

Page 140: Functional Interplay Between Subthreshold Ion Channels in

during hypertension remains unclear. As summarized above, results from our work

demonstrate that the IA/IT balance in PVN-RVLM neurons is a key factor influencing

their overall activity. Thus, another major goal of this work was to determine whether a

disbalance in the interplay between these opposing subthreshold conductances during

hypertension may contribute to increased neuronal excitability during this condition. To

this end, we used a renovascular hypertensive rat model, 2-kidney 1-clip, at a chronic

stage of hypertension, characterized by enhanced sympathetic nerve activity [136].

4.3.1. Altered balance between IA and IT in PVN-RVLM neurons, during hypertension

Our results indicate an overall reduction in the magnitude of IA through a wide

range of membrane potentials in PVN-RVLM neurons during hypertension. In addition to

changes in current amplitude, some critical voltage-dependent properties of IA were also

altered during hypertension. For example, the degree of IA steady-state inactivation was

enhanced in PVN-RVLM neurons during hypertension, and a hyperpolarizing shift in the

V1/2 inactivation was also observed in this condition. An increased degree of steady-state

inactivation during hypertension indicates that at a particular membrane potential, fewer

channels will be available for activation. Thus, the larger degree of voltage-dependent

inactivation of IA may be a contributing factor to the reduced magnitude of the evoked IA

during hypertension.

Another potential mechanism contributing to diminished IA availability during

hypertension is a reduction in single channel conductance. Our results using non-

stationary noise analysis of whole cell currents, a well-established approach

[5,163,176,177,197], do in fact support a diminished single channel conductance of A-

type K+ channels during hypertension. The values for A-type K+ single channel

126

Page 141: Functional Interplay Between Subthreshold Ion Channels in

conductance obtained in our studies are in general agreement with previous reports based

on single-channel recordings [40,97,100,202], further supporting the validity of this

approach.

Concomitant with a decreased IA magnitude during hypertension, both the current

amplitude and density of IT were found to be enhanced during hypertension. Thus, a

diminished IA along with an increased IT magnitude strongly supports a shift in the IA/IT

balance towards an inward current predominance during hypertension. Several lines of

evidence support such an imbalance during hypertension. Firstly, while 4-AP increased

the IT current amplitude in sham cells, 4-AP failed to do so in hypertensive cells. Single-

cell RT-PCR data also lends credence to this, as the mRNA levels of Kv4.3 and Cav3.1

were significantly decreased and increased, respectively, during hypertension. As well,

the relative expression of Kv4.3 to Cav3.1 mRNA in a given cell was significantly

reduced, during hypertension, indicating that the relative level of Cav3.1 mRNA

expression was nearly identical to that of Kv4.3. In summary, a shift in the balance

between IA and IT, favoring IT, as supported by current density and mRNA data, occurs in

PVN-RVLM neurons during hypertension.

4.3.2. Altered IA/IT balance during hypertension influences PVN-RVLM neuronal

activity

Consistent with a diminished IA availability during hypertension, our data also

demonstrated wider and slower-decaying action potentials in PVN-RVLM neurons

during hypertension. This is also supported by the fact that during hypertension, the

effects on action potential waveform of two procedures that diminish IA availability (i.e.,

membrane depolarization and the A-type K+ channel blocker 4-AP [180]) were blunted.

127

Page 142: Functional Interplay Between Subthreshold Ion Channels in

Prolongation of action potential duration due to diminished IA availability during

hypertension resulted in enhanced action potential-dependent Ca2+ entry. As previously

reported in other neuronal types [75], 4-AP enhanced the action-potential evoked Ca2+

transient, an effect that was diminished in hypertensive rats. It is unclear whether this

change in calcium entry is mediated in part by IT. However, since we demonstrated that

IT amplitude is enhanced by 4-AP application, at similar potentials that IA is active, it is

likely that IT contributed in part to the elevated action potential-dependent Ca2+ transient

observed during hypertension.

As stated above, the diminished IA availability, along with an increased

availability of IT, resulted in a shift in the IA/IT balance towards an inward predominance.

Our data indicates that this subthreshold disbalance had additional impacts on the

excitability of PVN-RVLM neurons during hypertension. In this sense, the magnitude of

the IT – mediated LTS (peak and area), as well as the LTS-mediated increase in

intracellular calcium levels, were significantly elevated during hypertension The blunted

effect of 4-AP on the LTS and LTS-mediated changes in intracellular calcium levels

further supports a contribution of the diminished IA to the IT predominance.

Since IT has been shown to regulate spontaneous firing activity in a variety of

CNS neurons [93,159], we wanted to determine if the IA/IT disbalance also contributed to

an enhanced ongoing firing discharge of PVN-RVLM neurons during hypertension

[4,120,180]. Our studies indicated that the basal spontaneous firing activity was

significantly increased in hypertensive cells compared to control, and that a low

concentration of NiCl2 (100 μM), significantly reduced the firing activity in the

hypertensive cells, but had no effect on control cells. This data supports an enhanced IT

128

Page 143: Functional Interplay Between Subthreshold Ion Channels in

contribution to the elevated spontaneous firing activity observed during hypertension in

PVN-RVLM neurons.

In summary, the work of this thesis indicates that there is an opposing competition

between IA and IT in PVN-RVLM neurons, which is shifted in favor of IT activity during

hypertension, resulting in an increased magnitude of low-threshold spikes, elevated

activity-dependent levels of intracellular calcium, as well as enhanced firing activity

during hypertension.

4.4. Source of Disbalanced Interplay between IA and IT During Hypertension

Our studies indicate that the disbalanced IA and IT interplay observed in PVN-

RVLM neurons, during hypertension, could be in part due to changes in the relative

expression of ion channel subunits underlying IA and IT (i.e. the diminished Kv4.3

subunit mRNA and the enhanced Cav3.1 subunit mRNA, respectively). It is currently

unclear what the source causing these observed changes is, although a possible

contributor is the neurotransmitter angiotensin II (AngII). ANGII is a key

neurotransmitter involved in the central control of cardiovascular function [184]. Within

the PVN, ANGII has been shown to increase sympathetic activity [214], and abundant

evidence supports enhanced PVN AngII activity during hypertension[80,96]. Numerous

studies indicate that ANGII can mediate its actions by modulating the expression/activity

of IA and IT. For example, AngII has been shown to diminish IA current amplitude and

diminish Kv4.3 mRNA in the RVLM of rats with congestive heart failure [71]. As well,

there is evidence indicating that AngII inhibits IA [125,148,202] and enhances the IT-

dependent LTS activity [181] within the hypothalamus. Furthermore, it has been

suggested that an AngII induced increase of the duration of the action potential waveform

129

Page 144: Functional Interplay Between Subthreshold Ion Channels in

is mediated by block of K+ conductances that contribute to the resting membrane

potential [181]. Results from our studies indicate that IA contributes to maintaining the

resting membrane potential, as well as suppressing the action potential duration in PVN-

RVLM neurons[180], and that IA influences are blunted during hypertension. Taken

together, it is reasonable to speculate that an enhanced AngII activity within the PVN

during hypertension, may contribute to the subthreshold shift between IA and IT, unveiled

by our studies in PVN-RVLM neurons during hypertension.

4.5. Future Studies

Certainly, existing evidence about AngII effects on the subthreshold ion channels,

IA and IT, and how this interaction may contribute to hypertension, would warrant further

studies on this important subject. AngII receptors have been shown to be densely

distributed throughout the PVN [2,73,117], and more specifically, has been shown to

acutely increase firing activity in PVN-RVLM neuron, in part through disinhibition of

GABAergic inhibitory currents [118] or through activation of a mixed cationic

conductance [26]. However, it is unclear whether AngII may also influence activity of

PVN-RVLM neurons through interactions with intrinsic mechanisms, such as IA and IT,

as has been reported in other hypothalamic regions [125,148,181,202]. AngII activity has

been reported to be elevated in the PVN, during hypertension, and since AngII has been

demonstrated to increase firing activity in PVN-RVLM neurons (see above), it would be

important to test in future studies the hypothesis that enhanced AngII activity in PVN-

RVLM neurons contributes to increased firing activity and sympathetic output, during

hypertension, by disbalancing IA and IT, in favor of IT.

130

Page 145: Functional Interplay Between Subthreshold Ion Channels in

If this mechanism is proven, understanding the underlying mechanism(s) by

which AngII may differentially act upon IA and IT would be important. It is well known

that AngII acts through AT1 receptors in the PVN [26,113,125,202], and that AT1

receptors are G-protein-coupled receptors [44]. In the hypothalamus, AngII elicits G-

protein-mediated stimulation of phosphoinositide hydrolysis, through AT1 receptors, and

subsequent activation of protein kinase C (PKC), resulting in a decrease in voltage-

dependent K+ currents and an increase in voltage-dependent Ca2+ currents [190]. Thus,

evaluating the role of PKC in the altered IA/IT balance during hypertension would be

important.

Finally, it would be extremely relevant to confirm at the whole animal level

whether the disbalance of IA/IT reported in our studies contributes to enhanced blood

pressure and sympathoexcitation during hypertension. This could be performed by

simultaneously recording blood pressure and renal sympathetic nerve activity in

anesthetized rats, while addressing the effects of local microinjections of compounds

within the PVN (e.g., 4-AP, NiCl2). Alternatively, instead of transiently affecting the

IA/IT balance pharmacologically, we could molecularly affect the balance for a prolonged

period of time using short interference RNA (siRNA) technology.

Another route that would be worthwhile to continue would be study of the

PVN-DVC neurons. As mentioned above, the PVN can influence cardiovascular

function through the regulation of both sympathetic and parasympathetic output. A

major target innervated by the PVN is the dorsal vagal complex (DVC) [55,72,98,178],

an autonomic center comprising parasympathetic neurons known to regulate tonic and

reflex cardiac activity [86,144,150,200]. One of the subnuclei that the DVC houses is the

131

Page 146: Functional Interplay Between Subthreshold Ion Channels in

NTS, a major visceroreceptive autonomic integrating center and key modulator of the

baroreflex [69,111,138,154]. NTS-projecting PVN neurons (PVN-NTS), upon

stimulation, have been shown to suppress the baroreflex through inhibition of the NTS

[33,55,56,145], inducing an elevation in heart rate and cardiac output during the defense

reaction or exercise. This phenomenon is also commonly observed during chronic

hypertension [14,66,78]. Our data indicated that both IA amplitude and current density

were significantly reduced in these neurons during hypertension. Thus, it would be

interesting to continue this line of study, and to determine the contribution of such change

to blunted baroreflex during hypertension.

4.6. General Conclusions

Hypertension is a highly deleterious disease, causing a huge health and economic

burden to the public health system. Sympathoexcitation has been identified as a key

pathophysiological mechanism during hypertension, contributing to morbidity and

mortality in this disorder. Despite its importance, little is known about the precise

underlying mechanisms contributing to hypertension-related sympathoexcitation. Results

from our work identified altered interactions among subthreshold membrane

conductances as a key mechanism contributing to exacerbated neuronal activity in

presympathetic hypothalamic neurons known to contribute to sympathoexcitation during

hypertension. It is expected that this information will help in the device of improved

and/or novel therapeutic approaches for the treatment of this prevalent cardiovascular

disorder.

132

Page 147: Functional Interplay Between Subthreshold Ion Channels in

Appendix A Abbreviations

4-AP 4-aminopyridine ABC avidin-biotin complex ACSF artificial cerebrospinal fluid AHP after-hyperpolarizing potential AngII angiotensin II [Ca2+]ic intracellular calcium levels CNS central nervous system CVLM caudal ventrolateral medulla DAB diaminobenzidine tetrahydrochloride DMX dorsal motor nucleus of the vagus DVC dorsal vagal complex F/F0 fractional fluorescence HAP hyperpolarizing afterpotential HVA high threshold calcium current IA A-type potassium current IKDR delayed rectifier potassium current IML intermediolateral column of the spinal cord IT T-type calcium current LJP liquid junction potential LTS low threshold spike NA nucleus ambiguus NTS nucleus of the solitary tract OT oxytocin PBS phosphate buffered saline PKC protein kinase C PVN paraventricular nucleus PVN-DVC paraventricular nucleus neurons projecting to the dorsal vagal complex PVN-NTS paraventricular nucleus neurons projecting to the nucleus of the solitary tract PVN-RVLM paraventricular nucleus neurons projecting to the rostral ventrolateral medulla RVLM rostral ventrolateral medulla SCN suprachiasmatic nucleus SFO subfornical organ SHR spontaneously hypertensive rat TEA tetraethyl ammonium TOR transient outward rectification TTX tetrodotoxin VLM ventrolateral medulla

133

Page 148: Functional Interplay Between Subthreshold Ion Channels in

Vm membrane potential VP vasopressin

134

Page 149: Functional Interplay Between Subthreshold Ion Channels in

Appendix B Methodological Considerations

Retrograde injections

To identify PVN-RVLM neurons, rhodamine-labeled fluorescent latex

microspheres were injected in the RVLM. This retrograde tracer results in highly

restricted and well-defined injection sites [102], and is commonly used to trace CNS

pathways, including PVN projections to brainstem nuclei [26,28,118,119,209]. While

this tracer is not taken up by fibers in passage [102,103,109], we cannot rule out potential

labeling of severed axons in the area of the injection, which could include PVN

projections to the spinal cord [132]. Labeling of severed axons in our studies seems

however unlikely, because injections that were misplaced, either rostrally, caudally or

laterally to the RVLM, in areas containing PVN descending axons running towards the

RVLM and/or the spinal cord [132], failed to retrogradely label neurons in the PVN.

Voltage-clamp recordings

Voltage- and space-clamp problems are often associated with voltage-clamp

recordings in brain slices. We believe however that these potential errors were

minimized in this study by the use of rigorous selection criteria for inclusion of neurons

(see Methods), along with the fact that PVN neurons are known to be relatively

electrotonically compact [133,185]. The good quality of our voltage-clamp experiments

is supported by several lines of evidence, including: A) stable and relatively low series

135

Page 150: Functional Interplay Between Subthreshold Ion Channels in

resistance throughout the recordings (~10 MΩ, compared to neuronal input resistance of

~1000 MΩ), B) small voltage errors (~1-3 mV) associated with uncompensated series

resistance, and C) lack of dependence of IA rates of activation and inactivation on varying

conditioning steps [133]. Although voltage errors cannot be completely eliminated when

recording intact neurons in a slice preparation, we believe the results reported here to be

relatively accurate. This is also supported by previous studies from hypothalamic

neurons in which similar IA voltage and kinetic properties were reported, including

studies from acutely dissociated [36,87] or intact neurons in a brain slice preparation

[133].

Pharmacological considerations

The voltage-dependent properties of IA are known to overlap with other voltage-

dependent conductances, in particular the T-type Ca2+ current, previously reported to be

present in parvocellular and identified preautonomic PVN neurons [133,185]. Thus, in

order to isolate IA from the low-threshold T-type and other voltage-dependent Ca2+

current, voltage-clamp recordings were obtained in the presence of nominal 0mM

extracellular Ca2+, the Ca2+ chelator EGTA and the broad spectrum Ca2+ channel blocker

Cd2+. A drawback of this approach however, is that previous studies in other neuronal

populations reported that extracellular divalent cations, including Ca2+ and Cd2+, could

affect the voltage-dependence of IA activation and inactivation [51,87,179,203]. Thus, it

is possible that the reported voltage-dependent activation and inactivation values of IA do

not reflect the physiological condition. This becomes important when comparing data

obtained from voltage-clamp and current-clamp experiments in this study, since the latter

were obtained in the presence of 2 mM Ca2+ and in the absence of Cd2+.

136

Page 151: Functional Interplay Between Subthreshold Ion Channels in

IT has previously been described in non-identified parvocellular PVN neurons

[133] to activate at a threshold of -59 mV. However, we observed the threshold of

activation of IT to be -54 mV in PVN-RVLM neurons. This difference may be due to

population differences between PVN-RVLM neurons and unidentified parvocellular

neurons. Observed differences may also be due in part to variations in extracellular

calcium concentrations as it has been described that increases in extracellular calcium

concentrations can shift the gating of T-type channels in a positive direction [70]. While

we used an extracellular calcium concentration of 5 mM, the study by Luther & Tasker

utilized 10 mM suggesting that it is more likely that the differences observed in threshold

are due to population differences as opposed to differences in extracellular calcium.

Early studies determined that T-type channels were more sensitive to inhibition

by low concentrations of nickel than HVA [68,77]. However, the HVA R-type calcium

channels have also been shown to be relatively nickel sensitive [160,210]. We, therefore,

attempted to specifically inhibit IT by using the selective T-type channel blocker NNC

55-0396 (10-50 μM, Sigma Aldrich) [92,123] via bath application. Unfortunately, no

effect was observed (data not shown). Therefore, a low concentration of nickel (100 μM)

was used in order to reduce the possible impact of inhibition on HVA. Interestingly,

nickel did not have an effect under sham conditions, only in hypertensive PVN-RVLM

neurons, in which we have demonstrated an increase in IT activity. However, we cannot

rule out the possibility of nickel partially inhibiting HVA and contributing to the

enhanced spontaneous firing activity observed during hypertension.

137

Page 152: Functional Interplay Between Subthreshold Ion Channels in

V. REFERENCES

[1] Aghajanian, G.K. and Rasmussen, K., Intracellular studies in the facial nucleus illustrating a simple new method for obtaining viable motoneurons in adult rat brain slices, Synapse, 3 (1989) 331-8.

[2] Aguilera, G., Young, W.S., Kiss, A. and Bathia, A., Direct regulation of hypothalamic corticotropin-releasing-hormone neurons by angiotensin II, Neuroendocrinology, 61 (1995) 437-44.

[3] Akine, A., Montanaro, M. and Allen, A.M., Hypothalamic paraventricular nucleus inhibition decreases renal sympathetic nerve activity in hypertensive and normotensive rats, Auton Neurosci, 108 (2003) 17-21.

[4] Allen, A.M., Inhibition of the hypothalamic paraventricular nucleus in spontaneously hypertensive rats dramatically reduces sympathetic vasomotor tone, Hypertension, 39 (2002) 275-80.

[5] Alvarez, O., Gonzalez, C. and Latorre, R., Counting channels: a tutorial guide on ion channel fluctuation analysis, Adv Physiol Educ, 26 (2002) 327-41.

[6] An, W.F., Bowlby, M.R., Betty, M., Cao, J., Ling, H.P., Mendoza, G., Hinson, J.W., Mattsson, K.I., Strassle, B.W., Trimmer, J.S. and Rhodes, K.J., Modulation of A-type potassium channels by a family of calcium sensors, Nature, 403 (2000) 553-6.

[7] Andrew, R.D. and Dudek, F.E., Intrinsic inhibition in magnocellular neuroendocrine cells of rat hypothalamus, J Physiol, 353 (1984) 171-85.

[8] Badoer, E., Hypothalamic paraventricular nucleus and cardiovascular regulation, Clin Exp Pharmacol Physiol, 28 (2001) 95-9.

[9] Bains, J.S. and Ferguson, A.V., Paraventricular nucleus neurons projecting to the spinal cord receive excitatory input from the subfornical organ, Am J Physiol, 268 (1995) R625-33.

[10] Bains, J.S. and Ferguson, A.V., Activation of N-methyl-D-aspartate receptors evokes calcium spikes in the dendrites of rat hypothalamic paraventricular nucleus neurons, Neuroscience, 90 (1999) 885-91.

[11] Beck, E.J., Bowlby, M., An, W.F., Rhodes, K.J. and Covarrubias, M., Remodelling inactivation gating of Kv4 channels by KChIP1, a small-molecular-weight calcium-binding protein, J Physiol, 538 (2002) 691-706.

[12] Beck, H., Ficker, E. and Heinemann, U., Properties of two voltage-activated potassium currents in acutely isolated juvenile rat dentate gyrus granule cells, J Neurophysiol, 68 (1992) 2086-99.

[13] Belugin, S. and Mifflin, S., Transient voltage-dependent potassium currents are reduced in NTS neurons isolated from renal wrap hypertensive rats, J Neurophysiol, 94 (2005) 3849-59.

138

Page 153: Functional Interplay Between Subthreshold Ion Channels in

[14] Berecek, K.H., Okuno, T., Nagahama, S. and Oparil, S., Altered vascular reactivity and baroreflex sensitivity induced by chronic central administration of captopril in the spontaneously hypertensive rat, Hypertension, 5 (1983) 689-700.

[15] Bergamaschi, C., Campos, R.R., Schor, N. and Lopes, O.U., Role of the rostral ventrolateral medulla in maintenance of blood pressure in rats with Goldblatt hypertension, Hypertension, 26 (1995) 1117-20.

[16] Bisset, G.W. and Chowdrey, H.S., Control of release of vasopressin by neuroendocrine reflexes, Q J Exp Physiol, 73 (1988) 811-72.

[17] Blessing, W., The Lower Brainstem and Bodily Homeostasis, (1997). [18] Bossu, J.L., Capogna, M., Debanne, D., McKinney, R.A. and Gahwiler, B.H.,

Somatic voltage-gated potassium currents of rat hippocampal pyramidal cells in organotypic slice cultures, J Physiol, 495 (Pt 2) (1996) 367-81.

[19] Bourque, C.W., Transient calcium-dependent potassium current in magnocellular neurosecretory cells of the rat supraoptic nucleus, J Physiol, 397 (1988) 331-47.

[20] Bourque, C.W., Randle, J.C. and Renaud, L.P., Calcium-dependent potassium conductance in rat supraoptic nucleus neurosecretory neurons, J Neurophysiol, 54 (1985) 1375-82.

[21] Bourque, C.W. and Renaud, L.P., Membrane properties of rat magnocellular neuroendocrine cells in vivo, Brain Res, 540 (1991) 349-52.

[22] Bouskila, Y. and Dudek, F.E., A rapidly activating type of outward rectifier K+ current and A-current in rat suprachiasmatic nucleus neurones, J Physiol, 488 (Pt 2) (1995) 339-50.

[23] Budzikowski, A.S., Vahid-Ansari, F. and Leenen, F.H., Chronic activation of brain areas by high-sodium diet in Dahl salt-sensitive rats, Am J Physiol, 274 (1998) H2046-52.

[24] Buijs, R.M., la Fleur, S.E., Wortel, J., Van Heyningen, C., Zuiddam, L., Mettenleiter, T.C., Kalsbeek, A., Nagai, K. and Niijima, A., The suprachiasmatic nucleus balances sympathetic and parasympathetic output to peripheral organs through separate preautonomic neurons, J Comp Neurol, 464 (2003) 36-48.

[25] Carvalho, T.H., Bergamaschi, C.T., Lopes, O.U. and Campos, R.R., Role of endogenous angiotensin II on glutamatergic actions in the rostral ventrolateral medulla in Goldblatt hypertensive rats, Hypertension, 42 (2003) 707-12.

[26] Cato, M.J. and Toney, G.M., Angiotensin II excites paraventricular nucleus neurons that innervate the rostral ventrolateral medulla: an in vitro patch-clamp study in brain slices, J Neurophysiol, 93 (2005) 403-13.

[27] Cavelier, P., Desplantez, T., Beekenkamp, H. and Bossu, J.L., K+ channel activation and low-threshold Ca2+ spike of rat cerebellar Purkinje cells in vitro, Neuroreport, 14 (2003) 167-71.

[28] Cham, J.L., Owens, N.C., Barden, J.A., Lawrence, A.J. and Badoer, E., P2X purinoceptor subtypes on paraventricular nucleus neurones projecting to the rostral ventrolateral medulla in the rat, Exp Physiol, 91 (2006) 403-11.

[29] Chaves, G.Z., Caligiorne, S.M., Santos, R.A., Khosla, M.C. and Campagnole-Santos, M.J., Modulation of the baroreflex control of heart rate by angiotensin-(1-7) at the nucleus tractus solitarii of normotensive and spontaneously hypertensive rats, J Hypertens, 18 (2000) 1841-8.

139

Page 154: Functional Interplay Between Subthreshold Ion Channels in

[30] Chemin, J., Mezghrani, A., Bidaud, I., Dupasquier, S., Marger, F., Barrere, C., Nargeot, J. and Lory, P., Temperature-dependent Modulation of CaV3 T-type Calcium Channels by Protein Kinases C and A in Mammalian Cells, J Biol Chem, 282 (2007) 32710-8.

[31] Chen, C., beta-Amyloid increases dendritic Ca2+ influx by inhibiting the A-type K+ current in hippocampal CA1 pyramidal neurons, Biochem Biophys Res Commun, 338 (2005) 1913-9.

[32] Chen, Q., Li, D.P. and Pan, H.L., Presynaptic alpha1 adrenergic receptors differentially regulate synaptic glutamate and GABA release to hypothalamic presympathetic neurons, J Pharmacol Exp Ther, 316 (2006) 733-42.

[33] Chen, Y.L., Chan, S.H. and Chan, J.Y., Participation of galanin in baroreflex inhibition of heart rate by hypothalamic PVN in rat, Am J Physiol, 271 (1996) H1823-8.

[34] Ciriello, J., Caverson, M.M. and Calaresu, F.R., Lateral hypothalamic and peripheral cardiovascular afferent inputs to ventrolateral medullary neurons, Brain Res, 347 (1985) 173-6.

[35] Ciriello, J., Kline, R.L., Zhang, T.X. and Caverson, M.M., Lesions of the paraventricular nucleus alter the development of spontaneous hypertension in the rat, Brain Res, 310 (1984) 355-9.

[36] Cobbett, P., Legendre, P. and Mason, W.T., Characterization of three types of potassium current in cultured neurones of rat supraoptic nucleus area, J Physiol, 410 (1989) 443-62.

[37] Coetzee, W.A., Amarillo, Y., Chiu, J., Chow, A., Lau, D., McCormack, T., Moreno, H., Nadal, M.S., Ozaita, A., Pountney, D., Saganich, M., Vega-Saenz de Miera, E. and Rudy, B., Molecular diversity of K+ channels, Ann N Y Acad Sci, 868 (1999) 233-85.

[38] Colbert, C.M., Magee, J.C., Hoffman, D.A. and Johnston, D., Slow recovery from inactivation of Na+ channels underlies the activity-dependent attenuation of dendritic action potentials in hippocampal CA1 pyramidal neurons, J Neurosci, 17 (1997) 6512-21.

[39] Connor, J.A. and Stevens, C.F., Prediction of repetitive firing behaviour from voltage clamp data on an isolated neurone soma, J Physiol, 213 (1971) 31-53.

[40] Cooper, E. and Shrier, A., Single-channel analysis of fast transient potassium currents from rat nodose neurones, J Physiol, 369 (1985) 199-208.

[41] Coote, J.H., A role for the paraventricular nucleus of the hypothalamus in the autonomic control of heart and kidney, Exp Physiol, 90 (2005) 169-73.

[42] Coote, J.H., Yang, Z., Pyner, S. and Deering, J., Control of sympathetic outflows by the hypothalamic paraventricular nucleus, Clin Exp Pharmacol Physiol, 25 (1998) 461-3.

[43] Cowley Jr., A., Quillen, E. and Barber, B., Further evidence for lack of baroreceptor control of long-term level of arterial pressure., In: Sleight, P (ed). Arterial Baroreceptors and Hypertension. (1980) 391-399.

[44] Crawford, K.W., Frey, E.A. and Cote, T.E., Angiotensin II receptor recognized by DuP753 regulates two distinct guanine nucleotide-binding protein signaling pathways, Mol Pharmacol, 41 (1992) 154-62.

140

Page 155: Functional Interplay Between Subthreshold Ion Channels in

[45] Cunningham, J.T., Herrera-Rosales, M., Martinez, M.A. and Mifflin, S., Identification of active central nervous system sites in renal wrap hypertensive rats, Hypertension, 49 (2007) 653-8.

[46] Czyzewska-Szafran, H., Wutkiewicz, M., Remiszewska, M., Jastrzebski, Z., Czarnecki, A. and Danysz, A., Down-regulation of the GABA-ergic system in selected brain areas of spontaneously hypertensive rats (SHR), Pol J Pharmacol Pharm, 41 (1989) 619-27.

[47] Dampney, R.A., Coleman, M.J., Fontes, M.A., Hirooka, Y., Horiuchi, J., Li, Y.W., Polson, J.W., Potts, P.D. and Tagawa, T., Central mechanisms underlying short- and long-term regulation of the cardiovascular system, Clin Exp Pharmacol Physiol, 29 (2002) 261-8.

[48] Dampney, R.A., Czachurski, J., Dembowsky, K., Goodchild, A.K. and Seller, H., Afferent connections and spinal projections of the pressor region in the rostral ventrolateral medulla of the cat, J Auton Nerv Syst, 20 (1987) 73-86.

[49] Dampney, R.A., Polson, J.W., Potts, P.D., Hirooka, Y. and Horiuchi, J., Functional organization of brain pathways subserving the baroreceptor reflex: studies in conscious animals using immediate early gene expression, Cell Mol Neurobiol, 23 (2003) 597-616.

[50] Davern, P.J. and Head, G.A., Fos-related antigen immunoreactivity after acute and chronic angiotensin II-induced hypertension in the rabbit brain, Hypertension, 49 (2007) 1170-7.

[51] Davidson, J.L. and Kehl, S.J., Changes of activation and inactivation gating of the transient potassium current of rat pituitary melanotrophs caused by micromolar Cd2+ and Zn2+, Can J Physiol Pharmacol, 73 (1995) 36-42.

[52] Dayas, C.V., Buller, K.M. and Day, T.A., Hypothalamic paraventricular nucleus neurons regulate medullary catecholamine cell responses to restraint stress, J Comp Neurol, 478 (2004) 22-34.

[53] De Koninck, Y. and Mody, I., Noise analysis of miniature IPSCs in adult rat brain slices: properties and modulation of synaptic GABAA receptor channels, J Neurophysiol, 71 (1994) 1318-35.

[54] DiCarlo, S.E., Zheng, H., Collins, H.L., Rodenbaugh, D.W. and Patel, K.P., Daily exercise normalizes the number of diaphorase (NOS) positive neurons in the hypothalamus of hypertensive rats, Brain Res, 955 (2002) 153-60.

[55] Duan, Y.F., Kopin, I.J. and Goldstein, D.S., Stimulation of the paraventricular nucleus modulates firing of neurons in the nucleus of the solitary tract, Am J Physiol, 277 (1999) R403-11.

[56] Duan, Y.F., Winters, R., McCabe, P.M., Green, E.J., Huang, Y. and Schneiderman, N., Cardiorespiratory components of defense reaction elicited from paraventricular nucleus, Physiol Behav, 61 (1997) 325-30.

[57] Edwards, M.A., Loxley, R.A., Powers-Martin, K., Lipski, J., McKitrick, D.J., Arnolda, L.F. and Phillips, J.K., Unique levels of expression of N-methyl-D-aspartate receptor subunits and neuronal nitric oxide synthase in the rostral ventrolateral medulla of the spontaneously hypertensive rat, Brain Res Mol Brain Res, 129 (2004) 33-43.

141

Page 156: Functional Interplay Between Subthreshold Ion Channels in

[58] Egger, V., Svoboda, K. and Mainen, Z.F., Dendrodendritic synaptic signals in olfactory bulb granule cells: local spine boost and global low-threshold spike, J Neurosci, 25 (2005) 3521-30.

[59] Esler, M. and Kaye, D., Increased sympathetic nervous system activity and its therapeutic reduction in arterial hypertension, portal hypertension and heart failure, J Auton Nerv Syst, 72 (1998) 210-9.

[60] Ferguson, A.V. and Bains, J.S., Actions of angiotensin in the subfornical organ and area postrema: implications for long term control of autonomic output, Clin Exp Pharmacol Physiol, 24 (1997) 96-101.

[61] Ferguson, A.V., Day, T.A. and Renaud, L.P., Subfornical organ efferents influence the excitability of neurohypophyseal and tuberoinfundibular paraventricular nucleus neurons in the rat, Neuroendocrinology, 39 (1984) 423-8.

[62] Ferguson, A.V. and Kasting, N.W., Electrical stimulation in subfornical organ increases plasma vasopressin concentrations in the conscious rat, Am J Physiol, 251 (1986) R425-8.

[63] Filosa, J.A., Bonev, A.D. and Nelson, M.T., Calcium dynamics in cortical astrocytes and arterioles during neurovascular coupling, Circ Res, 95 (2004) e73-81.

[64] Fisher, T.E. and Bourque, C.W., Properties of the transient K+ current in acutely isolated supraoptic neurons from adult rat, Adv Exp Med Biol, 449 (1998) 97-106.

[65] Fisher, T.E., Voisin, D.L. and Bourque, C.W., Density of transient K+ current influences excitability in acutely isolated vasopressin and oxytocin neurones of rat hypothalamus, J Physiol, 511 (Pt 2) (1998) 423-32.

[66] Floras, J.S., Hassan, M.O., Jones, J.V., Osikowska, B.A., Sever, P.S. and Sleight, P., Consequences of impaired arterial baroreflexes in essential hypertension: effects on pressor responses, plasma noradrenaline and blood pressure variability, J Hypertens, 6 (1988) 525-35.

[67] Flores-Hernandez, J., Galarraga, E., Pineda, J.C. and Bargas, J., Patterns of excitatory and inhibitory synaptic transmission in the rat neostriatum as revealed by 4-AP, J Neurophysiol, 72 (1994) 2246-56.

[68] Fox, A.P., Nowycky, M.C. and Tsien, R.W., Kinetic and pharmacological properties distinguishing three types of calcium currents in chick sensory neurones, J Physiol, 394 (1987) 149-72.

[69] Frigero, M., Bonagamba, L.G. and Machado, B.H., The gain of the baroreflex bradycardia is reduced by microinjection of NMDA receptor antagonists into the nucleus tractus solitarii of awake rats, J Auton Nerv Syst, 79 (2000) 28-33.

[70] Fukushima, Y. and Hagiwara, S., Currents carried by monovalent cations through calcium channels in mouse neoplastic B lymphocytes, J Physiol, 358 (1985) 255-84.

[71] Gao, L., Wang, W., Mann, E., Finch, M., Li, Y., Liu, D., Schultz, H.D. and Zucker, I.H., Sympathoexcitation in chronic heart failure: AngII induced inhibition of voltage-gated K+ channel, an in vivo and in vitro study, The FASEB Journal, 20 (2006) A1202-A1203.

[72] Gardiner, S.M. and Bennett, T., Endogenous vasopressin and baroreflex mechanisms, Brain Res, 396 (1986) 317-34.

142

Page 157: Functional Interplay Between Subthreshold Ion Channels in

[73] Gehlert, D.R., Gackenheimer, S.L. and Schober, D.A., Autoradiographic localization of subtypes of angiotensin II antagonist binding in the rat brain, Neuroscience, 44 (1991) 501-14.

[74] Goldberg, J.H., Lacefield, C.O. and Yuste, R., Global dendritic calcium spikes in mouse layer 5 low threshold spiking interneurones: implications for control of pyramidal cell bursting, J Physiol, 558 (2004) 465-78.

[75] Goldberg, J.H., Tamas, G. and Yuste, R., Ca2+ imaging of mouse neocortical interneurone dendrites: Ia-type K+ channels control action potential backpropagation, J Physiol, 551 (2003) 49-65.

[76] Goto, A., Ikeda, T., Tobian, L., Iwai, J. and Johnson, M.A., Brain lesions in the paraventricular nuclei and catecholaminergic neurons minimize salt hypertension in Dahl salt-sensitive rats, Clin Sci (Lond), 61 Suppl 7 (1981) 53s-55s.

[77] Gu, X. and Spitzer, N.C., Low-threshold Ca2+ current and its role in spontaneous elevations of intracellular Ca2+ in developing Xenopus neurons, J Neurosci, 13 (1993) 4936-48.

[78] Guo, G.B. and Abboud, F.M., Impaired central mediation of the arterial baroreflex in chronic renal hypertension, Am J Physiol, 246 (1984) H720-7.

[79] Guo, Z.L. and Moazzami, A.R., Involvement of nuclei in the hypothalamus in cardiac sympathoexcitatory reflexes in cats, Brain Res, 1006 (2004) 36-48.

[80] Gutkind, J.S., Kurihara, M., Castren, E. and Saavedra, J.M., Increased concentration of angiotensin II binding sites in selected brain areas of spontaneously hypertensive rats, J Hypertens, 6 (1988) 79-84.

[81] Guyenet, P.G., The sympathetic control of blood pressure, Nat Rev Neurosci, 7 (2006) 335-46.

[82] Hamill, O.P., Marty, A., Neher, E., Sakmann, B. and Sigworth, F.J., Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches, Pflugers Arch, 391 (1981) 85-100.

[83] Hardy, S.G., Hypothalamic projections to cardiovascular centers of the medulla, Brain Res, 894 (2001) 233-40.

[84] Haselton, J.R., Goering, J. and Patel, K.P., Parvocellular neurons of the paraventricular nucleus are involved in the reduction in renal nerve discharge during isotonic volume expansion, J Auton Nerv Syst, 50 (1994) 1-11.

[85] Herzig, T.C., Buchholz, R.A. and Haywood, J.R., Effects of paraventricular nucleus lesions on chronic renal hypertension, Am J Physiol, 261 (1991) H860-7.

[86] Higa, K.T., Mori, E., Viana, F.F., Morris, M. and Michelini, L.C., Baroreflex control of heart rate by oxytocin in the solitary-vagal complex, Am J Physiol Regul Integr Comp Physiol, 282 (2002) R537-45.

[87] Hlubek, M.D. and Cobbett, P., Outward potassium currents of supraoptic magnocellular neurosecretory cells isolated from the adult guinea-pig, J Physiol, 502 (Pt 1) (1997) 61-74.

[88] Hlubek, M.D. and Cobbett, P., Differential effects of K(+) channel blockers on frequency-dependent action potential broadening in supraoptic neurons, Brain Res Bull, 53 (2000) 203-9.

[89] Hoffman, D.A., Magee, J.C., Colbert, C.M. and Johnston, D., K+ channel regulation of signal propagation in dendrites of hippocampal pyramidal neurons, Nature, 387 (1997) 869-75.

143

Page 158: Functional Interplay Between Subthreshold Ion Channels in

[90] Hosoya, Y., Sugiura, Y., Okado, N., Loewy, A.D. and Kohno, K., Descending input from the hypothalamic paraventricular nucleus to sympathetic preganglionic neurons in the rat, Exp Brain Res, 85 (1991) 10-20.

[91] Hotson, J.R. and Prince, D.A., A calcium-activated hyperpolarization follows repetitive firing in hippocampal neurons, J Neurophysiol, 43 (1980) 409-19.

[92] Huang, L., Keyser, B.M., Tagmose, T.M., Hansen, J.B., Taylor, J.T., Zhuang, H., Zhang, M., Ragsdale, D.S. and Li, M., NNC 55-0396 [(1S,2S)-2-(2-(N-[(3-benzimidazol-2-yl)propyl]-N-methylamino)ethyl)-6-fluo ro-1,2,3,4-tetrahydro-1-isopropyl-2-naphtyl cyclopropanecarboxylate dihydrochloride]: a new selective inhibitor of T-type calcium channels, J Pharmacol Exp Ther, 309 (2004) 193-9.

[93] Huguenard, J.R., Low-threshold calcium currents in central nervous system neurons, Annu Rev Physiol, 58 (1996) 329-48.

[94] Huguenard, J.R., Coulter, D.A. and Prince, D.A., A fast transient potassium current in thalamic relay neurons: kinetics of activation and inactivation, J Neurophysiol, 66 (1991) 1304-15.

[95] Judy, W.V., Watanabe, A.M., Henry, D.P., Besch, H.R., Jr., Murphy, W.R. and Hockel, G.M., Sympathetic nerve activity: role in regulation of blood pressure in the spontaenously hypertensive rat, Circ Res, 38 (1976) 21-9.

[96] Jung, J.Y., Lee, J.U. and Kim, W.J., Enhanced activity of central adrenergic neurons in two-kidney, one clip hypertension in Sprague-Dawley rats, Neurosci Lett, 369 (2004) 14-8.

[97] Kang, J., Huguenard, J.R. and Prince, D.A., Voltage-gated potassium channels activated during action potentials in layer V neocortical pyramidal neurons, J Neurophysiol, 83 (2000) 70-80.

[98] Kannan, H. and Yamashita, H., Electrophysiological study of paraventricular nucleus neurons projecting to the dorsomedial medulla and their response to baroreceptor stimulation in rats, Brain Res, 279 (1983) 31-40.

[99] Kannan, H. and Yamashita, H., Connections of neurons in the region of the nucleus tractus solitarius with the hypothalamic paraventricular nucleus: their possible involvement in neural control of the cardiovascular system in rats, Brain Res, 329 (1985) 205-12.

[100] Kasai, H., Kameyama, M., Yamaguchi, K. and Fukuda, J., Single transient K channels in mammalian sensory neurons, Biophys J, 49 (1986) 1243-7.

[101] Katsunuma, N., Tsukamoto, K., Ito, S. and Kanmatsuse, K., Enhanced angiotensin-mediated responses in the nucleus tractus solitarii of spontaneously hypertensive rats, Brain Res Bull, 60 (2003) 209-14.

[102] Katz, L.C., Burkhalter, A. and Dreyer, W.J., Fluorescent latex microspheres as a retrograde neuronal marker for in vivo and in vitro studies of visual cortex, Nature, 310 (1984) 498-500.

[103] Katz, L.C. and Iarovici, D.M., Green fluorescent latex microspheres: a new retrograde tracer, Neuroscience, 34 (1990) 511-20.

[104] Kiely, J.M. and Gordon, F.J., Role of rostral ventrolateral medulla in centrally mediated pressor responses, Am J Physiol, 267 (1994) H1549-56.

[105] Kim, J., Wei, D.S. and Hoffman, D.A., Kv4 potassium channel subunits control action potential repolarization and frequency dependent broadening in hippocampal CA1 pyramidal neurons, J Physiol, 569 (2005) 41-57.

144

Page 159: Functional Interplay Between Subthreshold Ion Channels in

[106] Kirkpatrick, K. and Bourque, C.W., Dual role for calcium in the control of spike duration in rat supraoptic neuroendocrine cells, Neurosci Lett, 133 (1991) 271-4.

[107] Kloppenburg, P., Levini, R.M. and Harris-Warrick, R.M., Dopamine modulates two potassium currents and inhibits the intrinsic firing properties of an identified motor neuron in a central pattern generator network, J Neurophysiol, 81 (1999) 29-38.

[108] Koshiya, N., Huangfu, D. and Guyenet, P.G., Ventrolateral medulla and sympathetic chemoreflex in the rat, Brain Res, 609 (1993) 174-84.

[109] Krug, K., Smith, A.L. and Thompson, I.D., The development of topography in the hamster geniculo-cortical projection, J Neurosci, 18 (1998) 5766-76.

[110] Kubo, T., Hagiwara, Y., Sekiya, D., Chiba, S. and Fukumori, R., Cholinergic inputs to rostral ventrolateral medulla pressor neurons from hypothalamus, Brain Res Bull, 53 (2000) 275-82.

[111] Laguzzi, R., Nosjean, A., Mazarguil, H. and Allard, M., Cardiovascular effects induced by the stimulation of neuropeptide FF receptors in the dorsal vagal complex: an autoradiographic and pharmacological study in the rat, Brain Res, 711 (1996) 193-202.

[112] Lambert, R.C., McKenna, F., Maulet, Y., Talley, E.M., Bayliss, D.A., Cribbs, L.L., Lee, J.H., Perez-Reyes, E. and Feltz, A., Low-voltage-activated Ca2+ currents are generated by members of the CavT subunit family (alpha1G/H) in rat primary sensory neurons, J Neurosci, 18 (1998) 8605-13.

[113] Latchford, K.J. and Ferguson, A.V., Angiotensin depolarizes parvocellular neurons in paraventricular nucleus through modulation of putative nonselective cationic and potassium conductances, Am J Physiol Regul Integr Comp Physiol, 289 (2005) R52-8.

[114] Lebrun, C.J., Blume, A., Herdegen, T., Mollenhoff, E. and Unger, T., Complex activation of inducible transcription factors in the brain of normotensive and spontaneously hypertensive rats following central angiotensin II administration, Regul Pept, 66 (1996) 19-23.

[115] Lee, S., Han, T.H., Ryu, P.D. and Lee, S.Y., Molecular identification of ion channels for electrophysiological properties of rat hypothalamic neurons, Neuroscience Meeting Planner (2007) Program # 784.16.

[116] Levitan, I.B., Modulation of ion channels by protein phosphorylation and dephosphorylation, Annu Rev Physiol, 56 (1994) 193-212.

[117] Li, D.P., Chen, S.R. and Pan, H.L., Angiotensin II stimulates spinally projecting paraventricular neurons through presynaptic disinhibition, J Neurosci, 23 (2003) 5041-9.

[118] Li, D.P. and Pan, H.L., Angiotensin II attenuates synaptic GABA release and excites paraventricular-rostral ventrolateral medulla output neurons, J Pharmacol Exp Ther, 313 (2005) 1035-45.

[119] Li, D.P. and Pan, H.L., Plasticity of GABAergic Control of Hypothalamic Presympathetic Neurons in Hypertension, Am J Physiol Heart Circ Physiol (2005).

[120] Li, D.P. and Pan, H.L., Plasticity of GABAergic control of hypothalamic presympathetic neurons in hypertension, Am J Physiol Heart Circ Physiol, 290 (2006) H1110-9.

145

Page 160: Functional Interplay Between Subthreshold Ion Channels in

[121] Li, D.P. and Pan, H.L., Glutamatergic inputs in the hypothalamic paraventricular nucleus maintain sympathetic vasomotor tone in hypertension, Hypertension, 49 (2007) 916-25.

[122] Li, D.P. and Pan, H.L., Role of gamma-aminobutyric acid (GABA)A and GABAB receptors in paraventricular nucleus in control of sympathetic vasomotor tone in hypertension, J Pharmacol Exp Ther, 320 (2007) 615-26.

[123] Li, M., Hansen, J.B., Huang, L., Keyser, B.M. and Taylor, J.T., Towards selective antagonists of T-type calcium channels: design, characterization and potential applications of NNC 55-0396, Cardiovasc Drug Rev, 23 (2005) 173-96.

[124] Li, Y., Zhang, W. and Stern, J.E., Nitric oxide inhibits the firing activity of hypothalamic paraventricular neurons that innervate the medulla oblongata: role of GABA, Neuroscience, 118 (2003) 585-601.

[125] Li, Z. and Ferguson, A.V., Electrophysiological properties of paraventricular magnocellular neurons in rat brain slices: modulation of IA by angiotensin II, Neuroscience, 71 (1996) 133-45.

[126] Lien, C.C., Martina, M., Schultz, J.H., Ehmke, H. and Jonas, P., Gating, modulation and subunit composition of voltage-gated K(+) channels in dendritic inhibitory interneurones of rat hippocampus, J Physiol, 538 (2002) 405-19.

[127] Lindau, M. and Neher, E., Patch-clamp techniques for time-resolved capacitance measurements in single cells, Pflugers Arch, 411 (1988) 137-46.

[128] Livak, K.J. and Schmittgen, T.D., Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method, Methods, 25 (2001) 402-8.

[129] Llinas, R. and Yarom, Y., Properties and distribution of ionic conductances generating electroresponsiveness of mammalian inferior olivary neurones in vitro, J Physiol, 315 (1981) 569-84.

[130] Locke, R.E. and Nerbonne, J.M., Three kinetically distinct Ca2+-independent depolarization-activated K+ currents in callosal-projecting rat visual cortical neurons, J Neurophysiol, 78 (1997) 2309-20.

[131] Lovick, T.A. and Coote, J.H., Electrophysiological properties of paraventriculo-spinal neurones in the rat, Brain Res, 454 (1988) 123-30.

[132] Luiten, P.G., ter Horst, G.J., Karst, H. and Steffens, A.B., The course of paraventricular hypothalamic efferents to autonomic structures in medulla and spinal cord, Brain Res, 329 (1985) 374-8.

[133] Luther, J.A. and Tasker, J.G., Voltage-gated currents distinguish parvocellular from magnocellular neurones in the rat hypothalamic paraventricular nucleus, J Physiol, 523 Pt 1 (2000) 193-209.

[134] Mack, S.O., Kc, P., Wu, M., Coleman, B.R., Tolentino-Silva, F.P. and Haxhiu, M.A., Paraventricular oxytocin neurons are involved in neural modulation of breathing, J Appl Physiol, 92 (2002) 826-34.

[135] Magee, J., Hoffman, D., Colbert, C. and Johnston, D., Electrical and calcium signaling in dendrites of hippocampal pyramidal neurons, Annu Rev Physiol, 60 (1998) 327-46.

[136] Martinez-Maldonado, M., Pathophysiology of renovascular hypertension, Hypertension, 17 (1991) 707-19.

146

Page 161: Functional Interplay Between Subthreshold Ion Channels in

[137] Martins, A.S., Crescenzi, A., Stern, J.E., Bordin, S. and Michelini, L.C., Hypertension and exercise training differentially affect oxytocin and oxytocin receptor expression in the brain, Hypertension, 46 (2005) 1004-9.

[138] Matsumura, K., Averill, D.B. and Ferrario, C.M., Angiotensin II acts at AT1 receptors in the nucleus of the solitary tract to attenuate the baroreceptor reflex, Am J Physiol, 275 (1998) R1611-9.

[139] McCobb, D.P. and Beam, K.G., Action potential waveform voltage-clamp commands reveal striking differences in calcium entry via low and high voltage-activated calcium channels, Neuron, 7 (1991) 119-27.

[140] McKinley, M.J. and Johnson, A.K., The physiological regulation of thirst and fluid intake, News Physiol Sci, 19 (2004) 1-6.

[141] Mei, L., Zhang, J. and Mifflin, S., Hypertension alters GABA receptor-mediated inhibition of neurons in the nucleus of the solitary tract, Am J Physiol Regul Integr Comp Physiol, 285 (2003) R1276-86.

[142] Meis, S., Biella, G. and Pape, H.C., Interaction between low voltage-activated currents in reticular thalamic neurons in a rat model of absence epilepsy, Eur J Neurosci, 8 (1996) 2090-7.

[143] Melnick, I.V., Santos, S.F., Szokol, K., Szucs, P. and Safronov, B.V., Ionic basis of tonic firing in spinal substantia gelatinosa neurons of rat, J Neurophysiol, 91 (2004) 646-55.

[144] Mendelowitz, D., Firing properties of identified parasympathetic cardiac neurons in nucleus ambiguus, Am J Physiol, 271 (1996) H2609-14.

[145] Michelini, L.C., Vasopressin in the nucleus tractus solitarius: a modulator of baroreceptor reflex control of heart rate, Braz J Med Biol Res, 27 (1994) 1017-32.

[146] Molineux, M.L., Fernandez, F.R., Mehaffey, W.H. and Turner, R.W., A-type and T-type currents interact to produce a novel spike latency-voltage relationship in cerebellar stellate cells, J Neurosci, 25 (2005) 10863-73.

[147] Monteil, A., Chemin, J., Bourinet, E., Mennessier, G., Lory, P. and Nargeot, J., Molecular and functional properties of the human alpha(1G) subunit that forms T-type calcium channels, J Biol Chem, 275 (2000) 6090-100.

[148] Nagatomo, T., Inenaga, K. and Yamashita, H., Transient outward current in adult rat supraoptic neurones with slice patch-clamp technique: inhibition by angiotensin II, J Physiol, 485 (Pt 1) (1995) 87-96.

[149] Nakata, T., Takeda, K., Itho, H., Hirata, M., Kawasaki, S., Hayashi, J., Oguro, M., Sasaki, S. and Nakagawa, M., Paraventricular nucleus lesions attenuate the development of hypertension in DOCA/salt-treated rats, Am J Hypertens, 2 (1989) 625-30.

[150] Nosaka, S., Yamamoto, T. and Yasunaga, K., Localization of vagal cardioinhibitory preganglionic neurons with rat brain stem, J Comp Neurol, 186 (1979) 79-92.

[151] Osborn, J.W., Pathogenesis of hypertension in the sinoaortic-denervated spontaneously hypertensive rat, Hypertension, 18 (1991) 475-82.

[152] Osborn, J.W., Hypothesis: set-points and long-term control of arterial pressure. A theoretical argument for a long-term arterial pressure control system in the brain rather than the kidney, Clin Exp Pharmacol Physiol, 32 (2005) 384-93.

147

Page 162: Functional Interplay Between Subthreshold Ion Channels in

[153] Palkovits, M., Interconnections between the neuroendocrine hypothalamus and the central autonomic system. Geoffrey Harris Memorial Lecture, Kitakyushu, Japan, October 1998, Front Neuroendocrinol, 20 (1999) 270-95.

[154] Papas, S. and Ferguson, A.V., Electrophysiological evidence of baroreceptor input to area postrema, Am J Physiol, 261 (1991) R9-13.

[155] Pape, H.C., Budde, T., Mager, R. and Kisvarday, Z.F., Prevention of Ca(2+)-mediated action potentials in GABAergic local circuit neurones of rat thalamus by a transient K+ current, J Physiol, 478 Pt 3 (1994) 403-22.

[156] Pinato, G. and Midtgaard, J., Dendritic sodium spikelets and low-threshold calcium spikes in turtle olfactory bulb granule cells, J Neurophysiol, 93 (2005) 1285-94.

[157] Porter, J.P. and Brody, M.J., Neural projections from paraventricular nucleus that subserve vasomotor functions, Am J Physiol, 248 (1985) R271-81.

[158] Pyner, S. and Coote, J.H., Identification of branching paraventricular neurons of the hypothalamus that project to the rostroventrolateral medulla and spinal cord, Neuroscience, 100 (2000) 549-56.

[159] Raman, I.M. and Bean, B.P., Ionic currents underlying spontaneous action potentials in isolated cerebellar Purkinje neurons, J Neurosci, 19 (1999) 1663-74.

[160] Randall, A. and Tsien, R.W., Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons, J Neurosci, 15 (1995) 2995-3012.

[161] Ranson, R.N., Motawei, K., Pyner, S. and Coote, J.H., The paraventricular nucleus of the hypothalamus sends efferents to the spinal cord of the rat that closely appose sympathetic preganglionic neurones projecting to the stellate ganglion, Exp Brain Res, 120 (1998) 164-72.

[162] Reis, D.J., Ruggiero, D.A. and Morrison, S.F., The C1 area of the rostral ventrolateral medulla oblongata. A critical brainstem region for control of resting and reflex integration of arterial pressure, Am J Hypertens, 2 (1989) 363S-374S.

[163] Robinson, H.P., Sahara, Y. and Kawai, N., Nonstationary fluctuation analysis and direct resolution of single channel currents at postsynaptic sites, Biophys J, 59 (1991) 295-304.

[164] Rogawski, M.A., The A-current: how ubiquitous a feature of excitable cells is it? Trends Neurosci, 8 (1985) 214-219.

[165] Ross, C.A., Ruggiero, D.A., Joh, T.H., Park, D.H. and Reis, D.J., Rostral ventrolateral medulla: selective projections to the thoracic autonomic cell column from the region containing C1 adrenaline neurons, J Comp Neurol, 228 (1984) 168-85.

[166] Rudy, B., Diversity and ubiquity of K channels, Neuroscience, 25 (1988) 729-49. [167] Sah, P. and Davies, P., Calcium-activated potassium currents in mammalian

neurons, Clin Exp Pharmacol Physiol, 27 (2000) 657-63. [168] Saha, S., Spary, E.J., Maqbool, A., Asipu, A., Corbett, E.K. and Batten, T.F.,

Increased expression of AMPA receptor subunits in the nucleus of the solitary tract in the spontaneously hypertensive rat, Brain Res Mol Brain Res, 121 (2004) 37-49.

148

Page 163: Functional Interplay Between Subthreshold Ion Channels in

[169] Sawchenko, P.E. and Swanson, L.W., The organization of noradrenergic pathways from the brainstem to the paraventricular and supraoptic nuclei in the rat, Brain Res, 257 (1982) 275-325.

[170] Schreihofer, A.M., Ito, S. and Sved, A.F., Brain stem control of arterial pressure in chronic arterial baroreceptor-denervated rats, Am J Physiol Regul Integr Comp Physiol, 289 (2005) R1746-55.

[171] Schreihofer, A.M., Stornetta, R.L. and Guyenet, P.G., Regulation of sympathetic tone and arterial pressure by rostral ventrolateral medulla after depletion of C1 cells in rat, J Physiol, 529 Pt 1 (2000) 221-36.

[172] Segal, M., Rogawski, M.A. and Barker, J.L., A transient potassium conductance regulates the excitability of cultured hippocampal and spinal neurons, J Neurosci, 4 (1984) 604-9.

[173] Serodio, P. and Rudy, B., Differential expression of Kv4 K+ channel subunits mediating subthreshold transient K+ (A-type) currents in rat brain, J Neurophysiol, 79 (1998) 1081-91.

[174] Shafton, A.D., Ryan, A. and Badoer, E., Neurons in the hypothalamic paraventricular nucleus send collaterals to the spinal cord and to the rostral ventrolateral medulla in the rat, Brain Res, 801 (1998) 239-43.

[175] Shao, L.R., Halvorsrud, R., Borg-Graham, L. and Storm, J.F., The role of BK-type Ca2+-dependent K+ channels in spike broadening during repetitive firing in rat hippocampal pyramidal cells, J Physiol, 521 Pt 1 (1999) 135-46.

[176] Sigworth, F.J., The variance of sodium current fluctuations at the node of Ranvier, J Physiol, 307 (1980) 97-129.

[177] Sigworth, F.J., Covariance of nonstationary sodium current fluctuations at the node of Ranvier, Biophys J, 34 (1981) 111-33.

[178] Smith, P.M. and Ferguson, A.V., Paraventricular nucleus efferents influence area postrema neurons, Am J Physiol, 270 (1996) R342-47.

[179] Song, W.J., Tkatch, T., Baranauskas, G., Ichinohe, N., Kitai, S.T. and Surmeier, D.J., Somatodendritic depolarization-activated potassium currents in rat neostriatal cholinergic interneurons are predominantly of the A type and attributable to coexpression of Kv4.2 and Kv4.1 subunits, J Neurosci, 18 (1998) 3124-37.

[180] Sonner, P.M. and Stern, J.E., Functional role of A-type potassium currents in rat presympathetic PVN neurones, J Physiol, 582 (2007) 1219-38.

[181] Spanswick, D. and Renaud, L.P., Angiotensin II induces calcium-dependent rhythmic activity in a subpopulation of rat hypothalamic median preoptic nucleus neurons, J Neurophysiol, 93 (2005) 1970-6.

[182] Spyer, K.M., Neural organisation and control of the baroreceptor reflex, Rev Physiol Biochem Pharmacol, 88 (1981) 24-124.

[183] Standish, A., Enquist, L.W., Escardo, J.A. and Schwaber, J.S., Central neuronal circuit innervating the rat heart defined by transneuronal transport of pseudorabies virus, J Neurosci, 15 (1995) 1998-2012.

[184] Steckelings, U., Lebrun, C., Qadri, F., Veltmar, A. and Unger, T., Role of brain angiotensin in cardiovascular regulation, J Cardiovasc Pharmacol, 19 Suppl 6 (1992) S72-9.

149

Page 164: Functional Interplay Between Subthreshold Ion Channels in

[185] Stern, J.E., Electrophysiological and morphological properties of pre-autonomic neurones in the rat hypothalamic paraventricular nucleus, J Physiol, 537 (2001) 161-77.

[186] Stern, J.E. and Armstrong, W.E., Electrophysiological differences between oxytocin and vasopressin neurones recorded from female rats in vitro, J Physiol, 488 (Pt 3) (1995) 701-8.

[187] Storm, J.F., Action potential repolarization and a fast after-hyperpolarization in rat hippocampal pyramidal cells, J Physiol, 385 (1987) 733-59.

[188] Stornetta, R.L., Morrison, S.F., Ruggiero, D.A. and Reis, D.J., Neurons of rostral ventrolateral medulla mediate somatic pressor reflex, Am J Physiol, 256 (1989) R448-62.

[189] Stuhmer, W., Ruppersberg, J.P., Schroter, K.H., Sakmann, B., Stocker, M., Giese, K.P., Perschke, A., Baumann, A. and Pongs, O., Molecular basis of functional diversity of voltage-gated potassium channels in mammalian brain, Embo J, 8 (1989) 3235-44.

[190] Sumners, C., Zhu, M., Gelband, C.H. and Posner, P., Angiotensin II type 1 receptor modulation of neuronal K+ and Ca2+ currents: intracellular mechanisms, Am J Physiol, 271 (1996) C154-63.

[191] Sundaram, K., Johnson, S.M. and Felder, R.B., Altered expression of delayed excitation in medial NTS neurons of spontaneously hypertensive rats, Neurosci Lett, 225 (1997) 205-9.

[192] Swanson, L.W. and Kuypers, H.G., The paraventricular nucleus of the hypothalamus: cytoarchitectonic subdivisions and organization of projections to the pituitary, dorsal vagal complex, and spinal cord as demonstrated by retrograde fluorescence double-labeling methods, J Comp Neurol, 194 (1980) 555-70.

[193] Swanson, L.W. and Sawchenko, P.E., Paraventricular nucleus: a site for the integration of neuroendocrine and autonomic mechanisms, Neuroendocrinology, 31 (1980) 410-7.

[194] Swanson, L.W. and Sawchenko, P.E., Hypothalamic integration: organization of the paraventricular and supraoptic nuclei, Annu Rev Neurosci, 6 (1983) 269-324.

[195] Tagawa, T. and Dampney, R.A., AT(1) receptors mediate excitatory inputs to rostral ventrolateral medulla pressor neurons from hypothalamus, Hypertension, 34 (1999) 1301-7.

[196] Tanaka, J., Hayashi, Y., Sakamaki, K., Okumura, T. and Nomura, M., Activation of the subfornical organ enhances extracellular noradrenaline concentrations in the hypothalamic paraventricular nucleus in the rat, Brain Res Bull, 54 (2001) 421-5.

[197] Traynelis, S.F., Silver, R.A. and Cull-Candy, S.G., Estimated conductance of glutamate receptor channels activated during EPSCs at the cerebellar mossy fiber-granule cell synapse, Neuron, 11 (1993) 279-89.

[198] Turton, W.E., Ciriello, J. and Calaresu, F.R., Changes in forebrain hexokinase activity after aortic baroreceptor denervation, Am J Physiol, 251 (1986) R274-81.

[199] van der Heyden, M.A., Wijnhoven, T.J. and Opthof, T., Molecular aspects of adrenergic modulation of the transient outward current, Cardiovasc Res, 71 (2006) 430-42.

150

Page 165: Functional Interplay Between Subthreshold Ion Channels in

[200] Varga, K. and Kunos, G., Inhibition of baroreflex bradycardia by ethanol involves both GABAA and GABAB receptors in the brainstem of the rat, Eur J Pharmacol, 214 (1992) 223-32.

[201] Wang, D. and Schreurs, B.G., Characteristics of IA currents in adult rabbit cerebellar Purkinje cells, Brain Res, 1096 (2006) 85-96.

[202] Wang, D., Sumners, C., Posner, P. and Gelband, C.H., A-type K+ current in neurons cultured from neonatal rat hypothalamus and brain stem: modulation by angiotensin II, J Neurophysiol, 78 (1997) 1021-9.

[203] Wickenden, A.D., Tsushima, R.G., Losito, V.A., Kaprielian, R. and Backx, P.H., Effect of Cd2+ on Kv4.2 and Kv1.4 expressed in Xenopus oocytes and on the transient outward currents in rat and rabbit ventricular myocytes, Cell Physiol Biochem, 9 (1999) 11-28.

[204] Yang, Z., Bertram, D. and Coote, J.H., The role of glutamate and vasopressin in the excitation of RVL neurones by paraventricular neurones, Brain Res, 908 (2001) 99-103.

[205] Yang, Z. and Coote, J.H., Influence of the hypothalamic paraventricular nucleus on cardiovascular neurones in the rostral ventrolateral medulla of the rat, J Physiol, 513 (Pt 2) (1998) 521-30.

[206] Yao, W.D. and Wu, C.F., Distinct roles of CaMKII and PKA in regulation of firing patterns and K(+) currents in Drosophila neurons, J Neurophysiol, 85 (2001) 1384-94.

[207] Yarowsky, P. and Weinreich, D., Loss of accommodation in sympathetic neurons from spontaneously hypertensive rats, Hypertension, 7 (1985) 268-76.

[208] Yu, Y.Q., Xia, Q. and Zhang, R.B., [Roles of the central nucleus of amygdala in the cardiovascular response elicited by the paraventricular nucleus of hypothalamus], Zhejiang Da Xue Xue Bao Yi Xue Ban, 35 (2006) 172-7.

[209] Zahner, M.R., Li, D.P. and Pan, H.L., Benzodiazepine inhibits hypothalamic presympathetic neurons by potentiation of GABAergic synaptic input, Neuropharmacology, 52 (2007) 467-75.

[210] Zamponi, G.W., Bourinet, E. and Snutch, T.P., Nickel block of a family of neuronal calcium channels: subtype- and subunit-dependent action at multiple sites, J Membr Biol, 151 (1996) 77-90.

[211] Zhang, T.X. and Ciriello, J., Effect of paraventricular nucleus lesions on arterial pressure and heart rate after aortic baroreceptor denervation in the rat, Brain Res, 341 (1985) 101-9.

[212] Zhang, T.X. and Ciriello, J., Kainic acid lesions of paraventricular nucleus neurons reverse the elevated arterial pressure after aortic baroreceptor denervation in the rat, Brain Res, 358 (1985) 334-8.

[213] Zhao, M.L. and Wu, C.F., Alterations in frequency coding and activity dependence of excitability in cultured neurons of Drosophila memory mutants, J Neurosci, 17 (1997) 2187-99.

[214] Zhu, G.Q., Patel, K.P., Zucker, I.H. and Wang, W., Microinjection of ANG II into paraventricular nucleus enhances cardiac sympathetic afferent reflex in rats, Am J Physiol Heart Circ Physiol, 282 (2002) H2039-45.

151