the role of the hypothalamic paraventricular nucleus in

199
Loyola University Chicago Loyola University Chicago Loyola eCommons Loyola eCommons Dissertations Theses and Dissertations 1989 The Role of the Hypothalamic Paraventricular Nucleus in Stress- The Role of the Hypothalamic Paraventricular Nucleus in Stress- Induced Renin and Corticosterone Secretion Induced Renin and Corticosterone Secretion Kathy Richardson Morton Loyola University Chicago Follow this and additional works at: https://ecommons.luc.edu/luc_diss Part of the Neuroscience and Neurobiology Commons Recommended Citation Recommended Citation Morton, Kathy Richardson, "The Role of the Hypothalamic Paraventricular Nucleus in Stress-Induced Renin and Corticosterone Secretion" (1989). Dissertations. 2653. https://ecommons.luc.edu/luc_diss/2653 This Dissertation is brought to you for free and open access by the Theses and Dissertations at Loyola eCommons. It has been accepted for inclusion in Dissertations by an authorized administrator of Loyola eCommons. For more information, please contact [email protected]. This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 3.0 License. Copyright © 1989 Kathy Richardson Morton

Upload: others

Post on 22-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Loyola University Chicago Loyola University Chicago

Loyola eCommons Loyola eCommons

Dissertations Theses and Dissertations

1989

The Role of the Hypothalamic Paraventricular Nucleus in Stress-The Role of the Hypothalamic Paraventricular Nucleus in Stress-

Induced Renin and Corticosterone Secretion Induced Renin and Corticosterone Secretion

Kathy Richardson Morton Loyola University Chicago

Follow this and additional works at: https://ecommons.luc.edu/luc_diss

Part of the Neuroscience and Neurobiology Commons

Recommended Citation Recommended Citation Morton, Kathy Richardson, "The Role of the Hypothalamic Paraventricular Nucleus in Stress-Induced Renin and Corticosterone Secretion" (1989). Dissertations. 2653. https://ecommons.luc.edu/luc_diss/2653

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

This work is licensed under a Creative Commons Attribution-Noncommercial-No Derivative Works 3.0 License. Copyright © 1989 Kathy Richardson Morton

,. , ... ·--:-i···,1Ty UBRJ.\i·:;y .. LG'(()L!\ l::·~ \ \i L.:n,::;:.

. MEDICAL CE:NTEH .: ... l'.:··

THE ROLE OF THE HYPOTHALAMIC PARAVENTRICULAR NUCLEUS

IN STRESS-INDUCED RENIN AND CORTICOSTERONE SECRETION

by

Kathy Richardson Morton

A Dissertation Submitted to the Faculty of the Graduate School

of Loyola University of Chicago in Partial Fulfillment

of the Requirements for the Degree of

Doctor of Philosophy

May

1989

ACKNOWLEDGEMENTS

I express my thanks to Dr. Louis Van de Kar, for providing the

support, guidance and friendship that has made this dissertation a

productive and enjoyable experience. Appreciation is also extended to my

committee members for their suggestions and insightful discussions during

the course of this dissertation work. Special thanks is given to Dr. Mark

Brownfield for his exhaustive help with the irnrnunocytochemistry and

histological analyses; to Dr. Stanley Lorens for use of the stress

chambers and assistance with the sotalol injections; to Dr. Celeste Napier

for experimental design suggestions and help with the microinj ection

techniques; to Dr. Stephen Jones for help with the fluorometric

catecholamine assay; and to Dr. Alexander Karczmar and Dr. Israel Hanin

for their guidance as Departmental Chairmen.

I am grateful to my laboratory co-workers, Ms. Kayoko Kunimoto and

Dr. Janice Urban, for their assistance on many occasions. I also thank

Ms. Heather Read for the preparation of the brain pathways diagram. I

express my gratitude to Ms. Diane Pikowski and her parents, Mr. and Mrs.

Carl Villafan, for their concern and support for both myself and my

daughter Lindsay. I also wish to thank Gerry and Dorothy Hunchak, Toni

Price, Lynn Rohrmann, Loek Van de Kar, Susan Schmitt, Celeste Napier, Ron

Price, Peter Rittenhouse and Denise Lokhorst for their kindness during my

trips to Chicago.

Finally, I gratefully acknowledge the support of the faculty, staff

and students of the Department of Pharmacology; the Graduate School; Dee

Miller; Mary Ann Jurgus; and the financial backing supplied by Loyola

University and the Arthur J. Schmitt Foundation.

ii

DEDICATION

This dissertation is dedicated to my parents,

Virgil and Audrey Bullington, for their constant

support and encouragement, and to my daughter, Lindsay.

iii

VITA

The author, Kathy Richardson Morton, was born on January 26, 1957,

in Knoxville, Tennessee. She completed her secondary education at West

High School in Knoxville, Tennessee in June, 1975. In September, 1975,

Ms. Morton entered the University of Tennessee, where she graduated magna

cum laude with a Bachelor of Arts degree in Biology in June, 1978.

Following graduation, she was employed as an Associate Chemist for the

Bristol-Myers Company in Cincinnati, Ohio. In 1982, Ms. Morton did a

summer rotation in the laboratory of Dr. Robert Ten Eick at Northwestern

University.

In January, 1983, Ms. Morton was accepted as a graduate student in

the Department of Pharmacology at Loyola University of Chicago Stritch

School of Medicine, where she was awarded a Loyola University Basic

Science Fellowship. In July, 1986, she received a Sigma Xi Grant-in-Aid

of Research, and in March, 1987, was awarded an Arthur J. Schmitt

Dissertation Fellowship. In May, 1988, Ms. Morton received an award for

Best Neuroscience Dissertation from Loyola University for the Chicago

Chapter of Neuroscience. The author also served as a Representative to

the Graduate Student Council (1983-84) and the Pharmacology Faculty

Meetings (1987). She is a student member of the Society for Neuroscience.

Ms. Morton has accepted a post-doctoral position in the laboratory

of Dr. Neil MacLusky in the.Department of Reproductive Science at the

University of Toronto and Toronto General Hospital. She has been awarded

a post-doctoral fellowship from the Medical Research Council of Canada.

iv

TABLE OF CONTENTS

ACKNOWLEDGEMENTS. ii

DEDICATION. iii

VITA .. .iv

LIST OF TABLES. .viii

LIST OF FIGURES .. ix

Chapter I . INTRODUCTION. . . . . . . . . . . . . . . . . . . . . . . . . . l

II. LITERATURE REVIEW . A. RENIN ..... .

1. Introduction. 2. Description of Renin. 3. Physiological Aspects of the Angiotensins

a. Angiotensinogen. b. Angiotensin I. . . .. . c. Angiotensin II . . .. .

4. Regulation of Renin Secretion a. Introduction ...... . b. Role of the Sympathetic Nervous System

in Renin Secretion . . . . . . . . c. Role of Extra-Renal Beta Receptors

in Renin Secretion . . . . . . . d. Role of Brain Serotonin Neurons

5 5 5 5 6 6 7 7 9 9

.10

.11

in Renin Secretion . . . . . . . .12 1. Role of Beta Receptors in the Serotonergic

Stimulation of Renin Secretion . . . . .13 e. Role of the Hypothalamic Paraventricular Nucleus

(PVN) in Renin Secretion . . . . . .15 f. Other Hypothalamic Areas Involved

in Renin Secretion . . . . . . . . 16 g. Extrahypothalamic Pathways Involved

in Renin Secretion . . . . . . . . . .17 h. Renin and Angiotensin in the Brain . .17 i. Role of Brain Catecholamines in Renin Secretion. .19

B. CORTICOSTERONE . . . . . . . . . . .21 1. Introduction. . . . . . . . . . .21 2. Corticotropin Releasing Factor (CRF). .22 3. Other Factors Affecting ACTH Release. .25 4. Role of Brain Catecholamines in

Corticosterone Secretion. .26 a. Introduction . . . . .26 b. Anatomical Evidence. .26

v

c. Role of the Ventral Noradrenergic Bundle in Corticosterone Secretion . . . . . . . . . .

d. Stimulation of CRF/Corticosterone Secretion by Injections of Epinephrine and Norepinephrine.

e. Role of Dopamine Neurons in Corticosterone Secretion . . . . . . . . .

5. Role of 5-HT Neurons in Corticosterone Secretion. C. STRESS .... .

1. Introduction .. . a. Definition .. b. Cannon's Alarm Reaction. c. Selye's General Adaptation Syndrome. d. Sympathetic and Endocrine Stress Responses e. Physiological Changes in Response to Stress.

2. The Conditioned Fear Paradigm ..... a. Description. . . . . . . . . . . . . b. Pavlov's Dog: Classical Conditioning c. Definition of Fear d. Measures of Fear . . . . . . . . . e. Emotional Factors ........ .

3. Cardiovascular Changes in Response to Classical Conditioning ................ .

4. Neuroendocrine Changes in Response to Stress. a. Stress-Induced Renin Secretion ..... b. Stress-Induced Corticosterone Secretion.

1. Introduction ....... . 2. Stress-Induced Secretion of CRF .. . 3. Stress-Induced Secretion of ACTH .. . 4. Role of Vasopressin in Stress-Induced

Corticosterone Secretion. . . . . . . 5. Role of Catecholamines in Stress-Induced

Corticosterone Secretion. . . i. Evidence for an Inhibitory Role.

ii. Evidence for a Stimulatory Role. iii. Role of Epinephrine ....

6. Role of Serotonin in Stress-Induced Corticosterone Secretion. . . . . .

c. Effect of Peripheral Sympathectomy on Stress­Induced Renin and Corticosterone Secretion .

5. Central Nervous System Pathways Involved in the Stress Response . . . a. Sensory Afferents ..... . b. The Limbic System ..... .

1. Septo-Hippocampal System. 2. Amygdala ........ . 3. Hypothalamic Ventromedial Nucleus

c. Role of Brain Catecholamines in Stress D. HYPOTHALAMIC PARAVENTRICULAR NUCLEUS (PVN)

1. Introduction ....... . 2. Morphology ........ . 3. Role of the PVN in Behavior 4. Afferent Connections to the PVN

vi

.29

.30

.31

.32

.35

.35

.35

.35

.36

.37

.38

.38

.38

.39

.40

.40

.41

.42

.45

.45

.47

.47

.47

.48

.so

.51

.51

.53

.57

.. 58

.61

.62

.62

.63

.63

.63

.65

.66

.68

.68

.69

.70

.71

5. Efferent Projections of the PVN ..... 6. Catecholaminergic Innervation of the PVN. 7. Neuroactive Substances in the PVN

a. Neuropeptide Y ........... .

III. METHODS . .

A. Animals B. Surgery C. Electrolytic Destruction of the PVN D. Electrolytic Destruction of the Nucleus Reuniens. E. Microinjection of Ibotenic Acid into the PVN F. Microinjections of 6-0HDA into the PVN. G. Microinjections of Sotalol into the PVN H. Histology ...... . I. Immunocytochemistry ... . J. Stress (CER) Procedure .. . K. Biochemical Determinations.

1. Plasma Renin Activity .. 2. Plasma Renin Concentration 3. Plasma Angiotensinogen 4. Plasma Corticosterone.

L. Statistics . . . . . . . 1. Analysis of Variance . 2. Student-Newman-Keuls' Test

IV. RESULTS ....

A. Electrolytic Lesions in the PVN. B. Electrolytic Lesions in the Nucleus Reuniens c. Injection of Ibotenic Acid into the PVN. D. Injection of 6-0HDA into the PVN E. Injection of Sotalol into the PVN. F. Comparison of Normal Plasma Pools to Experimental

V. DISCUSSION ...

Controls

.73

.74

.74

.75

. 77

. 77

. 77

.78

.78

.79

.80

.81

.82

.82

.83

.85

.85

.87

.88

.88

.89

.90

.91

.93

.93

.98 103 109 116 121

..... 126

A. Overview. . . . . . . . . . . . . . . . . 126 B. Justification of the Stress Paradigm. . . 126

1. Classical Conditioning of Fear Stimuli 127 C. Electrolytic and Ibotenic Acid Lesions in the PVN 129 D. Injection of 6-0HDA into the PVN. . . . . 132 E. Injection of Sotalol into the PVN . . . . 134 F. Hypothetical Organization of Brain Pathways Mediating

Stress-Induced Renin and Corticosterone Secretion . . 139

VI. LITERATURE CITED

vii

LIST OF TABLES

TABLE 1 Comparison of Mean Values: Normal PRA Pool versus Experimental Sham- or Vehicle-Treated Groups . . . . . . 122

TABLE 2 Newman-Keuls' Multiple Range Test for PRA ..... 123

TABLE 3 Comparison of Mean Values: Normal Corticosterone Pool versus Experimental Sham- or Vehicle-Treated Groups ... 124

TABLE 4 Newman-Keuls' Multiple Range Test for Corticosterone . . ....... 125

viii

LIST OF FIGURES

FIGURE 1 Coronal section of rat brain with an intact paraventricular nucleus ....... . . .. 94

FIGURE 2 Coronal section of rat brain with an electrolytic lesion in the paraventricular nucleus. . .. 94

FIGURE 3 The effect of stress on plasma renin activity in rats with electrolytic PVN lesions . . . . . . . . . .95

FIGURE 4 The effect of stress on plasma renin concentration in rats with electrolytic PVN lesions. . . . . . . .96

FIGURE 5 The effect of stress on plasma corticosterone levels in rats with electrolytic PVN lesions. . . . . . . .97

FIGURE 6 Coronal section of rat brain with an electrolytic lesion in the thalamic nucleus reuniens. . . . . . . .. 99

FIGURE 7 The effect of stress on plasma renin activity in rats with electrolytic lesions in the nucleus reuniens . 100

FIGURE 8 The effect of stress on plasma renin concentration in rats with electrolytic lesions in the nucleus reuniens .101

FIGURE 9 The effect of stress on plasma corticosterone levels in rats with electrolytic lesions in the nucleus reuniens. 102

FIGURE 10 Intact paraventricular nucleus .. ..... 104

FIGURE 11 Paraventricular nucleus following bilateral injection of ibotenic acid. . . . . . . . . . . . . . . . . .105

FIGURE 12 The effect of stress on plasma renin activity in rats with ibotenic acid lesions in the PVN. . . . . . .106

FIGURE 13 The effect of stress on plasma renin concentration in rats with ibotenic acid lesions in the PVN ......... 107

FIGURE 14 The effect of stress on plasma corticosterone levels in rats with ibotenic acid lesions in the PVN .108

FIGURE 15A Coronal section from rat injected with vehicle and stained with DBH antiserum. . . . . . . . . . . .111

ix

FIGURE lSB Coronal section from rat injected with 6-0HDA and stained with DBH antiserum, showing loss of DBH immunoreactivity. . . . . . . . . .. 111

FIGURE 16 The effect of stress on plasma renin activity in rats with 6-0HDA lesions . . . . . . . . . . . . . . . . .112

FIGURE 17 The effect of stress on plasma renin concentration in rats with 6-0HDA lesions . . . . . . . . . . . . .113

FIGURE 18 The effect of stress on plasma corticosterone levels in rats with 6-0HDA lesions ................ 114

FIGURE 19 The effect of stress on plasma angiotensinogen levels in rats with 6-0HDA lesions ................ 115

FIGURE 20 Coronal section of the paraventricular nucleus following bilateral injection of sotalol. . . . . . . .. 117

FIGURE 21 The effect of stress on plasma renin activity in rats injected with either vehicle or sotalol, or in non-inj ected rats . . . . . . . . . . . . . . . . . . . . 118

FIGURE 22 The effect of stress on plasma renin concentration in rats injected with either vehicle or sotalol, or in non-injected rats. . . . . . . . . . . . . . . . .119

FIGURE 23 The effect of stress on plasma corticosterone levels in rats injected with either vehicle or sotalol, or in non-injected rats .................. 120

FIGURE 24 Hypothetical brain pathways mediating stress-induced renin and corticosterone secretion ............. 146

x

CHAPTER I

INTRODUCTION

Stress has profound effects on neuroendocrine function,

increasing plasma corticosterone and prolactin levels, and plasma renin

activity (PRA). Neurons in the hypothalamic paraventricular nucleus (PVN)

play an important role as integrators of endocrine, autonomic and

behavioral functions. The purpose of this dissertation research project

was to investigate the role of neurons in, and afferents to, the PVN in

mediating stress-induced neuroendocrine responses, specifically increased

PRA, plasma renin concentration (PRC), and corticosterone levels. Male

Sprague-Dawley rats were used for these experiments. The stressor

consisted of a conditioned emotional (fear) response paradigm, which has

been shown to cause increases in renin and corticosterone secretion. In

this paradigm, each rat receives a footshock 10 minutes following

placement in a chamber, and then was returned to his home cage. This

procedure is repeated for three consecutive days. On the fourth day,

instead of receiving shock 10 minutes after being placed in the test

chamber, the rat is removed and sacrificed by decapitation. Trunk blood

was collected for hormone assays. The experimental control rats were

treated identically to rats in the treatment group with the exception that

shock was not administered at any time.

1

2

1. Effect of Electrolytic PVN Lesions on Stress-Induced Renin and

Corticosterone Secretion. Bilateral electrolytic lesions in the PVN

prevented the stress-induced increase in PRA, PRC and corticosterone

levels. In contrast, electrolytic lesions in the thalamic nucleus

reuniens, dorsal and caudal to the PVN, did not prevent the stress-induced

increase in either PRA, PRC or corticosterone levels. Since the majority

of the diencephalic corticotropin releasing factor (CRF) neurons are

located in the parvocellular subnuclei of the PVN, it is not surprising

that electrolytic PVN lesions prevented the stress-induced increase in

corticosterone secretion. However, the finding that the stress-induced

increases in PRA and PRC were also prevented was unexpected.

2. Effect of Ibotenic Acid Lesions of Cells in the PVN on Stress­

Induced Renin and Corticosterone Secretion. Ibotenic acid is a selective

neurotoxin that destroys cell bodies while leaving fibers of passage

intact. To determine whether cell bodies in the PVN or fibers of passage

mediate stress-induced increases in PRA, PRC and corticosterone levels,

the cells in the PVN were selectively destroyed by injection of ibotenic

acid. The corticosterone, PRA and PRC responses to stress were blocked

by ibotenic acid injection. These observations suggest that neuronal cell

bodies in the PVN mediate the stress-induced increase in PRA, PRC and

corticosterone.

3. Effect of 6-Hydroxydopamine-Induced Lesions in the PVN on

Stress-Induced Renin and Corticosterone Secretion. Since the data

indicated that cell bodies in the PVN mediate stress-induced renin and

3

corticosterone secretion, and the PVN is known to have an extensive

catecholaminergic innervation, the next approach was to determine if

catecholaminergic innervation of PVN neurons plays an important role in

the neuroendocrine response to stress. This was accomplished by injection

of the neurotoxin 6-hydroxydopamine (6-0HDA) into the PVN. 6-0HDA is

taken up by the catecholaminergic nerve endings resulting in degeneration

of these nerve terminals. Lesion placement and damage to noradrenergic

nerve terminals was verified by immunocytochemical methods, using an

antibody against dopamine E-hydroxylase. The 6-0HDA injections prevented

the stress-induced increase in PRA, PRC and corticosterone levels,

suggesting that intact catecholaminergic afferents to the PVN are critical

for the stress-induced increase in renin and corticosterone secretion.

4. Effect of Microinj ection of the E-Adrenoceptor Antagonist

Sotalol into the PVN on Stress-Induced Renin and Corticosterone Secretion.

The results from the experiment with 6-0HDA suggest that catecholaminergic

nerve terminals in the PVN are involved in mediating the effect of stress

on renin and corticosterone secretion. In previous studies, the H.-

adrenoceptor antagonist propranolol attenuated the stress- induced increase

in PRA. Since propranolol is known to cross the blood-brain barrier, it

is possible that &-receptors, located in the PVN, mediate the effect of

stress on renin and corticosterone secretion. To test this hypothesis,

the H.-adrenoceptor antagonist sotalol was microinj ected into the PVN.

Sotalol prevented the stress-induced increase in corticosterone levels,

but did not attenuate the stress-induced increase in PRA or PRC. The

results suggest that a catecholaminergic input to cells in the PVN mediate

4

the effect of stress on renin and corticosterone secretion. Furthermore,

stress-induced increases in corticosterone levels are mediated via &­

receptors, whereas stress-induced increases in PRA and PRC are mediated

by different receptors.

CHAPTER II

LITERATURE REVIEW

A. RENIN

1. Introduction

The renin-angiotensin system plays a role in the maintenance of

sodium balance (Parfrey et al., 1981) and the regulation of blood pressure

and volume (Fujii and Vatner, 1985). Renin is the enzyme responsible for

the conversion of angiotensinogen to angiotensin I (AI), a decapeptide.

AI is further converted to the octapeptide angiotensin II (All) by a

peptidyl dipeptidase, converting enzyme. This reaction is slow in vitro,

but rapid in vivo due to the activity of tissue-bound converting enzyme

present on the luminal aspect of vascular endothelial cells throughout the

body (Caldwell et al., 1976). Renin is the rate-limiting enzyme in the

synthesis of All. All is the physiologically active component of the

system and a potent pressor agent. All causes vasoconstriction by a

direct action on vascular smooth muscle (Fujii and Vatner, 1985) and

indirectly via the sympathetic nervous system (Peach et al., 1971).

2. Description of Renin

Renin is a glycoprotein with a molecular weight of about 40,000

(Galen et al., 1979). Specifically, it is an aspartyl proteinase (Dzau

and Pratt, 1986) that cleaves a leucine-leucine peptide bond between

5

6

residues 10 and 11 of angiotensinogen. Renin has a circulation half­

life of 4-15 minutes (Michelakis and Mizukoshia, 1971; Gordon and Sull­

ivan, 1969; Gutman et al., 1973; Oates et al., 1974; Oparil et al., 1970;

De Vito et al., 1977) and is metabolized by the liver and kidneys.

Inactive forms of renin (prorenin and activation intermediates) have been

found in the plasma of humans (Leckie, 1981), dogs (Wilczynski and Osmond,

1983) and rats (Osmond et al., 1982). Prorenin is a circulating, high

molecular weight form of renin with no intrinsic enzymatic activity

(Sealey et al., 1979; lnagami and Murakami, 1980). Prorenin release

following bilateral nephrectomy was reported to be under F..-receptor

regulation (Wilczynski and Osmond, 1986). Nielsen and Poulsen (1988)

reported that the kidney is the main source of inactive renin.

3. Physiological Aspects of the Angiotensins

a. Angiotensinogen (renin substrate)

The glycoprotein angiotensinogen (approximate molecular weight

60,000) is the substrate for renin. The main source of angiotensinogen

is the liver. lmmunocytochemical studies (Richoux et al., 1983) and

Northern Blot analysis with complementary mRNA sequences for angioten­

sinogen (lngelfinger et al., 1986) have shown that the liver contains and

synthesizes angiotensinogen. Angiotensinogen is widely distributed in the

blood and other extracellular fluids (Reid and Ramsay, 1975; Horky et al.,

1971; Keeton and Campbell, 1980).

Plasma levels of angiotensinogen can be decreased by adrenalectomy

(Carretero and Gross, 1967), hypophysectomy (Goodwin et al., 1970) and

7

diseases that impair liver function (Ayers, 1967). The administration of

glucocorticoids, mineralocorticoids and sodium replacement can prevent the

decrease in angiotensinogen levels produced by adrenalectomy (Nasjletti

and Masson, 1969). Nephrectomy increases plasma levels of angiotensinogen

(Hiwada et al., 1976; Carretero and Gross, 1967; Bing and Poulsen, 1969).

This increase is thought to be mediated by adrenocortical steroids.

Eggena and Barrett (1988) demonstrated an additive effect of nephrectomy

and dexamethasone on the stimulation of renin substrate secretion.

Freeman and Rostorfer (1972) observed that adrenalectomy caused a 70%

reduction of the increase in plasma angiotensinogen caused by nephrectomy.

Surgical stress, which stimulates the release of adrenocortical steroids,

also increases plasma levels of angiotensinogen for several days in

rabbits (Campbell et al., 1973; Romero et al., 1970; Morris et al., 1977).

These studies suggest that an intact pituitary-adrenal axis is necessary

for adequate substrate production.

b. Angiotensin I (AI)

The amino terminal decapeptide AI is cleaved from angiotensinogen

by renin in the blood. Angiotensin I has been reported to act in a

similar manner to All with respect to the adrenal medulla (stimulating

the release of catecholamines) and the central nervous system (inducing

thirst), but with lower potency (Fitzsimons, 1971; Peach, 1974).

c. Angiotensin II (AII)

All has many physiological effects, all of which are involved in

the maintenance of blood pressure and plasma volume. All is a potent

8

pressor agent, acting directly on the vascular smooth muscle to cause

vasoconstriction (Skeggs et al., 1956) and indirectly, via the sympathetic

nervous system (Peach et al., 1971; Scroop et al., 1971). Similarly, All

increases cardiac contractility in vitro by a direct action on the

myocardium and indirectly via norepinephrine release from the cardioac­

celerator nerves (Blumberg et al., 1975). All acts on the peripheral

arterioles to maintain arterial pressure in low cardiac output states

(Davis and Freeman, 1976).

All stimulates the secretion of aldosterone by a direct action on

the zona glomerulosa of the adrenal cortex (Bartter et al., 1961; Biron

et al., 1961). Blockade of All receptors decreases aldosterone production

to nondetectable levels (Davis and Freeman, 1976). All also stimulates

vasopressin secretion when administered intraventricularly in rats and

dogs (Share, 1979; Keil et al., 1975) or intravenously in conscious dogs

(Reid et al., 1982; Bonjour and Malvin, 1970), conscious rats (Knepel and

Meyer, 1980) and humans (Uhlich et al., 1975; Padfield and Morton, 1977).

Aldosterone and vasopressin act at the distal tubules of the kidney to

increase sodium and water reabsorption, respectively. Together, they act

to restore blood pressure and volume in response to the physiological

stress of hypotension, hemorrhage and/or hypovolemia.

All inhibits renin release when it binds to the juxtaglomerular

(JG) cells (Dzau and Pratt, 1986), and increases the concentration of

angiotensinogen in the plasma of dogs (Blair-West et al., 1974; Beaty et

al., 1976) and rats (Nasjletti and Masson, 1973; Khayyall et al., 1973).

All also stimulates secretion of angiotensinogen in vitro from rat

hepatocytes (Klett and Hackenthal, 1987). It is likely that the

9

stimulation of angiotensinogen by All occurs via enhanced glucocorticoid

secretion. All stimulates glucocorticoid secretion by a direct action on

the adrenal cortex (Kaplan, 1965) and indirectly by potentiating the

effect of CRF on ACTH secretion (Keller-Wood et al., 1986; Maran and

Yates, 1977; Spinedi and Negro-Vilar, 1983). Reid et al. (1978) suggested

that All stimulates angiotensinogen production to provide a positive

feedback mechanism to prevent lack of substrate supply during periods of

elevated renin secretion.

4. Regulation of Renin Secretion

a. Introduction

Renin secretion from the kidney is regulated by a number of

neural, chemical and mechanical factors. In general, factors that cause

a decrease in either renal perfusion pressure, blood volume or plasma

sodium concentration will stimulate renin secretion.

Renin is secreted by the granular juxtaglomerular cells located in

the afferent arterioles of the kidney (Taugner et al., 1979; Taugner et

al., 1984). The juxtaglomerular cells are in close proximity to the

macula densa, a specialized portion of the distal tubule that is sensitive

to changes in sodium reabsorption across the tubular epithelium. The

macula densa monitors the ionic environment in the tubular fluid and

relays the message to either increase or decrease renin secretion.

Studies by ltoh et al. (1985) suggest that adenosine, via the activation

of A1 receptors, may be a signal from the macula densa to the juxta­

glomerular cells to inhibit renin release. It is reported that a decrease

in the amount of sodium (Churchill et al., 1978), potassium (Shade et al.,

10

1972) or chloride (Rostand et al., 1985) that crosses the macula densa

cells will stimulate renin secretion.

Renin secretion is also stimulated by a decrease in renal arterial

pressure. The idea that a "pressor substance" (later to be known as

renin) was secreted in response to a decrease in renal blood flow was

first postulated by Goldblatt et al. (1934). This idea was further

developed by Tobian et al. (1959) who suggested the existence of a renal

baroreceptor that was responsible for the inverse relationship between

blood pressure and renin secretion.

Studies by Fahri et al. (1982) have shown that stimulation of renin

secretion following a decrease in renal arterial pressure occurs via an

intrarenal baroreceptor interacting with a renal B-receptor. Their

studies have demonstrated that elevating plasma epinephrine concentration

without changing blood pressure does not increase plasma renin activity.

However, if epinephrine secretion accompanies a stimulus (e.g., hemor­

rhage) which also lowers blood pressure, then renin secretion will be

increased. This system allows for plasma renin activity to be either

stimulated or unaffected by high physiological levels of circulating

epinephrine, depending on whether or not a hypotensive threat is involved.

b. Role of the Sympathetic Nervous System in Renin Secretion

The sympathetic innervation of the rat kidney consists of

sympathetic cell bodies in paravertebral (thoracic T6 through lumbar L4)

and prevertebral (renal, greater splanchnic, and celiac) ganglia

(Sripairoj thikoon and Wyss, 1987). Stimulation of the renal nerves

increases renin secretion (Di Bona, 1985; Blair et al., 1985).

11

This

stimulatory effect is mediated, at least in part, via norepinephrine

released from the endings of the postganglionic sympathetic neurons that

innervate the kidneys (Johnson et al., 1971; Taber et al., 1976).

Circulating epinephrine can also stimulate renin secretion. The effect

of stimulation of the renal nerves can be enhanced by activating beta (,8)

receptors (Vander, 1965). Catecholamines are believed to stimulate renin

secretion via a ,B-adrenergic mechanism since /3-antagonists have been shown

to inhibit their effect both in vivo (Loeffler et al., 1972; Pettinger and

Keeton, 1975) and in vitro (Weinberger et al., 1975; Vandongen et al.,

1973). The renin stimulating potency profile is isoproterenol >

epinephrine > norepinephrine according to studies in conscious dogs by

Johnson et al. (1979). Studies by Healy et al. (1985) and Osborn et al.

(1981) suggest that the sympathetic nervous system stimulates renin

release by activating B1 receptors on the juxtaglomerular cells.

c. Role of Extra-Renal Beta Receptors in Renin Secretion

Although there is evidence supporting intrarenal /3 receptor

regulation of renin secretion, this does not exclude a role of extrarenal

/3-adrenergic receptors that provide a powerful stimulatory effect on renin

secretion. For example, Reid et al. (1972) observed increases in plasma

renin activity and renin secretion following isoproterenol infusion into

the femoral artery of dogs. Renal perfusion pressure was kept constant

with an aortic clamp. However, infusion of the same doses of isoproter­

enol into the renal artery had no effect on plasma renin activity or renin

secretion, suggesting that an extrarenal ,B receptor mechanism is involved

12

in regulating renin secretion. Johnson et al. (1979) obtained similar

results with infusion of epinephrine and norepinephrine into the inferior

vena cava (IVG) and renal artery. Epinephrine infusion into the IVG

resulted in a 3. 5-fold increase in plasma renin activity while IVG

infusion of norepinephrine resulted in a 1.5-fold increase in plasma renin

activity. Infusion of epinephrine or norepinephrine directly into the

renal artery to achieve similar arterial concentrations resulted in no

increases in plasma renin activity. Fahri et al. (1982) suggested that

the effect of epinephrine in the intact conscious dog can be explained as

the combination of an extrarenal and intrarenal action. The extrarenal

event is a generalized vasodilation that results in a drop in blood

pressure which is sensed by the renal baroreceptor. The intrarenal action

is mediated through intrarenal /3 receptors which affect the stimulus-

response curve of the renal baroreceptor. At the present time the

location of the extrarenal /3 receptor has not been established.

d. Role of Brain Serotonin (5-HT) Neurons in Renin Secretion

In 1980, Zimmermann and Ganong demonstrated that injection of the

5-HT precursors 5-hydroxytryptophan (5-HTP) or tryptophan to anesthetized

dogs increased plasma renin activity. In humans, the 5-HT precursor L­

tryptophan was reported to increase renin secretion (Modlinger et al.,

1979), and the 5-HT antagonist cyproheptadine inhibited the secretion of

renin induced by furosemide (Epstein and Hamilton, 1977). In rats, the

5-HT releasers p-chloroamphetamine (PGA) and fenfluramine dose-dependently

increased plasma renin activity via release of 5-HT from serotonergic

nerve terminals (Van de Kar et al., 1981; 1985). The 5-HT2 antagonist

13

LY53857 dose-dependently inhibited the increase in plasma renin activity

and plasma renin concentration following administration of fenfluramine

and MK-212, a 5-HT agonist, suggesting that stimulation of 5-HT2 receptors

enhances renin release (Lorens and Van de Kar, 1987).

The effect of PCA on plasma renin activity was prevented by either

pretreatment with the 5-HT synthesis inhibitor p-chlorophenylalanine

(PCPA) or by a chemical lesion of the dorsal raphe nucleus with the 5-HT

neurotoxin, 5, 7-dihydroxytryptamine (Van de Kar et al., 1982). A

mechanical lesion in the mediobasal hypothalamus, as well as postero­

lateral deafferentation of the hypothalamus (separating the hypothalamus

from the midbrain) also blocked the effect of PCA on plasma renin activity

(Karteszi et al., 1982). Anterolateral section through the retrochias­

matic area was ineffective in blocking the effect of PCA on plasma renin

activity (Karteszi et al., 1982). These results suggest that a serotoner­

gic pathway, with cells in the dorsal raphe nucleus in the midbrain and

terminals in the mediobasal hypothalamus, stimulates renin secretion.

1. Role of Beta Receptors in the Serotonergic

Stimulation of Renin Secretion

There is evidence for f3 receptor involvement in mediating the

serotonergic stimulation of renin secretion. Van de Kar and Richardson

Morton (1986) demonstrated that the B1 and B2 antagonist sotalol and the

B1 antagonist atenolol prevented the increase in plasma renin activity

after the injection of PCA and fenfluramine. These findings supported

those of Alper and Ganong (1984) who demonstrated that pretreatment with

the sotalol and propranolol prevented the increase in plasma renin

14

activity produced by PCA. These studies suggested a possible role for

the sympathetic nervous system in mediating the increase in renin

secretion following the administration of 5-HT releasers.

However, in rats which were adrenal medullectomized and sympathec­

tomized by chronic injections of 6-0HDA, there was no inhibition of the

PCA- and fenfluramine-induced increase in plasma renin activity. Renal

norepinephrine content was reduced to undetectable levels, suggesting that

it is not likely that the serotonergic stimulation of renin secretion is

mediated via the sympathetic nervous system or adrenal catecholamines (Van

de Kar and Richardson Morton, 1986). These results seem to be in

contradiction with the fact that the ~-adrenergic receptor antagonists

sotalol and atenolol prevented the PCA- or fenfluramine-induced increase

in plasma renin activity. However studies by Johnson (1984) and Johnson

et al. (1979), as described earlier, provide evidence that~ receptors

that mediate the increase in plasma renin activity are located outside the

kidney, suggesting an entirely different mechanism of stimulating renin

secretion. In addition, many beta receptor antagonists are known to bind

to 5-HT receptors (Middlemiss et al., 1977) and it is possible that

sotalol and atenolol are present in sufficient concentrations in the brain

to block the 5-HT receptors and inhibit the serotonin-mediated increase

in plasma renin activity (Lemmer et al., 1985; Garvey and Ram, 1975). To

summarize, the data suggest that brain 5-HT neurons regulate renin

secretion via a mechanism that is not related to the autonomic sympathetic

nervous system or adrenal catecholamines.

e. Role of the Hypothalamic Paraventricular Nucleus

in Renin Secretion

15

The PVN plays a role in neurogenic and genetic models of hyperten­

sion (Ciriello et al., 1984; Zhang and Ciriello, 1985). The PVN is a

logical central site to influence renin secretion as it receives afferent

information from the baroreceptor region as well as having efferent

projections to the sympathetic areas of the spinal cord (Sawchenko and

Swanson, 1982).

Electrolytic lesions in the PVN have been shown to prevent the

increase in plasma renin activity produced by the 5-HT releasing drug PCA

as well as preventing the increase in plasma renin activity that follows

immobilization stress (Gotoh et al., 1987). Previous studies (Gotoh et

al., 1985) demonstrated that bilateral knife cuts behind the hypothalamic

paraventricular nucleus (PVN) reduced the renin response to immobilization

and the gravitational stress of head-up tilting, suggesting a role for the

PVN in both stress-induced and 5-HT- stimulated renin secretion.

Electrical stimulation of the PVN in conscious rats increased

plasma renin activity in a frequency-related manner. There were no

significant changes reported in heart rate, mean arterial pressure or

renal vascular resistance at any of the frequencies that produced an

increase in plasma renin activity (Porter, 1986). These results suggest

that neurons in the PVN stimulate renin secretion.

Additional studies by Porter (1988) demonstrated the relationship

between renal perfusion pressure and plasma renin activity during

continuous low-level stimulation of the PVN. He found that for any given

16

decrease in renal perfusion pressure, the plasma renin activity increase

was greater during the ongoing PVN stimulation. The effect of stimula-

tion of the PVN was prevented by both renal denervation and pretreatment

with propranolol. These studies suggest that stimulation of the PVN

elicits increased responsiveness of the kidneys to decreases in renal

perfusion pressure, and that this effect is mediated via the sympathetic

nervous system.

f. Other Hypothalamic Areas Involved in Renin Secretion

The lateral and posterior hypothalamic areas have long been known

as sites where pressor responses can be elicited with electrical

stimulation (Brody et al., 1986; Ciriello and Calaresu, 1977). Therefore

it is not surprising that increases in plasma renin activity are also

observed following electrical stimulation of the lateral (Zanchetti and

Stella, 1975) and posterior (Natcheff et al., 1977) hypothalamus. Unlike

the studies by Porter (1986) involving the PVN, these increases in plasma

renin activity were accompanied by increases in blood pressure and

required intact renal nerves. It is likely that the lateral and posterior

hypothalamus exert their cardiovascular effects through long descending

neuronal projections to the NTS (Kuypers et al., 1973) or ventrolateral

medulla (Brody et al., 1986).

Gotoh et al. (1987) reported that bilateral electrolytic lesions

of the ventromedial nucleus prevented the increase in plasma renin

activity produced by each of the following stimuli: the 5-HT-releaser PCA,

immobilization, head-up tilt under inactin anesthesia, and a low-sodium

diet. The authors hypothesize that the stimuli that increase plasma renin

17

activity via the ventromedial nucleus act by increasing sympathetic

activity since the P-adrenergic receptor antagonist propranolol blocks the

renin response to PCA, immobilization and head-up tilt. The PCA response

is also abolished by another P-adrenergic receptor antagonist, sotalol,

and by ganglionic blockade (Alper and Ganong, 1984). In addition,

stimulation of the ventromedial nucleus increases blood pressure (Bunag

and Inoue, 1985), while bilateral lesions of this nucleus prevent the

development of experimental hypertension (Brody and Johnson, 1980).

g. Extrahypothalamic Pathways Involved in Renin Secretion

Other central nervous system sites play a role in renin release.

Electrical stimulation of the ventrolateral medulla (Richardson et al.,

1974), dorsal medulla (Passo et al., 1971) and mesencephalic dorsal

periaqueductal grey (Ueda et al., 1967) produced increases in plasma renin

activity that were mediated via the renal nerves.

h. Renin and Angiotensin in the Brain

Administration of renin or All into the cerebral ventricles

stimulates vasopressin secretion (Keil et al., 1975), increases drinking

behavior and increases blood pressure (Reid and Ramsay, 1975). The

anterior forebrain and nuclei of the anteroventral third cerebral

ventricle (AV3V) play a role in the All-induced stimulation of drinking

behavior, vasopressin release and increased arterial pressure. These

brain areas are richly innervated by catecholaminergic fibers arising from

cell bodies in the A2 area of the nucleus tractus solitarius, and other

brainstem regions involved with central cardiovascular regulation (Saper

18

et al., 1983; Palkovits et al., 1974; Meldrum et al., 1984). Bellin et

al. (1987) have shown that rats with 6-0HDA injections into the lateral

cerebral ventricle had significantly attenuated drinking and blood

pressure responses following both central and systemic All administration.

Rats pretreated with desmethylimipramine (DMI) prior to injection of 6-

0HDA (to protect noradrenergic terminals) had no thirst or pressor

response deficits following All administration. These results suggest a

role for norepinephrine in mediating All-induced drinking and blood

pressure responses.

The distributions of angiotensin I and II, angiotensinogen,

angiotensin-converting enzyme (ACE) and renin in the brain have been

studied by various biochemical and irnrnunocytochemical techniques

(Brownfield et al., 1982; Changaris et al., 1977; Fuxe et al., 1976; Fuxe

et al., 1980; Basso et al., 1982; Fischer-Ferraro et al., 1971), although

the physiological role for these substances is not clear. Ganong et al.

(1979) suggested that All-like irnrnunoreactive material in the brain may

be involved in modulating the release and action of brain catecholamines.

The presence of a brain renin-angiotensin system is highly

debatable. Fuxe et al. (1980) demonstrated renin-like irnrnunoreactivity

in the paraventricular, periventricular and supraoptic nuclei in the rat

and mouse. Dabsys et al. (1988) demonstrated an increase in angiotensin­

converting enzyme activity levels in the amygdaloid complex following

bilateral destruction of the ventral noradrenergic pathway with 6-0HDA.

There was a concomitant increase in arterial blood pressure that

significantly correlated with the ACE activity levels in the amygdala.

Studies in nephrectomized animals revealed All-like irnrnunoreactivity in

19

brain tissue (Phillips and Stenstrom, 1983) and significant elevations in

angiotensinogen content in the hypothalamus and midbrain (Gregory et al.,

1982). However, the All-like material from rat brain has been shown to

be different from authentic All by both gel filtration (Meyer et al.,

1982) and anion exchange HPLC techniques (Pohl et al., 1988). These

studies indicated that the All-like material has a molecular weight of

5000 7000 (compared to 1046 for All) and is more polar but less

positively charged than All.

Although indirect studies suggest the presence of each of the

components of the renin-angiotensin system in the brain (Printz et al.,

1982) there is no evidence to support a functional brain renin-angiotensin

system. Brownfield et al. (1982) have demonstrated that although

angiotensin and converting enzyme immunoreactivities are present in the

brain, they are not co-distributed, suggesting that there is no brain

pathway for the formation of angiotensin.

i. Role of Brain Catecholamines in Renin Secretion

Norepinephrine in the brain was proposed to inhibit secretion of

renin from the kidney through changes in sympathetic activity and

secretion of vasopressin (Blair et al. , 1977; Reid et al. , 1978). The

antihypertensive drug clonidine is an alpha (a) 2-agonist with a biphasic

effect on blood pressure: a transient increase followed by a prolonged

decrease. lntracerebroventricular injection of clonidine lowers plasma

renin activity at a dose which is ineffective when given intravenously

(Reid et al., 1975). Blair et al. (1977) have demonstrated that

administration of L-DOPA to dogs pretreated with the peripheral decar-

20

boxylase inhibitor carbidopa resulted in a drop in blood pressure and an

inhibition of renin secretion. Since this treatment also increases brain

catecholamine content, this study suggests that catecholamines formed from

L-DOPA can act within the central nervous system to cause a decrease in

renin secretion. In this study the decrease in plasma renin activity was

dependent upon the presence of the renal nerves.

The 5-HT releaser fenfluramine has a biphasic effect on plasma

renin activity: an initial increase followed by a sustained decrease.

Studies by Van de Kar et al. (1985) suggested that the initial increase

in plasma renin activity following fenfluramine administration is mediated

via release of serotonin, while the delayed, long-lasting suppression of

plasma renin activity is mediated via a catecholaminergic mechanism.

However, the long-term suppression of plasma renin activity following

fenfluramine administration is not dependent on the peripheral sympathetic

nervous system, as it was not prevented by adrenal enucleation combined

with 6-0HDA-induced sympathectomy (Van de Kar et al., 1988). To

determine whether a 2 -adrenoceptors mediate the suppressive effects of

fenfluramine, the a 2-antagonists yohimbine and rauwolscine were injected

prior to fenfluramine administration. Neither a 2-antagonist prevented the

long-term suppressive effect of fenfluramine on plasma renin activity or

plasma renin concentration. The data suggest that fenfluramine does not

inhibit renin secretion by activating a 2-adrenoceptors in the CNS. This

is in contrast to clonidine and a-methyl DOPA, which inhibit sympathetic

outflow to the kidneys through activation of central a 2-adrenoceptors.

Privitera et al. (1979) have shown that low doses of intracister­

nally-injected propranolol produced dose-dependent decreases in plasma

21

renin activity and mean arterial pressure (MAP), whereas identical doses

given intravenously had no significant effect on plasma renin activity or

MAP. A concomitant reduction in circulating norepinephrine levels was

also seen. This study suggests that propranolol could be acting at a

central ~-adrenergic site to suppress renin release.

It is clear that central catecholamines play a role in mediating

the suppression of plasma renin activity, but the site(s) of action and

type of receptors involved have yet to be identified.

B. CORTICOSTERONE

1. Introduction

In 1856 Brown-Sequard discovered that the adrenal glands are essential

to life. By the 1930's it became clear that the adrenal cortex is the

source of two vital classes of hormones: one, the glucocorticoids,

regulate carbohydrate and protein metabolism, and the ability to tolerate

prolonged stress (Selye, 1946; Selye, 1943; Cori and Cori, 1927; Britton

and Silvette, 1931; Long et al., 1940). The second class is the

mineralocorticoids, which control water and electolyte metabolism and the

ability of the kidney to reabsorb sodium (Loeb et al., 1933; Harrop et

al., 1933). Aldosterone is the primary salt-retaining hormone of the

adrenal cortex and is synthesized in the glomerulosa layer of cortical

cells. Cortisol (in humans) and corticosterone (in rats) are the main

glucocorticoids secreted by the adrenal gland and are synthesized in the

fasciculata layer. The fasiculata layer of cells in the adrenal cortex

is stimulated by adrenocorticotropin hormone (ACTH) from the anterior lobe

of the pituitary gland (Li et al., 1943; Sayers et al., 1943). Ingle et

22

al., (1938) found that ACTH secretion from the anterior pituitary is

inhibited by secretions of the adrenal cortex. The role of the hypothal­

amus in integrating stimulatory and inhibitory inputs to the adenohypo­

physis was studied by Harris (1948) and Rasmussen (1938). Characteriza­

tion of the hypothalamic factor that stimulates corticotropin secretion

(CRF) was completed by Vale et al. (1981).

2. Corticotropin-Releasing Factor (CRF)

CRF from ovine hypothalamic tissue is a 41-amino acid peptide that

releases ACTH and ~-endorphin from anterior pituitary cells (Vale et al.,

1981; Plotsky, 1985; Rivier et al., 1983; Rivier and Vale, 1983). This

peptide is 5-10 times more potent in releasing ACTH and ~-endorphin than

vasopressin, which was originally thought to act as a corticotropin

releasing factor (Saffran and Schally, 1977). Swanson et al. (1983)

described the localization of CRF immunoreactive cells and fibers in three

distinct systems in the rat brain. The first group is in the hypothalamic

paraventricular nucleus (PVN). Approximately 2,000 CRF-stained cells are

found in each of the eight subdivisions of the PVN, although most (80%)

of the cells are concentrated in the parvocellular division (especially

the periventricular and medial parts), and about 15% are found in the

magnocellular division where oxytocinergic cells predominate. These CRF­

containing neurons are the only ones that demonstrated increased staining

intensity following adrenalectomy. The second group of CRF-containing

neurons are involved in mediating autonomic reponses. They include the

central nucleus of the amygdala, bed nucleus of the stria terminalis,

substantia innominata, lateral hypothalamic area, medial and lateral

23

preoptic areas, central periaqueductal grey, laterodorsal tegmental

nucleus, locus coeruleus, dorsal vagal complex, parabrachial nucleus and

regions containing the Al and AS catecholamine cell groups. The third

group of CRF- immunoreactive cells are found in the cerebral cortex,

especially layers II and III of the neocortex, and limbic regions such as

the cingulate gyrus, prefrontal areas and areas bordering the rhinal

fissure (Swanson et al., 1983). CRF-like immunoreactive neurons were also

observed in the cerebellum and spinal cord (Olschowka et al., 1982).

Outside the central nervous system, CRF-like immunoreactivity has been

found in endocrine cells of the pancreas and gastrointestinal system, also

in the liver, pituitary, adrenal, lung and placenta (Petrusz et al.,

1985).

The parvocellular neurons of the PVN contain CRF (Merchenthaler et

al., 1982; Liposits et al., 1983), which is transported along axons to the

median eminence where it is released into the hypophyseal portal vessels.

Makara et al. (1981) have shown that the CRF fibers in the stalk-median

eminence region either originate from or run through the PVN. In the

pituitary gland, CRF is then bound to specific membrane receptors of ACTH­

containing cells (Wynn et al., 1984; Leroux and Pelletier, 1984). ACTH

release is stimulated which, in turn, increases steroid hormone production

in the adrenal cortex. The glucocorticoids then exert negative feedback

effects on both the hypothalamus (Yates and Urquhart, 1962) and anterior

pituitary (Russell et al., 1969; Mulder and Smelik, 1977) via glucocor­

ticoid receptors (Gustafsson et al., 1983). ACTH immunoreactive fibers

are found in most parts of the parvocellular division of the PVN and in

the oxytocinergic cells of the magnocellular division (Sawchenko et al.,

24

1982; Swanson and Sawchenko, 1983).

Afferent connections to CRF neurons include presynaptic boutons

containing immunoreactive ACTH 1-39 (Liposits et al., 1985) and CRF

(Liposits et al., 1985), suggesting an autoregulatory process in the

control of CRF release. Parvocellular CRF neurons also receive input from

magnocellular neurons in the same nucleus (Leranth et al., 1983) and send

processes to these magnocellular elements (Liposits et al., 1985). CRF

neurons have been shown to have angiotensin II reactivity in adrenalec­

tomized, colchicine-treated rats (Lind et al., 1985).

Rivier and Vale (1983) suggested that vasopressin may act

synergisticallywith CRF under certain physiological conditions to promote

ACTH secretion. In addition, in vitro studies have shown that vasopressin

acts directly on the anterior pituitary corticotropes to stimulate ACTH

secretion (Vale et al., 1983). Together, these results suggest that

vasopressin may act synergistically with CRF under certain physiological

conditions to promote ACTH secretion.

Sawchenko (1987) demonstrated that adrenalectomy (ADX) enhanced

both CRF and vasopressin immunoreactivity suggesting that they may be

colocalized in parvocellular neurosecretory neurons of the PVN. Other

studies support this observation (Tramu et al., 1983; Kiss et al., 1984;

Sawchenko et al., 1984; Piekut and Joseph, 1985). Dual immunocytochemical

studies have shown the co-localization of CRF and vasopressin in the

parvocellular neuronal perikarya in the PVN. In general, CRF immunos­

tained cells are concentrated in the medial parvocellular part of the PVN,

while the vasopressin neurons are predominantly in the magnocellular part

of the PVN. However, in adrenalectomized, colchicine-treated rats, a

25

dense accumulation of vasopressin-imrnunoreactive cells were seen in the

medial parvocellular part of the PVN, in a similar distribution to that

seen for CRF- containing cell bodies (Piekut and Joseph, 1986). Extra­

hypothalamic sites of ADX-enhanced CRF imrnunoreactivity included the

cerebral cortex, arnygdala and the bed nucleus of the stria terminalis.

These extra-hypothalamic sites contained no vasopressin imrnunoreactivity

colocalized with the CRF (Sawchenko, 1987). Hypophysectomy as well as ADX

has been shown to enhance CRF imrnunostaining in PVN cell bodies (Bugnon

et al., 1983; Merchenthaler et al., 1983; Paull and Gibbs, 1983; Swanson

et al., 1983). Sawchenko (1987) reported that hypophysectomy produced

results comparable to ADX-enhanced CRF and vasopressin staining in the

PVN. Low doses of dexamethasone attenuated and high doses prevented the

ADX-induced enhancement of CRF and vasopressin imrnunoreactivity, with the

adrenal steroid potency profile as dexamethasone > corticosterone >

deoxycorticosterone > aldosterone. These results suggest that adrenal

steroids, particularly glucocorticoids, play a primary role as regulators

of peptide expression.

3. Other Factors Affecting ACTH Release

Other neuropeptides that stimulate the release of ACTH include

angiotensin II (All) (Ramsay et al., 1978; Steele et al., 1981) and

oxytocin (Beny and Baertschi, 1980; Pearlmutter et al., 1974). In vitro

dose-response studies by Spinedi and Negro-Vilar (1983) ranked the

corticotropin-releasing activity as follows: CRF > vasopressin > oxytocin

> AII. The stimulatory effects of AII may involve the stimulation of AII

receptors in the anterior pituitary (Hauger et al., 1982; Mukherjee et

26

al., 1982) or the hypothalamic PVN (Saavedra et al., 1986). The density

of the All binding sites in the PVN have been shown to increase following

repeated immobilization stress (Castren and Saavedra, 1988). It is

possible that All receptors in the PVN play a role in modulating stress­

induced corticosterone secretion.

4. Role of Brain Catecholamines in Corticosterone Secretion

a. Introduction

In general, the ascending noradrenergic pathways appear to relay

afferent visceral information to the forebrain (Cunningham and Sawchenko,

1988). There has been considerable controversy as to whether catechol­

amines play an excitatory or inhibitory role in corticosterone secretion.

Norepinephrine and/or epinephrine have been implicated as stimulators of

the secretion of virtually every anterior pituitary hormone (Sawchenko and

Swanson, 1982). With respect to ACTH secretion, brain catecholamines were

originally postulated to have an inhibitory role (van Loon et al., 1971;

Weiner and Ganong, 1978; Cuello et al., 1973; Scapagnini et al., 1970).

However, in recent years more accurate and reliable anatomical and

pharmacological studies suggest a stimulatory role for catecholamines in

corticosterone secretion (Szafarczyk et al., 1985; Feldman et al., 1988;

Alonso et al., 1986; Guillaume et al., 1987).

b. Anatomical Evidence

The predominantly noradrenergic (Sawchenko and Swanson, 1981)

nucleus of the solitary tract (NTS) sends fiber projections to the

parvocellular PVN, substantia innominata, central nucleus of the amygdala

27

and bed nucleus of the stria terminalis (Ricardo and Koh, 1978), all of

which contain CRF-stained cells (Swanson et al., 1983). Swanson et al.

(1983) suggested that the nucleus of the solitary tract, which receives

direct vagal and glossopharyngeal afferents (Beckstead and Norgren, 1979)

could serve as a major relay point of visceral sensory information to CRF

cell bodies in the basal forebrain. Likewise the parabrachial nucleus,

which receives a large input from the NTS (Norgren, 1978), and the locus

coeruleus also project to many of the sites that contain CRF-immunoreac­

tive cell bodies, including the parvocellular PVN, the bed nucleus of the

stria terminalis and the central nucleus of the amygdala. These nuclei

may also be involved in the relay of visceral sensory information to CRF

cell bodies (Swanson et al., 1983).

Another fiber projection is described from the predominantly

noradrenergic cells in the Al region of the ventrolateral medulla (which

receives an input from the NTS) to the magnocellular (vasopressinergic)

and parvocellular PVN. Reciprocal connections have also been described

from the parvocellular PVN to the dorsal vagal complex and spinal cord

(Hosoya and Matsushita, 1979; Swanson and Kuypers, 1980). Electrophysio­

logical studies by Kannan and Yamashita (1985) have also demonstrated

reciprocal connections between neurons in the NTS region and the PVN.

Studies by Skofitsch et al. (1985) have demonstrated binding sites for CRF

in sensory areas of the rat hindbrain and spinal cord, specifically in the

posterior part of the NTS, the substantia gelatinosa nervi trigemini and

laminae I and II of the spinal cord. Since these areas are rich in nerve

endings of primary sensory neurons, this observation suggests a possible

role for CRF in mediating peripheral sensory processes.

28

Cunningham and Sawchenko (1988) used an immunofluorescence double­

labeling procedure to determine which anterogradely labeled fibers and

terminals in the PVN also displayed dopamine P-hydroxylase immunoreac­

tivity, indicating the presence of catecholaminergic neurons. Specifical­

ly, projections from the Al (caudal ventrolateral medulla) region were

found to synapse on magnocellular vasopressinergic neurons in the PVN,

while projections from the A2 (medial part of the NTS) region were located

primarily throughout the parvicellular division of the PVN. The

projections were the most dense in the dorsal medial parvicellular region,

which is known to contain a large population of CRF- immunoreactive

neurons. A less-dense projection was found in the magnocellular division

of the PVN. The A6 (locus coeruleus) projections were found almost

exclusively in the parvicellular division of the PVN, specifically in the

periventricular zone, which contains dopamine, somatostatin and thyrotrop-

in-releasing hormone-containing neurons. These findings provide

anatomical evidence that the noradrenergic cells in the medial NTS

innervate CRF-immunoreactive cells in the PVN.

Liposits et al. (1986) showed that PNMT-immunoreactive axon

terminals have synaptic connections with dendrites, somata and spinous

structures of CRF-immunoreactive neurons. The parvocellular subnuclei

demonstrated a more intense adrenergic innervation than the magnocellular

subnuclei. The mammalian anterior pituitary does not receive catechol­

aminergic innervation, unlike the neural and intermediate lobes which

receive noradrenergic and dopaminergic innervation (Bjorklund et al.,

1967; Saavedra et al., 1975).

c. Role of the Ventral Noradrenergic Bundle in

Corticosterone Secretion

29

The ventral noradrenergic bundle (VNAB) originates from the

medullary Al and A2 cell groups, and the pontine locus coeruleus. It

extensively innervates the hypothalamus, conveying most of the catechol­

aminergic innervation to the PVN (Moore and Bloom, 1979; Palkovits, 1981;

Swanson and Sawchenko, 1980). The VNAB is thought to be the source for

the noradrenergic modulation of ACTH secretion (Szafarczyk et al., 1985;

Feldman et al., 1988). Also, afferents from the epinephrine-containing

cell groups (Cl-C3) reach the hypothalamus via the VNAB (Ungerstedt,

1971).

Injections of 6-0HDA into the VNAB significantly reduced CRF levels

in hypophysial portal blood (HPB) as well as producing significant

depletion of hypothalamic norepinephrine and epinephrine (Guillaume et

al., 1987). Similar results were obtained by Eckland et al. (1988), who

found that 6-0HDA injections into the VNAB or the lateral ventricles

caused a reduction of HPB CRF concentration. However, HPB norepinephrine

was reduced only with the i.c.v. 6-0HDA injections and was unchanged in

the VNAB-lesioned group.

Further support for a stimulatory role of norepinephrine in CRF

secretion is provided by Plotsky (1987). The increase in hypophysial­

portal immunoreactive CRF (irCRF) following electrical stimulation of the

VNAB was prevented by i.c.v. pretreatment with the a 1-adrenoceptor

antagonist corynanthine, but not by the ~-adrenergic antagonist propran-

olol. Corynanthine also blocked the dose-dependent increases in irCRF

30

following icv administration of norepinephrine (0.1-5.0 nmol). In

contrast, 5 nmol or greater doses of norepinephrine resulted in a dose­

dependent inhibition of irCRF release that could be blocked by propran­

olol, but was unaffected by corynanthine. This suggests that low doses

of norepinephrine stimulate CRF release via an a-receptor mechanism,

whereas higher doses of norepinephrine inhibit CRF release via a &­

receptor.

d. Stimulation of CRF/Corticosterone Secretion by Epinephrine

and Norepinephrine

Microinjections of norepinephrine (40 nmoles) into the PVN

significantly increased serum corticosterone levels. Epinephrine (40

nrnoles) microinjections into the PVN were even more efficacious than

norepinephrine in increasing serum corticosterone levels, whereas dopamine

(40 nrnoles) had no effect (Leibowitz et al., 1986).

Intravenous infusion of either !-epinephrine or the P-adrenoceptor

agonist 1-isoproterenol into pentobarbital anesthetized rats resulted in

a dose-dependent increase in plasma corticosterone levels. The effect of

epinephrine and isoproterenol on plasma corticosterone levels was

associated with a parallel and dose-related increase in plasma ACTH

(Berkenbosch et al., 1983).

Tilders et al. (1985) found that intravenous infusion of epineph­

rine increased plasma ACTH and corticosterone to levels that are

comparable to those induced by stress. The stimulation of ACTH secretion

by epinephrine appears to be mediated via CRF and P-adrenergic receptors

since it was prevented by i.p. injections of propranolol or administration

31

of CRF antiserum.

In vitro studies with incubated rat hypothalami demonstrated that

norepinephrine produced a dose-dependent stimulation of CRF release which

was prevented by the ~-adrenoceptor antagonists propranolol and timolol.

The norepinephrine-induced CRF release was not affected by the a1

-

adrenoceptor antagonists thymoxamine, prazosin or corynanthine or the a2-

adrenoceptor antagonist idazoxan (Tsagarakis et al., 1988). This study

suggests a stimulatory role for norepinephrine on CRF release that is

mediated by~ receptors.

e. Role of Dopamine Neurons in Corticosterone Secretion

The PVN is uniformly innervated with dopaminergic neurons. The

dopaminergic innervation of the PVN originates from cells of the arcuate

nucleus (Al2), zona incerta (A13) and periventricular nucleus (Al4)

(Hokfelt et al., 1984). Immunocytochemical studies indicated that both

dopamine and norepinephrine terminals synapatically contact magnocellular

neurons on their dendrites or cell bodies (Decavel et al., 1987). In

addition, tyrosine hydroxylase immunoreactivity has been reported in both

fibers and neurons of the PVN (Swanson et al., 1981; Liposits et al.,

1986).

Dopamine has been shown to stimulate corticosterone secretion.

Injection (i.p.) of the dopamine agonist pergolide caused a dose-dependent

increase in plasma corticosterone levels, which was blocked by pretreament

with the dopamine antagonists spiperone and haloperidol. Spiperone also

partially inhibited the increase in plasma corticosterone levels following

administration of quipazine, a serotonin agonist, although much higher

32

doses were required than for pergolide. In addition, the peripheral

dopamine antagonist domperidone did not prevent the increase in plasma

corticosterone levels by pergolide (Fuller and Snoddy, 1981). These

results support a role for central dopamine receptors in corticosterone

secretion. Similar results were obtained by Jezova et al (1985), who

found that subcutaneous injection of the dopamine agonist apomorphine

increased plasma ACTH and corticosterone levels. Haloperidol pretreatment

completely inhibited the apomorphine-induced stimulation in plasma ACTH.

5. Role of 5-HT Neurons in Corticosterone Secretion

Many studies support the hypothesis that stimulation of serotonergic

receptors leads to increased corticosterone secretion in unstressed

animals (Buckingham and Hodges, 1979; Fuller and Snoddy, 1980; Fuller,

1981; Jones et al., 1976; Van de Kar et al., 1982). Gibbs and Vale (1983)

reported that administration of the 5-HT uptake inhibitor fluoxetine to

anesthetized rats caused increased release of CRF-41 into hypophyseal

portal blood. Lorens and Van de Kar (1987) demonstrated that plasma

corticosterone levels were increased by the 5-HT agonist MK-212, the 5-

HT1A agonists ipsapirone and 8-hydroxy-2-(di-N-propylamino)tetralin (8-0H-

DPAT) , and the serotonin releaser fenfluramine. The 5-HT2 antagonist

LY53857 did not prevent the effect of MK-212 and fenfluramine on

corticosterone secretion. These data suggest a possible role for 5-HT1A

receptors in the stimulation of corticosterone secretion. This conclusion

was confirmed by Koenig et al. (1987) and Gilbert et al. (1988).

The dorsal and median raphe nuclei in the midbrain are known to be

the major sources of serotonergic pathways to the hypothalamus (Azmitia

33

and Segal, 1978; Parent et al., 1981; Van de Kar et al., 1980). In

particular, the median raphe nuclei are the primary source of 5-HT fibers

to the suprachiasmatic nucleus, anterior hypothalamic area and medial

preoptic area, while the anterolateral hypothalamic area and arcuate

nucleus receive 5-HT inputs from both the dorsal and median raphe nuclei

(Van de Kar and Lorens, 1979).

With respect to the PVN, a relatively light serotonergic innerva­

tion has been described (Aghajanian et al., 1973; Steinbusch, 1981).

Immunocytochemical studies by Sawchenko et al. (1983) found that the major

source of ascending serotonergic projections to the PVN were from the

dorsal and median raphe nuclei and the medial lemniscus. A significant

decrease in PVN serotonin was reported following surgical destruction of

the dorsal raphe nucleus (Palkovits et al., 1977).

Immunocytochemical studies by Liposits (1987) demonstrated a

moderate serotonergic innervation to the PVN, with a prominent distribu­

tion to the parvocellular subnuclei. Serotonergic axons were found to

overlap with CRF-immunoreactive axons. Further ultrastructural examina­

tion revealed that serotonin-containing terminals formed axo-dendritic and

axo-somatic synapses with CRF-immunoreactive neurons. This study suggests

that serotonin can influence CRF-containing neurons via synaptic

transmission.

Midline injections of 5,7-dihydroxytryptamine (5,7-DHT) into the

raphe nuclei produced significant depletions of 5-HT in the hypothalamus,

and prevented the increase in corticosterone following stimulation of

cortical and limbic areas (dorsal hippocampus, basolateral amygdala,

mesencephalic reticular formation and medial septal nucleus). The ether-

34

induced increase in corticosterone levels was not affected by this lesion.

Injection of 5, 7-DHT into the PVN also prevented the adrenocortical

response to dorsal hippocampal and photic stimulation. These results

suggest that 5-HT in the PVN mediates the stimulation of corticosterone

secretion following either peripheral or central stimuli (Feldman et al.,

1987).

In contrast, Van de Kar et al. (1982) demonstrated that the PCA­

induced stimulation of corticosterone secretion was not mediated by either

the dorsal or median raphe nucleus. In addition, posterolateral cuts

which interrupted caudal inputs to the hypothalamus did not block the

serotonergic stimulation of corticosterone secretion by the serotonin­

releasing drug PCA (Van de Kar et al., 1985).

Serotonin has been shown to play a role in mediating the diurnal

variation in basal plasma corticosterone in rats. Plasma corticosterone

levels peak at the onset of the dark cycle in the rat (Krieger and Hauser,

1978; Wilkinson et al., 1979). Depletion of brain serotonin with p-

chlorophenylalanine (PCPA) raised the morning low and prevented the

evening increase in plasma corticosterone (Scapagnini et al., 1971;

Vernikos-Danellis et al., 1973). Similar results were observed with rats

fed a tryptophan-free diet for 21 days (Vernikos-Danellis et al., 1977).

In humans, plasma cortisol and ACTH levels also follow a diurnal rhythm,

except that the highest levels are found in the morning (0600h), with the

lowest values at night (2200h) (Watabe et al., 1987). This study

demonstrated a similar diurnal pattern for plasma CRF, with peak values

at 0600h and significantly lower levels at 1800h and 2200h. These results

suggest that the diurnal rhythm in CRF secretion plays a role in

35

regulating the diurnal rhythm in cortisol and ACTH secretion.

In conclusion, the studies presented in this section demonstrate the

presence of CRF cell bodies in the PVN, and a role for brain catechol­

amines in the stimulation of CRF secretion.

C. STRESS

1. Introduction

a. Definition

Stress has been defined by Hans Selye (1973) as the "non-specific

response of the organism to any demand made upon it" (Selye, 1973). By

this definition, the word "stress" describes a response, but it is often

used interchangeably with the stimulus or "stressor" that elicits the

response (Mills, 1985). The stress response can occur following a wide

variety of stimuli. Stressors can be classified as physiological,

environmental or psychological. Physiological stressors include surgery,

injection of foreign substances such as proteins, diseases, anesthesia,

loss of blood, exercise and trauma. Environmental stressors include

prolonged exposure to cold or heat, noise, radiation, pollutants and

chemicals. Psychological stressors include a threatening predator,

intense competition among members of the same species, prolonged conflict,

learning how to avoid a painful stimulus (such as electric shock),

conditions characterized by novelty, anticipation, fear, unpredictability

and change (Kidman, 1984; Gray, 1971; Mills, 1985).

b. Cannon's Alarm Reaction

Cannon (1914) described the role of the sympathetic nervous system

36

in the "fight or flight" emergency or "alarm reaction". This reaction

permits the body to perform more strenuous muscle activity than normally

possible. Cannon noted that when animals underwent stressful situations

there was a marked increase in the activity of sympathetic nerves. This

stimulation of the autonomic nervous system mobilizes the body's resources

to deal with the stressful situation. The alarm reaction is thought to

be transmitted from the hypothalamus through the reticular formation to

the spinal cord to evoke massive catecholamine release from the sym­

pathetic nerve endings (Guyton, 1981).

c. Selye's General Adaptation Syndrome

Selye (1936) was the first to describe a biological stress syndrome

produced by diverse noxious agents. His description subsequently became

known as the "general adaptation syndrome", a three-stage process that

describes the long-term adjustments elicited by prolonged stress. The

first stage is the "alarm reaction", when the body is exposed to diverse

stimuli to which it has not adapted. Following the initial alarm

reaction, a "stage of resistance" occurs in which resistance to the

original stressor (which continues to exist) is increased, but resistance

to other stressors is decreased. If the stressor continues for a long

enough time, the stage of resistance develops into a final "stage of

exhaustion", with a catastrophic decline in resistance to all other types

of stressors. If the stress reaction continues undiminished, death ensues

(Selye, 1976).

In cases of prolonged stress, Selye (1946) described a major role

for glucocorticoids during the alarm reaction. Glucocorticoids increase

37

sugar deposition in the liver and facilitate the conversion of proteins

and fats into sugars. They enhance the responsiveness of the blood

vessels to epinephrine and norepinephrine. This will provide the body

with energy sources to be rapidly utilized in the event of sudden

strenuous activity. Glucocorticoids also lower the body's resistance to

infection and have anti-inflammatory properties, including decreased

numbers of lymphocytes and eosinophils, decreased thymus weight and

inhibition of antibody formation. Stress appears to play a major role as

a suppressor of immune function.

Other physiological changes induced by prolonged stress include a

depression of thyroid activity, inhibition of body growth, and suppression

of sexual and reproductive behavior (Gray, 1971). In simple terms, the

body copes with prolonged stress by closing down any process that detracts

from immediate energy mobilization.

d. Sympathetic and Endocrine Stress Responses

The biological stress response is believed to be mediated by the

stimulation of two systems (Kidman, 1984). One, the pituitary-adrenal

axis, is activated by sensory nerves which relay information regarding

external stressors to the hypothalamus. Hypothalamic neurons contain CRF

and release it into the portal circulation, stimulating the anterior

pituitary gland to produce and release ACTH. In response, the adrenal

cortex produces corticosteroids. The glucocorticoids liberated into the

bloodstream inhibit the synthesis and/or release of ACTH from the

pituitary by a negative feedback mechanism.

The second system is the sympathetic adrenal medullary axis, which

38

plays an important role in the alarm reaction described by Cannon. In

this system, once the sensory nerves relay the information regarding the

external stressors to the hypothalamus, it relays a message through the

reticular formation to the spinal cord to evoke massive catecholamine

release from the sympathetic nerve endings and adrenal medulla.

e. Physiological Changes in Response to Stress

Selye (1950) emphasized the fact that the physiological changes

(primarily in the adrenal cortex) in response to stressful stimuli were

the same irrespective of the stimulus used. This suggests that whether

the stressor is physiological, environmental or psychological, the

biological stress response would always be the same. This general

hypothesis allows for comparison of physiological responses following a

variety of different stressors.

Since stress is abstractly defined, it is helpful to have

quantifiable parameters that are known to change when animals or humans

are exposed to a stressor. In the experiments presented in this

dissertation, the neuroendocrine responses to stress were measured with

respect to plasma renin activity and plasma corticosterone levels. The

stressor was a behavioral paradigm known as conditioned fear, or

conditioned emotional response (CER), designed to produce psychological

stress.

B. The Conditioned Fear Paradigm

a. Description

Each rat was transported from his home cage to the stress chamber.

39

Ten minutes following placement in the chamber, the experimental

(stressed) rat received an inescapable foot shock and was returned to his

home cage. This procedure was repeated for three consecutive days. In

previous experiments it was apparent that the stressed rats had learned

that their placement in the chamber will be followed by a foot shock. By

the third day, in contrast to the control animals, the stressed rats

defecated, urinated and demonstrated freezing (no movement) behavior. On

the fourth day, after 10 minutes in the test chamber, instead of receiving

shock the rat was removed, transported to a third area outside the stress

room and immediately sacrificed by decapitation. The control rats were

treated identically except that shock was not administered at any time.

A more detailed description of this procedure can be found in the Methods

section.

b. Pavlov's Dog: Classical Conditioning

Ivan Pavlov (1927) presented studies on classical conditioning in

which a dog was repeatedly exposed to a situation where an initially

neutral stimulus (e.g., a tone or light flash) was followed directly by

a stimulus of biological significance to the animal (food). The latter

"unconditioned stimulus" (UCS) elicited a range of "unconditioned

responses" (UCR) . As the pairings of stimuli were continued, the

originally neutral stimulus became a "conditioned stimulus" (CS), capable

of eliciting some part of the responses formerly elicited exclusively by

the UCS. This "conditioned response" (CR) in Pavlov's dogs was saliva­

tion.

40

c. Definition of Fear

Fear can be defined as an emotional response to an aversive

stimulus. In this respect, it is classified as an emotion. The common

element that emotions share is that they represent a reaction to a

"reinforcing event" or signals of reinforcing events. Reinforcing events

include both rewards and punishments. An example of a punishment would

be the removal of a reward, or the failure of an expected reward to occur.

The emotional responses will depend on the type of reinforcing event and

the animal's knowledge of these events (Gray, 1971).

d. Measures of Fear

Hall (1941) established the link between the study of fearfulness

and the use of defecation as a measure of fear. He developed a standard

test (the Open Field) for the measurement of emotional defecation. Parker

(1939) demonstrated that defecation occured in response to a variety of

situations which one would expect to elicit fear. He recorded the amount

of defecation elicited by six different situations: the Open Field test,

a buzzer, sudden dropping, a tilting box that caused the rat to slide down

an inclined plane, immobilization and forced swimming. High positive

correlations were obtained between the scores on every test. Further

factor analysis indicated a general factor (presumably fear) which all the

tests were measuring. These results support the statement by Selye (1950)

that physiological changes in response to stressful stimuli are the same

irrespective of the alarming stimulus used.

41

e. Emotional Factors

There are physiological events that occur within an animal to

influence its behavior in a given environment. In 1890, William James

and Carl Lange independently expressed a theory for the interrelationship

between physiological events and emotional responses. The James-Lange

theory, as it is known, defines emotions as the perception of the

physiological changes that occurred in response to a stimulus. For

example, the feeling of being frightened would consist solely of feeling

your heart pound, your breathing become faster, and so on. Unfortunately,

it is not possible to isolate specific physiological functions and

correlate them exclusively with a particular emotional state. However,

the role of emotional factors should not be overlooked. Studies by Mason

et al. (1976) recognize the sensitivity of the pituitary-adrenal cortical

system to psychological influences. Departing from Selye's non-specific

physiological response hypothesis (1950), Mason et al. (1976) proposed

that there are similar psychological reactions to various external

stimuli. Factors such as heat, cold, fasting or exercise do not produce

an adrenal cortical response if emotional arousal is carefully avoided

(Mason et al., 1976). To illustrate, on the first day of starvation,

food-deprived monkeys that could watch other monkeys being fed in nearby

cages had an elevation of blood corticosteroids that occurred too early

to be attributed to a physiological response due to a lack of food.

However when the monkeys were carefully adapted to their surroundings,

kept in a soundproof room, and maintained on regular "feeding" (non­

caloric, fruit-flavored pellets) schedules, there were none of the usual

42

changes in glucocorticosteroids associated with fasting periods. Mason

concluded that there is a considerable degree of emotional disturbance,

due to discomfort, inherent in the laboratory setting. Kidman (1984)

expanded this idea to a "cognitive theory of stress" which says that, in

spite of our best efforts to control hunger, temperature and other

environmental pressures, we still become frustrated over perceived threats

to our welfare and security. Thus, the stress response consists of a

variety of physiological and psychological factors. These factors are

most likely integrated through the central nervous system, in particular

the hypothalamus, as it coordinates the interaction between the autonomic

nervous system, the endocrine system and many forms of emotional behavior.

3. Cardiovascular Changes in Response to Classical Conditioning

Stressful conditions have been implicated in the etiology of

cardiovascular disease and hypertension (Hollenberg et al., 1981; Manuck

et al., 1981; Schulte et al., 1984). Changes in heart rate and arterial

blood pressure can be classically conditioned (Dykman and Gantt, 1956).

A neutral stimulus that elicits no cardiovascular response (conditioned

stimulus) is paired with another stimulus (unconditioned stimulus) that

evokes a vigorous cardiovascular response (unconditioned response). After

several pairings of the conditioned and unconditioned stimulus, the

previously neutral conditioned stimulus will eventually cause cardiovas­

cular changes in anticipation of the unconditioned stimulus (Cohen and

Randall, 1984).

Smith et al. (1984) have demonstrated that the perifornical area of

the hypothalamus (referred to as HAGER, the Hypothalamic Area Controlling

43

Emotional Responses) is responsible for the cardiovascular responses that

accompany emotional behavior in monkeys. Their operational definition of

emotion included cardiovascular responses (elevated arterial blood

pressure) and a behavioral response (suppression of lever pressing). 'When

the hypothalamic area (HAGER) was electrically stimulated, it produced the

same cardiovascular pattern as that associated with the emotional

situation. The cardiovascular but not the behavioral portion of the

emotional response was lost following ablation of the HAGER. Afferents

to the HAGER include the amygdala and septal area as well as brain stem

autonomic areas such as the central grey, nucleus tractus solitarius, and

dorsal motor nucleus of the vagus. Efferents include a direct projection

to the intermediolateral cell column cells of the thoracic cord, an

important site with respect to sympathetic control of cardiovascular

function. Other efferent projection sites include the raphe nuclei, the

parabrachial complex, the hypothalamic PVN and the periaqueductal grey.

Cardiovascular changes have been observed with the Sidman avoidance

behavioral testing (Galosy et al., 1982). The Sidman avoidance paradigm

(Sidman, 1953) is used frequently to evaluate cardiovascular changes

associated with the postponement of electric shock contingent upon an

appropriate reponse. The paradigm involves delivery of electric shock at

regular intervals unless a response (such as lever pressing) occurs to

delay the shock. If the animal presses the lever at least once during the

interval, it can avoid all shocks (Turkkan et al., 1982).

Anderson and Brady (1971, 1972) observed decreased heart rate and

increased pulse and arterial blood pressure during the anticipatory pre­

avoidance period. Once the avoidance test was underway, the heart rate

44

rapidly increased above pre-avoidance levels while pulse and arterial

blood pressure remained elevated. Further pharmacological studies

indicated that the ~-adrenoceptor blocker propranolol partially attenuated

the heart rate elevations during avoidance, but did not affect the

increases in blood pressure in either avoidance or pre-avoidance periods

(Anderson and Brady, 1976). In similar studies, the a-antagonist

phenoxybenzamine blocked the elevation in blood pressure during the

avoidance period but did not affect pre-avoidance changes in heart rate

and arterial blood pressure (Anderson et al., 1976). From these and other

studies, Galosy et al. (1982) suggested that reduced heart rate during the

pre-avoidance period is dominated by parasympathetic mechanisms, whereas

the arterial blood pressure levels may be mediated by peripheral vascular

mechanisms which do not involve a receptors, perhaps cholinergically-

mediated shifts in blood flow (Anderson et al., 1976). During the

avoidance period, a combination of increased sympathetic and decreased

parasympathetic activity leads to increased heart rates, while elevated

arterial pressure is due to continued sympathetic peripheral vasoconstric­

tion (Galosy et al., 1982).

In humans, the stress response, as elicited by a demanding

arithmetic test (Brod et al., 1959), consisted of the following:

vasodilatation in skeletal muscle, coronary vessels and brain, while most

other resistance vessels and veins underwent a neurogenic vasoconstric­

tion. Increased sympathetic and decreased parasympathetic influences on

the heart resulted in tachycardia, increased cardiac ouput and slightly

increased mean arterial pressure. Neuroendocrine responses included

increased epinephrine, ACTH, corticosteroids, vasopressin and renin-

45

angiotensin-aldosterone by means of sympathetic nerve stimulation of the

kidney (Brod, 1970). Brod et al. (1962) demonstrated that this cir­

culatory pattern exists in hypertension, especially the early stages.

Eliasson (1984) observed that the limbic-hypothalamic pattern of stress

responses is nearly identical in animals and man. He proposed that modern

humans develop hypertension because they are socially deprived of the

acting-out behavior which is a large part of the stress or defense

reaction. However, the autonomic and hormonal changes associated with

the stress reaction are not suppressible, therefore hypertension results.

4. Neuroendocrine Changes in Response to Stress

a. Stress-Induced Renin Secretion

The physiological response to stress involves the release of

catecholamines from the adrenal medulla and sympathetic nerve endings.

Stimulation of the renal nerves increases renin secretion (Di Bona, 1985;

Blair et al., 1985), and this stimulatory effect is mediated, at least in

part, via norepinephrine released at the endings of the postganglionic

sympathetic neurons that innervate the kidneys (Davis and Freeman, 1976;

Reid et al, 1978; Ganong and Barbieri, 1982). Circulating epinephrine and

norepinephrine can also stimulate renin secretion.

Many investigators have reported increases in plasma renin activity

in response to a variety of stressors. Clamage et al. (1976) reported

increases in plasma renin activity in rats following 30 minutes exposure

to either novel "open field" environment or the presence of a hungry cat.

Jindra et al. (1980) observed increases in plasma renin activity after

acute and repeated immobilization stress. Paris et al. (1987) reported

46

increased plasma renin activity by the following stressors: 12 minute

conditioned fear paradigm; 20 minute immobilization; 20 minute forced

swimming in cold water; 2, 12 and 22 minutes of intermittent footshock;

and a 3 minute conditioned fear paradigm with 0, 10, 30 and 60 minute

recovery. Blair et al. (1976) reported an elevation of plasma renin

activity in baboons during avoidance performance. Gotoh et al. (1985)

demonstrated that bilateral knife cuts behind the hypothalamic paraven­

tricular nucleus (PVN) reduced the renin response to immobilization and

the gravitational stress of head-up tilting. Plasma renin activity

clearly can be increased by many different types of stressors.

Sigg et al. (1978) found that restraint causes an increase in plasma

renin activity which was not blocked by pretreatment with the &-adrenergic

antagonists dl-propranolol or sotalol. However, propranolol pretreatment

prevented the increase in plasma renin activity following swim stress and

ether stress (Pettinger et al., 1972) and intermittent footshock (Leenen

and Shapiro, 1974). Also, Van de Kar et al. (1985) have shown that

propranolol significantly inhibits the increase in plasma renin activity

following a three-minute conditioned fear paradigm. Evidence for a role

of central catecholamines in stress-induced renin secretion was provided

by Lightman et al. (1984) who demonstrated that injections of the

catecholamine neurotoxin 6-0HDA into the dorsal noradrenergic bundle

markedly attenuated the increase in plasma renin activity in response to

hemorrhage. Kobayashi et al. (1975) have reported that afferent fibers

from the locus coeruleus to the PVN travel in the dorsal noradrenergic

bundle. These studies suggest that central catecholaminergic projections

to the PVN may be involved in the renin response to stress.

b. Stress-Induced Corticosterone Secretion

1. Introduction

47

Corticosterone levels have been shown to increase in response to

several stressors. For example, Armario et al. (1986) demonstrated that

the corticosterone response to stress is directly related to the intensity

of the stress stimulus. The stressors used in their study were graded

levels of low intensity stressors (either transfer to a novel cage or

room, or a loud noise). These results are in agreement with findings by

Kant et al. (1983), who demonstrated that increases in plasma cor­

ticosterone levels were related to the intensity of footshock. They found

that corticosterone secretion was very sensitive not only to an obvious

stressor (footshock) but also to any change in the environment accompany­

ing the procedure. Simply placing the rat in the shock box without

delivery of any current elicited a strong corticosterone response. These

studies suggest that the increase in corticosterone levels in response to

stress can represent a sensitive index of the intensity of emotional

arousal experienced by the animals.

Dohanics et al. (1986) found that the adrenocortical response to

ether stress occurs independently of neurohypophyseal function, and that

the full corticosterone and ACTH response to ether or ether plus

laparotomy stress requires not only an intact PVN but also an intact

neurointermediate lobe.

2. Stress-Induced Secretion of CRF

CRF has been shown to mediate stress-induced ACTH release in rats

48

(Rivier and Vale, 1983). CRF also has been shown to initiate other

biological actions in response to stress. For example, injections of CRF

into the cerebral ventricles produced significant increases in plasma

concentrations of epinephrine, norepinephrine, glucagon and glucose,

elevated mean arterial pressure and heart rate, and inhibited gastric acid

secretion. Pretreatment with the ganglionic blocking drug chlorisondamine

completely abolished the increases in plasma catecholamines and glucose,

heart rate and blood pressure (Lenz et al., 1987). CRF has been shown to

activate brain catecholamine systems. Studies by Dunn and Berridge (1987)

demonstrated that injections of CRF into the cerebral ventricles of mice

resulted in increased dihydroxyphenylacetic acid (DOPAC):dopamine ratios

in prefrontal cortex, septum, hypothalamus and brain stem compared to

saline controls. The ratio of 3-methoxy-4-hydroxyphenylethyleneglycol

(MHPG):norepinephrine was increased in the prefrontal cortex and

hypothalamus. DOPAC and MHPG are the major catabolites of dopamine and

norepinephrine, respectively. CRF release has been measured in conscious,

unrestrained rats using a push-pull cannula system implanted in the median

eminence (Ixart et al., 1987). Basal CRF release occurs in a pulsatile

pattern with peaks about every 45 minutes (peak mean 9.0 ± 0.7 pg vs.

trough 4.1 ± 0.3 pg). Ether elicited an increase in CRF which lasted

about 45 minutes with peak values of 54.3 ± 3.2 pg.

3. Stress-Induced Secretion of ACTH

Gibbs (1984) found that ACTH secretion increased in response to

restraint, cold or ether stress. However, Armario et al. (1985) have

found dissociations between ACTH and corticosterone responses: the ACTH

49

response to restraint stress, unlike the corticosterone response, was the

same in control and in chronically stressed rats. This suggests that the

increased corticosterone levels from restraint stress are not necessarily

due only to increased pituitary ACTH release.

The increase in ACTH following hemorrhage was prevented by PVN

lesions, while the responses of renin, epinephrine, norepinephrine, mean

arterial blood pressure and heart rate to hemorrhage were not affected

(Darlington et al., 1988). However, these lesions consisted of creating

an island of the PVN and dorsal hypothalamus by lowering a triangular

(Halasz-type) knife into the third ventricle to the level of the PVN and

then rotating. This may have disrupted fiber pathways to and from the

PVN without actual damage to PVN cell bodies.

Makara et al. (1986) proposed that ACTH hypersecretion following

adrenalectomy is driven primarily by CRF- and/or vasopressin-producing

neurons in and around the PVN, while either (a) other sources of CRF-41,

(b) increased pituitary sensitivity, or (c) other hypothalamic factors

may restore the stress-induced ACTH release in the absence of the PVN.

The effect of bilateral lesions of the PVN on circadian rhythm and

ether stress responsiveness of plasma ACTH and corticosterone were studied

in chronically cannulated female Sprague-Dawley rats. Three weeks after

PVN lesions, ACTH values had dropped by at least half, and the diurnal

maximum and amplitude for corticosterone were diminished by one third.

The peaks for both hormones following ether stress were only half as high

as the control peaks (Ixart et al., 1982). These results support the role

of the PVN as a mediator of stress- induced ACTH and corticosterone

secretion.

50

4. Role of Vasopressin in Stress-Induced Corticosterone Secretion

Vasopressin promotes the retention of water by the kidney, to

maintain plasma volume. There are two major vasopressin systems in the

hypothalamus. One system has cell bodies in the PVN and supraoptic

nuclei, projecting through the internal zone of the median eminence to

the pars neuralis of the pituitary gland. The other system consists of

cell bodies in the PVN with nerve terminals in the median eminence

(Silverman and Zimmerman, 1983). One difference between these systems is

that adrenalectomy resulted in increased staining intensity of vasopressin

immunoreactive nerve fibers in the external zone of the median eminence,

whereas the fibers of the neural lobe or intenal zone of the median

eminence were not affected (Zimmerman et al., 1977).

Vasopressin has been shown by Tilders et al. (1985) to have an

important role in the stress-induced activation of the pituitary-adrenal

system, possibly by potentiating the effects of CRF. Pretreatment of rats

with an antiserum to vasopressin inhibited the ACTH response to both

restraint and formalin stress. Vasopressin has also been shown to

potentiate the activity of CRF (Gillies et al., 1982). Vasopressin

secretion is stimulated by hemorrhage (Robertson, 1977), hypoglycemia

(Bayliss and Robertson, 1980), electric shock (Husain et al., 1979) and

ether (Gibbs, 1984).

Further evidence for the involvement of vasopressin in the ACTH

response to stress is that the vasopressin antagonist dpTyr(Me)arginine

vasopressin decreases ether stress-induced ACTH release, and administra­

tion of both the vasopressin antagonist and CRF antiserum completely

51

prevented stress-induced ACTH secretion (Rivier and Vale, 1983). However,

some studies reported that dpTyr(Me)arginine vasopressin does not block

either the intrinsic ACTH-releasing activity or its synergistic effects

with CRF (Baertschi et al., 1984; Knepel et al., 1984) and, therefore,

should be avoided in studies involving ACTH release (Tilders et al.,

1985).

5. Role of Catecholamines in Stress-Induced Corticosterone

Secretion

Controversy still exists as to whether catecholamines play an

excitatory or inhibitory role in stress-induced corticosterone secretion.

i. Evidence for an Inhibitory Role of Catecholamines

Brain norepinephrine has been postulated to inhibit ACTH secretion

(van Loon et al., 1971; Weiner and Ganong, 1978). Eisenberg (1975)

administered reserpine (systemically) or phenoxybenzamine (centrally) and

found an augmentation of the plasma corticosterone response to ether.

This suggests that norepinephrine may function in the central nervous

system to inhibit the stress response. However, reserpine is not a

selective noradrenergic depletor as it will also deplete serotonin stores

(Bugnon et al., 1983).

Suemaru et al. (1985) observed significant increases in serum ACTH

and corticosteorne levels 15 minutes after the onset of either ether­

laparotomy or cold-restraint stress. The hypothalamic CRF concentration

showed a rapid increase 2.5 minutes following ether-laparotomy stress,

followed by a significant decrease after 15 minutes, and a subsequent

52

increase. Norepinephrine concentration in the hypothalamus was reduced,

whereas dopamine concentration was increased. Epinephrine concentrations

did not show a significant change throughout the stress procedure. From

these observations they concluded that norepinephrine in the hypothalamus

may not be involved in stimulating hypothalamic CRF secretion in the early

phase of acute stress.

Glavin et al. (1985) depleted brain norepinephrine in rats by

injecting the dopamine-E.-hydroxylase inhibitor FI.A-63 (to prevent the

conversion of dopamine to norepinephrine) followed by R04-1284, a

reserpine-like vesicular depletor of norepinephrine, dopamine and

serotonin. This treatment produced significant depletion of norepineph-

rine in all brain regions examined. Norepinephrine depletion did not

prevent the increase in plasma corticosterone levels following cold-

restraint stress. However, in the non- stressed rats, there was no

significant difference in plasma corticosterone levels between the

norepinephrine-depleted and saline-treated rats. The authors suggest that

norepinephrine does not mediate cold restraint stress-induced increases

in plasma corticosterone levels, nor does it play an inhibitory role with

respect to corticosterone secretion. However, this study combines two

non-related stimuli (cold vs. immobilization) as stressors without

addressing how this could affect hormonal responses.

Reinstein et al. (1985) have demonstrated that acute, uncontroll­

able stress (restraint and intermittent footshock for one hour) depleted

norepinephrine in the rat brain (increased norepinephrine turnover).

Pretreatment with supplemental tyrosine (to prevent norepinephrine

depletion) prevented hypothalamic norepinephrine depletion and suppressed

53

the increase in plasma corticosterone, suggesting an inhibitory role for

norepinephrine. However, these corticosterone data were reported as

percent of non-stressed control values, which were quite high (37 .4

ug/dl). In conclusion, there is weak evidence in these studies to support

an inhibitory role for norepinephrine in corticosterone secretion.

ii. Evidence for a Stimulatory Role of Catecholamines

in Stress-Induced Gorticosterone Secretion

Gibson et al. (1986) found that bilateral injections of 6-0HDA into

the PVN prevented the restraint stress-induced increase in plasma

corticosterone levels. Pretreatment with the a 1-adrenoceptor antagonist

prazosin, at a dose which did not influence resting hormone levels,

inhibited the increase in plasma corticosterone following restraint

stress. The authors suggest an excitatory role for norepinephrine via a

receptors in CRF release during acute stress. Similarly, Feldman et al.

(1986) injected 6-0HDA into the PVN and found that PVN norepinephrine

depletion did not affect basal levels or ether stress-induced increases

in corticosterone levels, but it did significantly inhibit the adrenal

response to photic and acoustic stimuli and sciatic nerve stimulation,

suggesting that noradrenergic terminals in the PVN play a role in the

activation of adrenocortical responses following afferent neural stimuli.

Smythe et al. (1983) support the hypothesis that stress-induced ACTH

release from the pituitary is stimulated by hypothalamic noradrenergic

neuronal pathways. They measured norepinephrine activity in the

mediobasal hypothalamus by assessing the ratio of 3,4-dihydroxyphenyl­

ethyleneglycol (DHPG) /norepinephrine. Hypothalamic norepinephrine

54

activity was rapidly increased by cold swim stress in the rat, accompanied

by increased ACTH/corticosterone release into the blood. Norepinephrine

activity and ACTH release were also stimulated in rats following ether

stress. The significant correlation between serum corticosterone and the

ratio of hypothalamic DHPG/norepinephrine (shown by linear regression

analysis of the combined data for control and stressed rats in their

studies) is not evidence, but suggests that hypothalamic norepinephrine

neuronal activity may play an important stimulatory role in stress-induced

ACTH release.

Studies by Johnston et al. (1985) support the data presented by

Smythe et al. (1983). Norepinephrine, dopamine and 5-HT concentrations

as well as the concentrations of their respective metabolites MHPG, DOPAC

and 5-hydroxyindole-3-acetic acid (5-HIAA) were measured by HPLC. The

metabolism of norepinephrine, dopamine and 5-HT in different rat brain

nuclei was estimated by the ratio of the metabolite to its respective

amine. Ether stress produced large increases in plasma ACTH, accompanied

by significant increases in norepinephrine metabolism in the PVN, arcuate

and dorsomedial nuclei. Serotonin metabolism was decreased in the PVN,

supraoptic nucleus and caudal division of the arcuate nucleus, while

significant increases in 5-HT metabolism were detected in the suprachias-

matic nucleus and rostral division of the arcuate nucleus. Dopamine

metabolism was selectively increased in the rostral division of the

arcuate nucleus. The authors interpret the data as suggesting a

stimulatory role for norepinephrine, possibly in the PVN with respect to

ACTH secretion.

Studies by Mermet and Gonon (1988) used differential normal-pulse

55

voltammetry combined with carbon fiber electrodes to measure in vivo

extracellular catechols released by noradrenergic nerve terminals in the

PVN. In rats pretreated with the monoamine oxidase inhibitor pargyline,

ether stress produced an increase in the catechol signal corresponding to

extracellular norepinephrine. Electrical stimulation of the ventral

noradrenergic bundle also caused an immediate increase in the signal.

The increase in norepinephrine release following ether stress was

immediate and lasted as long as the stimulus duration. No dopamine

release was recorded from either the striatum or the Al3 dopaminergic cell

body group caudal to the PVN. These data suggest a facilitatory action

of norepinephrine on PVN cell bodies which are involved in the hormonal

responses to stress.

Immobilization stress (varying from 15 to 180 minutes) significant­

ly increased MHPG-S04 levels and decreased norepinephrine levels in the

following rat brain regions: hypothalamus, amygdala, hippocampus,

thalamus, pons, medulla and cerebral cortex, with the most rapid and

profound increase in MHPG-S04 found in the hypothalamus. However, the

basal ganglia demonstrated the reverse process- increases in norepineph-

rine levels with transient decreases in MHPG-S04 levels. There was a

concomitant increase in plasma corticosterone levels following immobiliza­

tion stress (Tanaka et al., 1982). This study suggests that norepineph­

rine released in the hypothalamus during immobilization stress may play

a role in the stress-induced increase in corticosterone levels.

Beaulieu et al. (1987) demonstrated that bilateral lesions of the

central nucleus of the amygdala significantly inhibited the stress-induced

ACTH secretion following immobilization. These lesions also reduced the

56

stress-induced noradrenergic activity in the anterior and lateral

hypothalamic areas, the PVN and arcuate nucleus, and the bed nucleus of

the stria terminalis. These results suggest that norepinephrine plays a

stimulatory role in ACTH secretion and is somehow regulated by neurons in

the central nucleus of the amygdala.

As stated earlier, the ventral noradrenergic bundle (VNAB) is

thought to be the source for the noradrenergic modulation of ACTH

secretion (see CORTICOSTERONE: Role of the VNAB). Szafarczyk et al.

(1985) have shown that injection of 6-0HDA into the VNAB obliterates the

circadian patterns for ACTH and corticosterone secretion, and also caused

an 80% inhibition of the ACTH response to stress.

Injections of 6-0HDA into the VNAB also prevented the increase in

corticosterone secretion following exposure to photic and acoustic stimuli

and sciatic nerve stimulation (Feldman et al., 1988). However, basal

levels of corticosterone and ether-induced increases in corticosterone

were not affected. The 6-0HDA-injected rats had average norepinephrine

depletions of 72% in the mediobasal hypothalamus and 78% in the paraven-

tricular nucleus when compared to vehicle-injected rats. These data

support previous studies (Feldman et al., 1986) in which 6-0HDA injections

into the PVN did not affect basal levels or ether-induced increases in

corticosterone levels, but did significantly inhibit the adrenal response

to photic and acoustic stimulation and sciatic nerve stimulation. These

studies suggest that norepinephrine plays a role in the activation of

adrenocortical responses following afferent neural stimuli.

In contrast, Velley et al. (1988) have shown that 6-0HDA lesions

in the locus coeruleus did not prevent the effect of novel environment

57

stress on corticosterone, although there was a significant (53%) depletion

of hypothalamic norepinephrine. However, this can be explained by the fact

that about 90% of the norepinephrine supply to the PVN originates from the

lower brainstem Al and A2 cell groups, while the locus coeruleus

projection to the PVN is only about 8% (Sawchenko and Swanson, 1982).

iii. Role of Epinephrine in Stress-Induced

Corticosterone Secretion

Szafarczyk et al. (1987) demonstrated a stimulatory role for

epinephrine (and norepinephrine) on corticosterone release when

intracerebroventricular (icv) injections of epinephrine or norepinephrine

induced ACTH release comparable to that occurring under conditions of

ether stress. Epinephrine was more efficacious than norepinephrine on an

equimolar basis. The stimulating effect of norepinephrine was reversed

by icv pretreatment with prazosin, an a 1- adrenoceptor antagonist, whereas

the effect of epinephrine was inhibited by either prazosin or the .&­

blocker propranolol. These results suggest that norepinephrine triggers

ACTH release via a1-adrenoceptors, whereas epinephrine acts via both a 1-

and B-adrenoceptors. It should be noted that they also administered a

combined pretreatment of prazosin and propranolol, and this combined

treatment was no more effective than either antagonist alone. They also

demonstrated that a unilateral juxta-PVN injection of either epinephrine

or norepinephrine produced an increase in plasma ACTH concentration with

a time course similar to that for an icv injection, and pretreatment with

anti-rCRF-41 serum caused a 50-60% reduction in the ACTH response to

either catecholamine. Since the majority of diencephalic CRF neurons are

58

located in the parvocellular subnuclei of the PVN (Swanson et al., 1983;

Liposits et al., 1985), these data suggest a role for a1

and B receptors

in the PVN in the regulation of corticosterone secretion.

Additional evidence was obtained by Spinedi et al. (1988) who

demonstrated that central inhibition of the enzyme that converts

norepinephrine to epinephrine, phenylethanolamine-N-methyltransferase

(PNMT), produced a selective decrease in hypothalamic epinephrine levels,

significantly decreased the plasma ACTH response to ether stress, and at

later times also caused a selective decrease in CRF content. In addition,

the altered ACTH response to ether stress was not due to changes in

arginine vasopressin levels or changes in pituitary responses to CRF.

These results support a stimulatory role of central epinephrine neurons

on CRF release.

6. Role of Serotonin (5-HT) in Stress-Induced

Corticosterone Secretion

The role of serotonin as a mediator of stress-induced corticosterone

secretion is not clear. Electrolytic lesions in the dorsal raphe nucleus

prevented the stress-induced increase in corticosterone levels as well as

increases in plasma renin activity and plasma renin concentration

following a ten-minute conditioned fear paradigm (Richardson Morton et

al., 1986). Other studies from our laboratory have shown that injection

of the neurotoxin 5,7-dihydroxytryptamine into the dorsal and median raphe

nucleus did not prevent the stress-induced increase in renin and

corticosterone secretion. Furthermore, i. p. injection of the 5-HT2

antagonist LY53857 did not prevent the stress-induced increase in renin

59

and corticosterone secretion (Richardson Morton et al., 1986), whereas low

doses of the 5-HT1 agonist buspirone and ipsapirone inhibited the stress­

induced increase in corticosterone and renin secretion (Van de Kar et al.,

1985; Urban et al., 1986; Lorens et al., 1989). These studies suggest

that a non-serotonergic (or possibly 5-HT1) mechanism is involved in

stress-induced renin and corticosterone secretion.

Feldman et al. (1987) reported that injection of the serotonergic

neurotoxin 5, 7-dihydroxytryptamine (5, 7-DHT) into the PVN blocked the

adrenocortical responses to cortical and extra-hypothalamic limbic

stimulation, but had no effect on the adrenocortical response to ether

stress. Immobilization and cold-exposure stress resulted in four-fold

increases in ACTH-like imrnunoreactivity in plasma. This increase was

potentiated by i.p. injections of the 5-HT reuptake inhibitor fluoxetine

and significantly reduced by the 5-HT antagonists metergoline and

cinanserin (Bruni et al., 1982). These results suggest that 5-HT may play

a role in stress-induced ACTH release. In contrast, Fuller and Snoddy

(1977) found that pretreatment with fluoxetine did not significantly alter

the increase in plamsa corticosterone levels following swim stress or

insulin-induced hypoglycemia. However, fluoxetine did enhance the

corticos terone response produced by an inj ec ti on of L- 5-hydroxytryptophan.

These results suggest that serotonergic neural pathways are involved in

mediating the adrenocortical response to some stressful stimuli (i­

mmobilization and cold exposure) but not others (swim stress and insulin­

induced hypoglycemia).

Beaulieu et al. (1986) support the studies by Bruni et al. (1982).

They found that bilateral lesions of the central nucleus of the amygdala

60

resulted in significantly decreased immobilization stress-induced ACTH

secretion with a concomitant increase in serotonergic activity in the PVN

and arnygdaloid nuclei. Additional studies by Beaulieu et al. (1987)

suggest a serotonergic involvement in immobilization stress. In these

studies, bilateral lesions of the central nucleus of the amygdala in

unstressed rats increased the serotonergic activity in the PVN, dor­

somedial nucleus, anterior hypothalamic area (AHA), cortical and medial

amygdaloid nuclei, and the lateral part of the basal amygdaloid nucleus.

Immobilization stress applied to lesion rats resulted in a decrease in

serotonergic activity to control levels in the PVN, AHA and the dor­

somedial nucleus, and below control levels in the ventromedial nucleus.

The serotonergic activity remained elevated in each of the amygdaloid

areas measured, which are rich in glucocorticoid receptors (Beaulieu et

al., 1986). Glucocorticoids have been reported to down-regulate their own

receptors in a site-specific manner in the amygdala, septum and hippocarn­

pus, but not in the hypothalamus or pituitary (Sapolsky et al., 1984).

Also noradrenergic cell bodies (Al-A7), serotonergic cell bodies (Bl-B9),

and PNMT-immunoreactive nerve cells of the adrenergic cell groups (Cl-C3)

have been shown to have strong glucocorticoid receptor immunoreactivity

(Harfstrand et al., 1986). De Kloet et al. (1986) have shown that

corticosterone selectively controls the 5-HT1 receptor density in the

dorsal raphe and dorsal hippocampus. These studies suggest that serotonin

may modulate the negative feedback of glucocorticoids on stress-induced

ACTH secretion. A moderate to strong nuclear glucocorticoid receptor

immunoreactivity has been demonstrated in the majority of CRF-immunoreac­

tive neurons of the parvocellular PVN, septohypothalamic thalamus, bed

61

nucleus of the stria terminalis, and the medial and central amygdaloid

nuclei (Cintra et al., 1987).

C. Effect of Peripheral Sympathectomy on Stress-Induced Renin and

Corticosterone Secretion

The physiological response to stress involves the release of

catecholamines from the adrenal medulla and sympathetic nerve endings.

These catecholamines stimulate the renal nerves to increase renin output.

It is generally accepted that the sympathetic nervous system mediates the

stress response from the brain to the kidney (Golin et al., 1988; Reid et

al., 1978). Studies by Van de Kar et al. (1985) have shown that the &­

adrenoceptor antagonist propranolol significantly inhibits but does not

completely block the stress-induced increase in plasma renin activity.

It should be noted, however, that this effect could be mediated by either

central or peripheral &-receptors since propranolol is known to cross the

blood brain barrier (Giarcovich et al., 1984; Lemmer et al., 1985). To

determine whether stress-induced renin secretion is mediated by the

peripheral sympathetic nervous system, rats were adrenal enucleated and

given weekly injections of 6-0HDA (i.p.). The adrenal medulla was removed

since it does not effectively take up 6-0HDA. The effectiveness of the

sympathectomy/adrenal enucleation was verified by determination of plasma

epinephrine (less than 1% of control) and norepinephrine (less than 10%

of control). The stressor induced a significant increase in plasma renin

activity and concentration. Sympathectomized-adrenal enucleated rats

showed no inhibition of the stress - induced increase in plasma renin

activity/concentration or plasma corticosterone levels (Richardson Morton

62

et al., 1988). The data suggest that the sympathetic nervous system and

adrenal catecholamines are not the only mediators of the stress-induced

increase in renin secretion.

5. Central Nervous System Pathways Involved in the

Stress Response

a. Sensory Afferents

With respect to the role of the central nervous system in the

conditioned fear response, it is important to consider how the sensory

stimulus (pain) is transmitted to the cerebral cortex for analysis.

Specialized nerve fibers found throughout the skin serve as end organs

for pain. These fibers enter the spinal cord and ascend through the

lateral spinothalamic tract to the posteroventral nucleus of the thalamus.

From here the fibers travel to the appropriate loci in the cerebral

cortex, depending on the somatic origin of the pain message.

The ascending reticular activating system (ARAS) is a network of

interconnective fibers that also travel from the top of the spinal cord

to the thalamus, and eventually to the cortex. This tract receives major

sensory inputs, and conveys information regarding alertness and excite­

ment.

There appears to be a close connection between pain sensation and

the ARAS. The midbrain and brainstem reticular formations receive many

pain afferents from the spinothalamic tract. Studies in animals (Olds

and Olds, 1965) and humans (Gray, 1971) have suggested that stimulation

through electrodes in the periventricular midbrain reticular system

generates pain sensation. Other pain reception fibers have been found in

63

the thalamic portion of the ARAS. Lesions in the thalamic intralaminar

nuclei and the thalamic nucleus centrum medianum have been shown to

alleviate intractable pain in humans (Gray, 1971).

b. The Limbic System

1. Septo-Hippocampal System

The limbic system, in particular the septo-hippocampal system,

appears to play a major role in pain reception. The hippocampus emits a

theta wave (4-9 cps rhythm) as a response to a painful stimulus in the

rat, and in the emotional states of disappointment and frustration in

humans (Gray, 1971). This theta wave can be abolished by lesions in the

medial nuclei of the septum, and the thalamic nucleus centrum medianum,

suggesting that the pain input route most likely travels from the midbrain

reticular formation, through the thalamus, to reach the septum and

hippocampus (Gray, 1971).

2. Amygdala

The amygdala is a limbic structure with many interconnections to the

hippocampus (Krettek and Price, 1978). The amygdaloid complex has been

shown to be involved in the regulation of fear-motivated behavior

(Blanchard and Blanchard, 1972; Spevack et al., 1975). In addition, the

amygdala appears to play a role in the stress response via the pituitary-

adrenal cortical axis. The central nucleus of the amygdala, in par-

ticular, contains large numbers of CRF-containing cell bodies and

terminals (Swanson et al., 1983; Moga and Gray, 1985). Injection of CRF

into the lateral ventricles of rats or dogs elicits autonomic and

64

behavioral changes that are characteristic of the stress or defense

reaction (Brown et al., 1982; Lenz et al., 1987; Britton

et al., 1982). Electrical stimulation of the amygdaloid complexes results

in an increase in adrenal glucocorticoid secretion in rats (Redgate and

Fahringer, 1973), cats (Redgate, 1970) and monkeys (Mason, 1959). The

amygdaloid complex is an essential part of the ACTH controlling mechanism

in rats (Knigge, 1961; Allen and Allen, 1974). Specifically it has been

shown to be necessary for the hypersecretion of ACTH following adrenalec­

tomy (Allen and Allen, 1975). In these studies, lesions which destroyed

the direct medial-projecting amgdalo-hypothalamic fibers blocked the

hypersecretion of ACTH following adrenalectomy, while lesions of the

septal region, stria terminalis and fornices did not.

Lesions in the amygdaloid central nucleus impaired one-way active

avoidance learning (Werka et al., 1978), produced marked deficits in

passive avoidance performance (Mcintyre and Stein, 1973; Grossman et al.,

1975), and significantly decreased ACTH secretion in response to

immobilization stress (Beaulieu et al., 1987). Axons from cells in the

olfactory bulb terminate in the amygdala (Powell et al., 1965) and the

medial hypothalamus receives axons from cells in the amygdala (Cowan et

al., 1965). Bilateral amygdalectomy significantly reduced the adrenocor­

tical responses to sciatic nerve and olfactory stimulation, without

affecting the corticosterone response to ether or to photic and acoustic

stimulation (Feldman and Conforti, 1981). This amygdalo-hypothalamic

pathway is believed to play a role in the adrenocortical response to a

number of somatosensory stimuli (Feldman et al., 1975). The amygdala has

also been shown to play an important role in regulating the cardiovascular

65

system during Sidman avoidance stress testing (Galosy et al., 1982).

3. Hypothalamic Ventromedial Nucleus

With respect to the central neural circuits involved in the

fight/flight behavior, de Molina and Hunsperger (1962) described a pathway

in cats that descends from the amygdala along the stria terminalis to the

ventromedial hypothalamus (VMN) with connections to the midbrain central

grey. Stimulation of sites within the hypothalamus, amygdala and midbrain

central grey could produce defensive attack while other loci immediately

adjacent to those stimulated would produce flight. When the VMN was

removed, stimulation of the amygdala no longer produced defensive attack,

and removal of the central grey prevented the flight behavior following

hypothalamic stimulation.

In addition, Kling et al. (1958) reported that cats made docile by

lesions in the amygdala could be made ferocious by subsequent lesions in

the VMN. The converse did not apply - cats first made ferocious by

lesions in the hypothalamus were not affected by a subsequent amygdalec­

tomy.

Based on these and other observations, Gray (1971) suggested that

the midbrain central grey is the ultimate controlling area for either

flight or defensive behavior. In addition, there is a tonic inhibitory

input to the central grey from the VMN. Further modulation of this

inhibitory input would come from the amygdala and the septo-hippocampal

system. Input from the amygdala would prevent the inhibition, resulting

in fight/flight behavior, and the septo-hippocampal input would intensify

passive avoidance behavior. The resulting behavior (either passive

66

avoidance or fight/flight) would depend on which of the inputs to the VMN

is stronger under a particular set of environmental conditions. This

scheme would provide a basis for the observation that, in times of danger,

both freezing and fight/flight patterns may be stimulated simultaneously

and alternate rapidly with each other.

Neurons in the VMN have been shown to play a role in stress-induced

renin secretion. Gotoh et al. (1988) found that bilateral electrolytic

lesions in the VMN prevented the increase in plasma renin activity

following immobilization stress. Lesions in the VMN led to excessive

levels of food intake and adiposity (Hetherington and Ranson, 1942), and

elicited a variety of metabolic and behavioral disturbances termed the VMN

syndrome, which has become a model for the study of obesity and motivated

behavior (Weingarten et al., 1985; Stellar, 1954).

c. Role of Brain Catecholamines in Stress

A variety of stressors including foot shock (Bliss et al., 1968;

Thierry et al., 1968, Iimori et al., 1981), immobilization (Corrodi et

al., 1968; Tanaka et al., 1981) and ether stress (Vellucci, 1977) have

been shown to decrease norepinephrine concentration and increase

norepinephrine turnover in the rat brain.

The metabolism of norepinephrine in the rat brain results in the

production of the sulfate conjugates MHPG-S04 , the major metabolite, and

DHPG-S04 (Schanberg et al., 1968; Sugden and Eccleston 1971). The

functional activity of central noradrenergic neurons has been shown to be

directly related to the production of MHPG-S04 (Stone, 1975). Levels of

MHPG-S04 in rat brain represent a useful index for norepinephrine turnover

67

in discrete brain regions (Kohno, 1981) as well as in the whole brain

(Korf et al., 1973; Stone, 1976).

Rats subjected to a mild psychological stressor (exposure to

responses of shocked rats) had decreased norepinephrine and increased

MHPG-S04 concentrations preferentially in the hypothalamus and amygdala.

Rats subjected to footshock stress had decreased norepinephrine and

increased MHPG-S04 not only in the hypothalamus and amygdala but also the

pons, medulla oblongata and cerebral cortex (Iimori et al., 1981). This

study suggests that the noradrenergic response to stress varies with the

nature or intensity of the stressor.

Brain noradrenergic systems, in particular the locus coeruleus­

noradrenergic system, have been shown to play a role in nociception,

anxiety and fear (Charney and Redmond, 1983). The locus coeruleus

receives innervation from sensory (pain) pathways in the spinal cord, and

relays information to the cerebral cortex and limbic structures (hypothal­

amus, hippocampus, amygdala) (Kuhar, 1986).

Rasmussen and Jacobs (1986) have recorded the electrical activity

of noradrenergic neurons in the locus coeruleus of freely moving cats.

During GER training, locus coeruleus units showed a large increase in

activity in response to a stimulus paired with a noxious air puff, whereas

no increase in unit activity is seen in response to a stimulus not paired

with the air puff. Also, electrical stimulation of the locus coeruleus

in monkeys elicited behavioral responses identical to those seen in

experimentally-induced or naturally-occurring fear states (Bertrand,

1969).

In humans, drugs that increase noradrenergic function have been

68

shown to produce anxiety. The a 2-adrenoceptor antagonists yohimbine and

piperoxane induced anxiety states in normal subjects (Soffer, 1954;

Holmberg and Gershon, 1961). These studies support the role of brain

norepinephrine in mediating the response to anxiety and fear.

The Kety hypothesis (Kety, 1970) proposed that forebrain nor­

epinephrine enhances cell firing in neurons that receive afferent

environmental information to "affectively important events". Further,

norepinephrine would produce a "persistent facilitation" of these inputs

which would reinforce the memory process for these events. The extensive

innervation of the limbic system (especially the hippocampus) by the locus

coeruleus-norepinephrine system suggests that noradrenergic neurons may

play a role in the memory or cognitive process of stress.

D. HYPOTHALAMIC PARAVENTRICULAR NUCLEUS (PVN)

A. Introduction

The PVN is a bilateral, triangular-shaped nucleus at the dorsal

end of the third ventricle in the hypothalamus. It is a morphologically

complex structure, with eight clearly distinguishable subdivisions, which

contain as many as thirty different putative neurotransmitters. Three

major efferent projection sites, arising from separate neuron groups,

include (1) the median eminence, (2) the posterior pituitary and (3) the

autonomic centers in the brain stem and spinal cord (Swanson and

Sawchenko, 1983; Saper et al., 1976). This suggests that the PVN serves

as an integrator of endocrine, autonomic and behavioral functions in a

variety of physiological systems.

69

B. Morphology

The PVN is composed of distinct parvocellular and magnocellular

divisions (Krieg, 1932). Swanson and Kuypers (1980) have further

characterized the nucleus into eight distinct subdivisions, three

magnocellular and five parvocellular. The anterior magnocellular part of

the PVN is ventromedial to the descending column of the fornix as it

enters the hypothalamus from the septum. The medial magnocellular

division is just caudal to the anterior magnocellular level. The largest

magnocellular division is the posterior magnocellular part. It first

appears in the ventromedial PVN, just caudal to the medial magnocellular

division. After dorsolateral expansion it forms a compact cell mass at

the lateral margin of the nucleus. Hatton et al., (1976) further

subdivide this group by defining two distinct populations of large cells

in the medial and lateral halves of the posterior magnocellular region.

The lateral group contains cells with multiple nucleoli that are larger

and more uniform in orientation, when compared to the medial group. The

posterior magnocellular cells number from 1300-2000 on either side of the

brain (Olivecrona, 1957; Bodian and Maren, 1951; Swanson and Sawchenko,

1983).

There are five distinct parts of the parvocellular division. The

periventricular part, medial to the rest of the PVN, contains small,

vertically-oriented fusiform cells. The anterior parvocellular group is

not well-defined but consists of small- to medium-sized cells at the level

of the anterior and medial magnocellular parts. Extending through the

caudal half of the nucleus is the medial parvocellular part, which

70

consists of densely packed cells. The dorsal parvocellular part is also

in the caudal half of the nucleus and consists of horizontally-oriented,

medium-sized cells in an elliptical group. The lateral parvocellular part

of the PVN is in the caudal third of the nucleus and is the lateral

expansion of the parvocellular part (the wings). These cells are

horizontally-oriented, medium-sized and fusiform shaped. At its maximum

extent the lateral parvocellular PVN extends laterally over the descending

column of the fornix, about 750-1000 um from the midline (Swanson and

Sawchenko, 1983).

Cells in the parvocellular division project primarily to the

external lamina of the median eminence (Wiegand and Price, 1980; Swanson

et al., 1980) and to the brainstem (dorsal vagal complex) and spinal cord

(Hosoya and Matsushita, 1979; Swanson and Kuypers, 1980). The PVN is also

known to project to those segments of the thoracic cord thought to contain

sympathetic preganglionic neurons which innervate the kidney (Swanson and

Mc Kellar, 1979). Cells in the magnocellular division project primarily

to the posterior level of the pituitary gland (Sherlock et al., 1975).

The parvocellular division of the PVN contains about 7000 neurons

in the following distribution: anterior 1650; dorsal 600; lateral 1300;

medial 2175; periventricular 1275 (Sawchenko and Swanson, 1981).

C. Role of the PVN in Behavior

The PVN plays a major role in the control of feeding. Bilateral

damage to the PVN results in increased levels of food intake and excesive

weight gain (Leibowitz et al., 1981). This eating behavior can be

elicited by hypothalamic injections of norepinephrine and other a-

71

noradrenergic agonists (Leibowitz, 1975). In particular, injection of

the a2-agonist clonidine into the PVN initiates a feeding response in

satiated rats (Marino et al., 1983). Goldman et al. (1985) observed a

potent feeding response in satiated animals following injection of

norepinephrine or clonidine into the PVN. This effect was blocked by the

a-antagonist phentolamine as well as the a2-antagonists rauwolscine and

yohimbine. These results suggest that a2 receptors in the PVN play a role

in mediating feeding behavior. Leibowitz et al. (1984) have demonstrated

that these a2 noradrenergic receptors fluctuate in density with respect to

the circadian cycle, exhibiting a unimodal peak at the onset of the dark

cycle, when feeding normally occurs. This shift in a 2 receptor density may

be correlated with plasma corticosterone levels, which also peak at the

onset of the dark cycle (Krieger and Hauser, 1978; Wilkinson et al.,

1979). Norepinephrine-stimulated feeding behavior was prevented by

adrenalectomy and restored by administration of corticosterone (Leibowitz

et al., 1984).

D. Afferent Connections to the PVN

Descending afferent connections to the PVN include a large input

from the limbic system structures, especially the lateral septa! nucleus

and the ventral portion of the subicular cortex, as well as input from

the amygdala, in particular, the medial nucleus (Silverman et al., 1981;

Tribollet and Dreifuss, 1981). The PVN receives afferent sensory

information from the telencephalic limbic system. The bed nucleus of the

stria terminalis receives a large input from the amygdala (Krettek and

Price, 1978) and smaller projection from the ventral part of the subiculum

72

(Swanson and Cowan, 1977). The bed nucleus has been shown to project to

the PVN by autoradiographic studies (Swanson and Cowan, 1979). This

projection innervates all parts of the parvocellular division and the

oxytocinergic parts of the magnocellular division (Sawchenko and Swanson,

1981). Electrophysiological studies have shown that stimulation of the

amygdala or septum can have inhibitory (and sometimes excitatory) effects

on the PVN magnocellular neurons (Renaud and Arnauld, 1979; Koizumi and

Yamashita, 1972; Pittman et al., 1981).

Anterograde autoradiographic techniques were used to describe brain

stem afferents to various PVN regions (McKellar and Loewy, 1981). The

parabrachial nuclei and locus coeruleus project to the posterior,

periventricular, parvocellular and dorsal divisions. The ventral medulla

projects to the lateral, medial, posterior, dorsal and parvocellular

divisions. The Al catecholamine cell group projects to the posterior,

lateral, medial and dorsal divisions, while the NTS projects to the

parvocellular and dorsal divisions. The projections from the Al and A2

(NTS) catecholaminergic cell groups to the PVN have been further

characterized by an immunofluorescence double-labeling procedure to

determine which anterogradely labeled fibers and terminals in the PVN also

displayed immunoreactive dopamine ~-hydroxylase, indicating the presence

of catecholaminergic neurons (Cunningham and Sawchenko, 1988). (See Role

of Brain Catecholamines in Corticosterone Secretion)

Studies using retrogradely-transported horseradish peroxidase

identified cell bodies of neurons that project to the PVN. Labelled cell

bodies were found in the mediobasal hypothalamus, the limbic system

(lateral septum, posterornedial arnygdala and ventral subiculum) and the

73

brain stem, including the dorsal raphe nucleus, locus coeruleus,

parabrachial nuclei, NTS and lateral reticular nucleus (Tribollet and

Dreifuss, 1981).

E. Efferent Projections of the PVN

Descending autonomic projections of the PVN were investigated using

the anterograde transport technique of Phaseolus vulgaris leuco-agglutinin

(Luiten et al., 1985). Two major fiber bundles were described. One

(fiber bundle 1) leaves the PVN dorsally and runs caudally in a dorsal

periventricular position to reach the midbrain periaqueductal grey, where

a number of terminal boutons are observed. This bundle then courses

laterally to join the lateral lemniscus. Some projections innervate the

dorsal raphe nucleus, the parabrachial nucleus, locus coeruleus and

pontine raphe. The second group of fibers (fiber bundle 2) runs medially

and caudally from the PVN and travels in a position dorsal to the

transition of the internal capsule and medial forebrain bundle. In the

midbrain, fiber bundle 2 moves laterally to cover the dorso-lateral pole

of the substantia nigra, and has some terminations in the ventral aspects

of the reticular formation. At the level of the pons, fiber bundle 2

moves ventrally to join the lateral lemniscus (and bundle 1) to reach the

floor of the brainstem. Posterior to the pons, all descending PVN

efferents are now in a ventral position adjacent to the pyramidal tracts.

While in the medulla, some fibers branch off to terminate in the reticular

and medullary raphe nuclei. Nearly all of the bundle fibers traverse the

lateral reticular nucleus, nucleus ambiguus, parvocellular reticular

nucleus and area postrema, with an especially dense amount of varicosities

74

and terminal boutons in the NTS and dorsal motor nucleus of the vagus.

F. Catecholaminergic Innervation of the Hypothalamic PVN

There is substantive evidence supporting extensive catecholaminergic

innervation of the PVN (Carlsson et al., 1962; Lindvall and Bjorklund,

1974; Hokfelt et al., 1974). Most of the adrenergic and noradrenergic

innervation to the PVN is provided by (1) the Al and Cl cell groups of the

ventral medulla, (2) the A2 and C2 cell groups in the medial NTS and (3)

the A6 cell group, the locus coeruleus of the pontine central gray. The

locus coeruleus receives major fiber projections from the Al region, and

projects to the most medial parts of the parvocellular PVN (Swanson et

al., 1981).

G. Neuroactive Substances in the PVN

The colocalization of CRF and vasopressin in the parvocellular PVN

has already been presented (see CORTICOSTERONE: CRF). These are only two

of many neuroactive substances found in the PVN by immunohistochemical

studies. Swanson and Sawchenko (1983) have compiled a thorough listing:

Cell Bodies include oxytocin and vasopressin in the magnocellular PVN

(Swaab et al., 1975; Vandesande and Dierickx, 1975); somatostatin in the

periventricular part of the parvocellular PVN (Dierickx and Vandesande,

1979); dopamine in the periventricular and adjoining parvocellular PVN

(Hokfelt et al., 1973); met- and leu-enkephalin in the magnocellular and

pavocellular PVN (Rossier et al., 1979; Finley et al., 1981); neurotensin

in the medial and periventricular parts of the parvocellular PVN (Kahn et

al., 1980); dynorphin coexists with vasopressin in the magnocellular PVN

(Watson et al., 1982); substance Pin the parvocellular PVN (Ljungdahl et

75

al., 1978); glucagon in the magnocellular PVN (Tager et al., 1980); renin

in the oxytocinergic parts of the magnocellular PVN (Fuxe et al., 1980);

and ACTH and i-endorphin in the magnocellular PVN (Joseph and Sternberger,

1979; 'Watkins, 1980). The PVN is also the major source of atrial

natriuretic peptide (ANF)-containing cells that project to the median

eminence (Palkovits et al., 1987). Double-labelling studies with antisera

to GAD and CRF in the PVN have indicated that a sub-population of CRF­

containing neurons contain the inhibitory neurotransmitter GABA (Meister

et al., 1987).

Nerve terminals and fibers in addition to norepinephrine,

epinephrine and serotonin (previously described) include "cholinoceptive"

acetylcholine neurons (Kimura et al., 1981); ACTH, i-endorphin and a-MSH

in the parvocellular and oxytocinergic divisions of the magnocellular PVN,

all of which originate in the arcuate nucleus (Joseph, 1980; Bloom et al.,

1978; Jacobowitz and O'Donohue, 1978); VIP in the periventricular part of

the parvocellular PVN (Sims et al., 1980); GABA (Perez de la Mora et al.,

1981); TRH in primarily parvocellular PVN (Hokfelt et al., 1975); and

sparse fibers of LHRH (Jennes and Stumpf, 1980); bradykinin (Correa et

al., 1979); prolactin (Fuxe et al., 1977) and angiotensin II (Fuxe et al.,

1976; Changaris et al., 1978).

a. Neuropeptide Y

The distribution of neuropeptide Y (NPY) immunoreactivity in the

PVN as well as its coexpression in noradrenergic and adrenergic brain stem

groups has been determined with the retrograde transport-double immunohis-

tochemical techniques (Sawchenko et al., 1985). Antisera to phenyl-

76

ethanolamine-N-methyltransferase (PNMT) was used as a marker for

adrenergic neurons, and dopamine-&-hydroxylase (DBH) antisera was used as

a marker for adrenergic and noradrenergic neurons. The results indicated

a dense NPY input to most parts of the parvocellular PVN in a similar, yet

more extensive distribution than that stained by DBH antisera. In the

magnocellular division, NPY-stained fibers were evenly distributed

throughout areas containing oxytocinergic and vasopressinergic fibers.

Catecholaminergic fibers were preferentially associated with vasopres­

sinergic fibers. Combined retrograde transport-double irnrnunohistochemical

labeling studies demonstrated that NPY irnrnunoreactivity is extensively

contained within the adrenergic Cl, C2 and C3 cell groups, with somewhat

less irnrnunoreactivity in the noradrenergic cell groups, mostly limited to

neurons in the Al cell group.

Leibowitz et al. (1986) found that NPY injected directly into the

PVN increased plasma corticosterone levels within 15 minutes of infusion.

The NPY response was similar to that for epinephrine and norepinephrine

(See CORTICOSTERONE: Stimulation by Norepinephrine and Epinephrine).

Also, NPY injected into the third ventricle results in robust feeding

behavior in satiated rats (Allen et al., 1975) in a similar manner to

adrenergic neurotransmitters (See PVN: Role in Behavior). Studies by Sahu

et al. (1988) suggest that NPY neurons in the brainstem project to the PVN

and mediate the increased sensitivity of the feeding response elicited by

NPY.

CHAPTER III

METHODS

A. Animals

Male Sprague-Dawley rats (200-250 g) were purchased from Sasco-

King Animal Laboratories (Oregon, WI). The rats were housed, two per

cage, in a temperature (20-22°C), humidity (50-55%) and illumination (12:12

hour light/dark cycle; lights on at 0700 h) controlled room. Water and

food (Wayne Lab Blox, Lab Mills Inc. , Chicago, IL) were available ad

libitum.

B. Surgery

All surgery was performed under pentobarbital anesthesia (50

mg/kg, i.p.). The rats received atropine sulfate (0.4 mg/kg, i.m.) to

reduce respiratory tract secretion and ampicillin ( 50 mg/kg, i. m.) to

prevent infection. Chloramphenicol (1% ophthalmic ointment) was topically

applied around the wound margin. The rats were euthanized by decapitation

with a guillotine in an area outside the rat room. Trunk blood was

collected into centrifuge tubes containing 0.5 ml 0.3 M ethylenediamine

tetra acetic acid (EDTA; pH 7.4).

77

78

c. Electrolytic Destruction of the PVN

Bilateral electrolytic lesions in the PVN were produced by passing

1. 0 mA (D. C.) current for 10 seconds through an intracranial cathode. The

anode was clipped to the wound margin. The cathode was a stainless steel

pin (0.25 mm diameter) insulated with Epoxylite except for 0.5 mm of its

tip. A Kopf stereotaxic apparatus and a Grass DC constant current lesion

maker were used. The coordinates for the lesion site were 6.0 mm rostral

to lambda, 0.5 mm lateral to the midsagittal suture, and 8.3 mm ventral

to the skull surface, with the incisor bar set at -4. 0 mm. Control

animals were treated identically except that the electrode was not lowered

into the skull. The rats were allowed to recuperate from surgery for 14

days, after which time they were subjected to a conditioned emotional

(fear) response paradigm (GER). The CER procedure is described in detail

below. The rats were sacrificed by decapitation immediately following the

last CER procedure. Trunk blood was collected and assayed as described

below. The brains were removed and saved for histological verification

of the lesion placement. Eight rats were included in the sham/control as

well as the sham/stress group, while 12 rats per group were used in both

the lesion/control and lesion/stress group.

D. Electrolytic Destruction of the Nucleus Reuniens

Bilateral electrolytic lesions of the nucleus reuniens were

produced by passing 1.0 mA current for 5 seconds through an intracranial

cathode while the anode was clipped to the wound margin. The coordinates

for the lesion site were 5.35 mm rostral to lambda, 0.4 mm lateral to the

midsagittal suture and 8. 0 mm ventral to the skull surface, with the

79

incisor bar set at -4.0 mm. Eight rats were included in the sham/control

as well as the sham/stress group, 10 rats were used in the lesion/control

and 12 rats were used in the lesion/stress group. The rest of the

protocol for this experiment was identical to that described in the

previous section (Electrolytic Destruction of the PVN).

E. Microiniections of Ibotenic Acid into the PVN

Bilateral stereotaxic microinjections of ibotenic acid (Sigma,

St. Louis, MO) were made using a Kopf stereotaxic apparatus with a 5 µl

Hamilton syringe and a Trent Wells hydraulic microdrive, to allow for a

slow delivery rate. Ibotenic acid (10 µg/µl) was dissolved in 0.01 M

phosphosaline buffer (PBS, pH 7.4), and injected in a volume of 0.3 µl

per side over five minutes (Fitzsimons and Ciriello, 1986). The injection

coordinates (incisor bar at -4.0 mm) were 5.7 mm anterior to lambda, 0.5

mm lateral to the midsagittal suture and 8. 5 mm ventral to the skull

surface. Control animals received an injection of the vehicle. After

completion of the injection (five minutes), the needle was left in situ

for an additional five minutes to prevent dorsal diffusion of the

neurotoxin along the needle track. The rats were allowed to recuperate

from surgery for 14 days, after which time they were subjected to the CER

procedure. The rats were sacrificed by decapitation immediately following

the CER procedure, and trunk blood was collected and later assayed. Nine

rats were included in the vehicle/control and eight rats in the vehicle/­

stress group, while 12 rats per group were used in both the ibotenic

acid/control and ibotenic acid/stress group.

80

F. Microinjections of 6-Hydroxydopamine into the PVN

Bilateral stereotaxic microinjections of 6-0HDA-HBr (Regis, Morton

Grove, IL) were made using a Kopf stereotaxic apparatus with a bilateral

cannula constructed of two 26 gauge stainless steel tubes (injectors) with

a 1. 0 mm distance between centers. The cannula was connected with

flexible tubing to two 25 ul Hamilton syringes and a Sage Instruments

syringe pump, model 355. 6-0HDA was dissolved in 0.1% ascorbic acid in

physiological saline, to prevent autooxidation. In addition, the vials

of drug solution were kept on ice and covered with aluminum foil to

prevent light exposure. 6-0HDA was injected bilaterally over a five

minute period, in a concentration of 8.0 µg free base in 1.5 µl saline

containing 0.1% ascorbic acid. Control animals received an injection of

the vehicle (0.1% ascorbic acid in physiological saline). After completion

of the injection, the needle was left in situ for an additional five

minutes to inhibit dorsal diffusion of the neurotoxin along the needle

track. The rats were injected with the monoamine oxidase inhibitor

pargyline ( 30 mg/kg, i. p. ; Sigma, St. Louis, MO) 30 minutes prior to

surgery, to augment the neurotoxic effect of 6-0HDA. The coordinates for

the injection site were 5.7 mm rostral to lambda, 0.5 mm lateral to the

midsagittal suture and 8. 7 mm ventral to the skull surface, with the

incisor bar set at -4.0 mm. The rats were allowed to recuperate from

surgery for 14 days, after which time they were subjected to the CER

paradigm, described below. The rats were sacrificed by decapitation

immediately following the last CER procedure. Trunk blood was collected

and assayed as described below. The extent of damage to noradrenergic

nerve terminals was determined by immunocytochemical techniques (see

81

below). Eight rats were included in the vehicle/control as well as the

vehicle/stress group, while 12 rats per group were used in both the 6-

0HDA/control and 6-0HDA/stress group.

G. Microinjections of Sotalol into the PVN

Rats were placed in a Kopf stereotaxic apparatus and implanted with

chronic bilateral guide cannulae (Plastic Products, Roanoke, VA). The

guide cannulae were constructed of two 26 gauge stainless steel tubes with

a 1.0 mm distance between centers. The inner stylets were constructed of

33 gauge stainless steel wire. The rats were allowed to recuperate from

surgery for 14 days, and were handled each day following the surgery to

condition them for handling during the injection. At the end of the 14

day period, the rats were subjected to the CER procedure, described below.

In this particular study, the rats were transported in their home cage to

the CER room, wrapped in a small towel, and held while the dust cap and

stylets were removed and replaced with a 33 gauge bilateral injector. The

rats received an injection of either vehicle or sotalol, or no injection

at all. All injections were performed approximately three minutes before

placement in the stress chamber, and 13 minutes before sacrifice. The

volume (0.5 µl per side) was injected over a 60 second interval. The

injector remained in situ for an additional 60 seconds to help prevent

dorsal diffusion of the drug. Sotalol was injected bilaterally (0.5 µl

per side), using two 25 ul Hamilton syringes and a CMA microinjection

pump. The concentration of sotalol was 2.0 µg in 0.5 µl of physiological

saline. Vehicle rats received an injection of saline. The non-injected

rats had their dust caps and stylets removed and injectors placed, but no

injection occured. The rats were sacrificed by decapitation immediately

82

following the CER procedure. Trunk blood was collected and assayed as

described below. Fourteen rats were included in the vehicle/control and

vehicle/stress groups. The non-injected/control and non-injected/stress

groups had six rats each. Sixteen rats were included in the sotalol/-

control and eighteen rats were in the sotalol/stress group.

H. Histology

Lesion placement and drug injection sites were verified histologi­

cally. The brains were preserved in buffered formalin for 48 hours. The

brains were then infiltrated with sucrose by successive four hour

intervals in 5% and 10% sucrose solutions, respectively. The brains were

stored in a 20% sucrose solution until they were cut on a cryostat (40

micron sections) and stained by the cresyl violet method and inspected by

a microscope to verify the position of the tip of the needle or extent of

lesion damage.

I. Immunocytochemistry

The brains from the 6-0HDA-treated rats were fixed for five hours

in a solution containing 4% paraformaldehyde-0.05% glutaraldehyde in 0.10

M phosphate buffer, pH 7.5 at 4°C. They were then transferred to a 15%

sucrose PBS solution where they were stored overnight at 4°C (Furness et

al., 1978). The brains were then mounted in a cryostat, frozen and cut

into 40 micron sections. These sections were floated in 0.05 M PBS and

mounted on slides and allowed to air dry. The slides were then placed in

air-tight slide boxes and kept at 0°C until processed (approximately 48

hrs). Sections were then washed and incubated in immunocytochemistry

83

(ICC) buffer consisting of phosphosaline buffer (PBS), 1.0% normal goat

serum (from second antibody host species), 0.1% gelatin, 0.01% thimerosal,

O - 0.002% neomycin, and 0 - 0.2% triton X-100. Sections were incubated

consecutively in the primary antibody, rabbit anti-bovine adrenal

medullary dopamine ~-hydroxylase (DBH, 1:1000 dilution) for 12 hours in

humidity-controlled chambers at 4°C. The DBH antibody was obtained from

two sources: (1) Eugene-Tech, Allendale, NJ and (2) Dr. Eric Schweitzer,

Madison, WI. The two DBH antibodies gave nearly identical results when

tested on control rats. This incubation was followed by addition of a

1:50 dilution of goat anti-rabbit IgG followed by an incubation with 1:200

peroxidase-antiperoxidase (rabbit PAP) solution for 30 minutes at room

temperature. Sections were washed between each immune incubation with

0.02 M PBS. Bound peroxidase was visualized by incubating sections in 25

mg/100 ml diaminobenzidine tetrahydrochloride (DAB-4HC1) and 0.01%

hydrogen peroxide for 5-7 minutes. To increase sensitivity this solution

also included 2.5% nickel ammonium sulfate. Sections were washed, mounted

with cover slips, and viewed on a Leitz Ortholux photomicroscope.

J. Stress (GER) Procedure

In each of the experiments, the rats had at least a two-week post­

operative recovery period before being subjected to the conditioned

emotional response (GER) paradigm. The GER was performed in a ten-inch

square, clear Plexiglass chamber with a grid floor composed of stainless

steel rods (0.3 inches in diameter) spaced 0.5 inches apart. The chamber

was illuminated by a 7.5 W incandescent lamp. The chamber was located

inside a larger, rectangular sound-attenuating box equipped with a two-

84

way mirror (13" x 9") so that the rat's behavior could be observed. The

sound-proof box also had a small fan in the back wall for ventilation.

Scrambled current (0.8 mA D.C.) shock for 10 seconds was delivered through

the grid floor by a Grayson-Stadler shock generator.

The rats were assigned at random before testing to a particular

treatment group (control or stress) such that cage mates were members of

the same group. Cage mates were tested sequentially and at the same time

each day (between 10:00 and 15:00 hours) on four consecutive days.

On each day, the rats were transported ten feet from their living

quarters to the stress chamber. Ten minutes following placement in the

chamber, the experimental animals received an inescapable foot shock, then

were returned to their home cage. This procedure was repeated for three

consecutive days. The behavior of the rats was observed on days 3 and 4

of the CER paradigm. It was apparent from these observations that the

stressed rats had learned that their placement in the chamber would be

followed by a foot shock. By the third day, in contrast to the control

animals, the stressed rats defecated, urinated and demonstrated freezing

(no movement) behavior. On the fourth day, after 10 minutes in the test

chamber, instead of receiving shock the rats were removed, transported to

a third area outside the stress room and immediately sacrificed by

decapitation. The control rats were treated identically except that shock

was not administered at any time.

The stress chambers were cleaned carefully after each rat had been

tested. Fecal boli and urine were removed, as well as washing the grid

floor and sides of the plexiglass chamber. This was to prevent the

analgesic effect described by Fanselow (1985) that can occur when an

85

unstressed rat is placed in a chamber containing the odor of a rat that

has received an electric shock. Rats have been shown to release an

innately recognized odor in response to many types of aversive events

(Brown, 1979). In particular, aversive electric shock elicits an odor

that conspecifics can discriminate from the odors of unstressed rats

(Valenta and Rigby, 1968). These odors have been shown to interfere with

the acquisition of one-way shock avoidance (Dua and Dobson, 1974).

K. Biochemical Determinations

1. Plasma Renin Activity (PRA)

PRA was measured by radioimmunoassay for generated angiotensin I

(Haber et al., 1969; Stockigt et al., 1971). The pH of the samples was

reduced to pH 6.5 by addition of 0.5 ml of 0.5 M phosphate buffer pH 6.0.

The conversion of angiotensin I to angiotensin II was prevented by

addition of 0.025 ml of 8-hydroxyquinoline (8-HQ; 10% suspension in 0.3

M EDTA; 3.4 mM final concentration; Mallinckrodt, St. Louis, MO) and 0.020

ml of phenylmethylsulfonyl fluoride (PMSF; 5 g in 150 ml ethanol; 2.5 mM

final concentration; Sigma, St. Louis, MO). Plasma samples were incubated

for three hours at 37°C to generate angiotensin I (AI). The incubation was

stopped by immersion of the samples in boiling water for three minutes.

The radioimmunoassay of AI was performed with an antiserum against AI at

a dilution of 1:30,000, and a total binding of 35%. The antibody was

added to the samples in a volume of 0.3 ml (AI antiserum in a 0.1 M Tris

buffer, pH 8.0, containing 0.1% gelatin) and incubated overnight at 4°c.

The tracer ( 125I labelled AI) was added to the samples the following

morning, in a volume of 0.4 ml per tube, approximately 10,000 counts per

86

minute per tube. On the next day, the antibody-bound (iodinated AI)

tracer was separated from the free tracer by addition of O. 5 ml of a

suspension containing activated charcoal and dextran T500 (Pharmacia,

Piscataway, NJ) and centrifugation at 2000 x g for 15 minutes. The pellet

containing the unbound tracer was counted on a micromedic 4/200 plus

automatic gamma counter which was connected via a buffer (Quadram

microfazer) to an IBM PC-AT computer. The data were analyzed by a program

(RIA-AID) that was purchased from Robert Maciel Associates (Arlington,

MA). The sensitivity limit of the RIA is 10 pg AI per tube and the intra­

assay variability was 4.0%. The inter-assay variability was 7.4% (Van de

Kar et al., 1984). The tracer (AI; Beckman, Arlington Heights, IL) was

iodinated via the chloramine T (Sigma, St. Louis, MO) method (Greenwood

et al., 1963). Phosphate buffer (0.05 M; 0.02 ml), 12 µl AI (0.5 µg/µl)

and 20 µl chloramine T (3.5 mg/ml phosphate buffer) were added to the vial

containing the 125I and mixed for three seconds. This was immediately

followed by addition of 25 µl of sodium metabisulfite (4. 5 mg/ml phosphate

buffer; J. T. Baker Chemical Co. , Phillipsburg, NJ) to terminate the

reaction. Chromatography of the iodinated AI was performed first by ion

exchange (Biorad AG-1X4, 200-400 mesh) chromatography. The column

consisted of a siliconized Pasteur pipette (5 3/4"; Scientific Products)

with a small amount of glass wool placed in the tip of the pipette to

support the anion exchange res in. Fractions (12 drops per tube) were

collected from the column and the radioactivity in each tube was measured.

The radioactive peak fractions eluted off this column were then placed on

a Sephadex G-15 column for gel filtration and eluted with 0.05 N acetic

acid containing 0.1% bovine serum albumin (Sigma, St. Louis, MO). Two ml

87

fractions were collected off the Sephadex column and measured for

radioactivity. The tubes containing the radioactive AI were pooled and

stored at -10°c.

2. Plasma Renin Concentration (PRC)

Renin is the enzyme responsible for the conversion of the

substrate angiotensinogen, produced by the liver, to

Angiotensin I. The measurement of plasma renin activity reflects the

concentration of renin as well as renin substrate (angiotensinogen) in

plasma. The normal angiotensinogen concentration is only about or less

than Km (Reid et al., 1978) and thus much less than the concentration

required to generate AI at maximum velocity (Vmax). Studies by Van de

Kar et al. (1981) have demonstrated that 5-HT dependent changes in plasma

renin activity are not accompanied by changes in plasma angiotensinogen

levels. However, since electrolytic lesions in the PVN could cause a

reduction in corticosterone levels, this could lead to decreased secretion

of angiotensinogen from the liver. This would cause plasma renin activity

values to be low even when renin levels were unaffected. The assay for

plasma renin concentration involves the addition of a saturating

concentration of exogenous renin substrate to the samples such that renin

is working at Vmax. The measurement of plasma renin concentration

eliminates the possibility that decreased substrate levels would be

misinterpreted as decreased levels of renin.

The renin substrate was plasma from nephrectomized rats which

received dexamethasone (0.2 mg/rat) 24 hours before sacrifice. The plasma

samples (0.2 ml) received 0.1 ml of nephrectomized plasma, 0.1 ml of 0.5

88

M phosphate buffer pH 6.0, and 5 µl each of PMSF and 8-hydroxyquinoline

(to a final concentration of 2.5 mM and 3.4 mM, respectively). The

mixture was incubated at 37°C for one hour. The reaction was stopped by

immersion of the tubes in a boiling water bath for three minutes. The

radioimmunoassay of AI is as described for PRA.

3. Plasma Angiotensinogen

This radioimmunoassay is similar to that for PRC except that a

saturating amount of the enzyme renin (from kidney homogenates) is added

to the plasma samples before incubation to allow conversion of endogenous

renin substrate to AI (Menard and Catt, 1972). Therefore any differences

in the production of AI will be attributed to differences in substrate,

not enzyme levels.

The plasma samples (0.1 ml) received 0.1 ml of 0.5 M phosphate

buffer pH 6.0, 0.1 ml renin (from rat kidneys homogenized in cold water

1 g/ml at a 1:1,000 dilution), and 5 µl each of PMSF and 8-HQ. The

mixture was incubated at 37°C for one hour. The reaction was stopped by

immersion of the tubes in a boiling water bath for three minutes. The

remainder of the radioimmunoassay for AI is as described for PRA.

4. Plasma Corticosterone

Corticosterone radioimmunoassays on unextracted plasma (5 and 10

µl) were performed using procedures and antiserum from Radioassay Systems

Laboratories (Carson, CA). The binding proteins were denatured by

incubation of the sample tubes and assay buffer in a hot water (80°C) bath

for twenty minutes. The standards were made of serial dilutions of

89

corticosterone (Steraloids, Wilton, NH) in assay buffer. The antiserum

(catalog # 1472, lot # 3R3-38) was used at a final dilution of 1:10,500

and total binding of 40%. The incubation of the samples (or standard)

with the antiserum and tracer (corticosterone l,2,6,7- 3H(N)-; NET-399,

New England Nuclear) occurred overnight at 4°C. The antibody-bound tracer

was separated from the unbound tracer by addition of a charcoal-Dextran

T70 suspension, centrifugation, and decanting of the supernatant into

scintillation vials containing 5 ml of a scintillation cocktail (4 g of

Omnifluor in 1 liter of toluene). The samples containing the supernatant

and scintillation cocktail were counted in a Packard liquid scintillation

counter. The data were stored on a diskette by a Diskstore (United

Technologies, Packard) and were analyzed using the RIA-AID software

(Robert Maciel Associates, Arlington, MA). The intra- and inter-assay

variability were 4.5% and 11.9% respectively (Van de Kar et al., 1985).

L. Statistics

The data are reported as mean ± S. E .M. (standard error of the

mean). The sample mean was the average of the number of animals in each

group (n). The standard error of the mean was calculated as described

below. Statistical analyses of the data was performed by a two-way

analysis of variance (ANOVA). Individual group means were compared by

Student Newman-Keul's multiple range test (Steel and Torrie, 1960).

The most common measurement of central tendency for given samples

of a population is the arithmetic average or mean. In addition to the

measure of central tendency, it is important to be aware of how the

samples differ in variability. This measure is the variance, or its

90

square root, the standard deviation (S.D.). This variance will provide

information as to whether an observation is an ordinary or an unusual

value in a specified population, giving a small value when the observa­

tions cluster closely about a central value, and a larger value when they

are spaced widely. The sum of squares is the sum of the squared

deviations from the arithmetic mean. The quantity (n-1) is known as

degrees of freedom. Sum of squares is divided by degrees of freedom to

give unbiased estimates of the population variances. Since sample means

are also subject to variation, the standard error of the mean is also

reported. It is calculated by dividing the S.D. by the square root of the

number of observations in the mean.

1. Analysis of Variance (ANOVA)

Analysis of variance is defined by Steel and Torrie (1960) as

" ... an arithmetic process for partitioning a total sum of squares into

components associated with recognized sources of variation." Using this

type of test, the data are classified according to their treatment group.

With respect to the present studies, a two-way analysis of variance was

the appropriate test, since the animals were subjected to two separate

treatments: (1) surgical (or drug) and (2) behavioral (stress).

The error mean square is an average of the components contributed

by the treatments. It is an estimate of a common variation among

observations of the same treatment group. Treatment mean square is an

independent estimate of the variance when the null hypothessis is true.

F is defined as the ratio of two independent estimates of the same

variance, thus, the F value is obtained by dividing the treatment mean

91

square by the error mean square. Once F has been calculated, it is

compared with the tabular F at the appropriate degrees of freedom at the

.05 (5%) probability level. If the calculated F is greater than the

tabulated F, then the null hypothesis (no difference between population

means) can be rejected and it can be concluded that there are real

differences among treatment means. With respect to the two-way ANOVAS of

the present studies, three F values were obtained: Fa for the surgical or

drug treatment, Fb for the stress treatment, and Fa x b for the interaction

between the drug and stress treatments. The F value reported was the one

that accounts for interaction between both the stress and surgical/drug

treatment groups.

2. Student-Newman-Keuls' Test

Multiple comparison tests can be used to compare each treatment

mean with every other treatment mean. Student-Newman-Keuls' test is a

multiple range test " ... for judging the significance of a set of

differences; it permits decisions as to which differences are significant

and which are not ... Error rate is seen to apply neither on an experiment­

wise nor a per-comparison basis since a particular range test may be made

for one or more comparisons in a particular experiment" (Steel and Torrie,

1960). This test also takes into account the number of treatments in the

experiment. To perform the Newman-Keuls' test, the mean square error from

the ANOVA is divided by the harmonic mean (average number of n per group).

The square root of this number is Sx, the standard error of the mean. Next

a tabular value, q, is selected. This value is based on a .05 level of

probability, and the appropriate p (number of treatment means) and degrees

92

of freedom. Q is multiplied by Sx to give a range of significance. If the

distance between two means is greater than the range of significance, then

the difference between the two means is significant at the 0.05 probabil­

ity level.

CHAPTER IV

RESULTS

A. Electrolytic Lesions in the PVN

Figure 1 is a coronal section of an intact PVN, stained with cresyl

violet. Figure 2 shows an electrolytic lesion in the PVN. These lesions

did not extend caudally beyond the PVN, and the dorsomedial nucleus was

intact. There was some damage to the caudal aspect of the anterior

hypothalamic area, as in Figure 2. This was not evident in all cases.

The lesions did not extend laterally to the fornix. Only lesions which

destroyed 90% or more of the PVN and did not extend dorsally into the

thalamus were considered in the statistical evaluation of hormonal data.

Following histological verification, the lesion control/group consisted

of five rats, while the lesion/stress group consisted of seven rats with

accurate and complete PVN lesions. One rat did not survive the anesthesia

in the sham/control group, making the total for this group seven rats.

Electrolytic lesions in the PVN prevented the stress-induced

increases in plasma renin activity (PRA) (Figure 3), plasma renin

concentration (PRC) (Figure 4), and in plasma corticosterone levels

(Figure 5). By comparison, rats with lesions which missed the PVN (n=S),

but destroyed areas rostral to the PVN, showed no inhibition of the effect

of stress on PRA (24.0 ± 3.2), PRC (33.7 ± 6.2) and corticosterone (29.3

± 6.1) levels.

93

FIGURE 1

CORONAL SECTION OF RAT BRAIN WITH AN INTACT PARAVENTRICULAR NUCLEUS

FIGURE 2

CORONAL SECTION OF RAT BRAIN WITH AN ELECTROLYTIC LESION IN THE PARAVENTRICULAR NUCLEUS

94

~ > 20 ~ /"\

L

0 £ <( I")

z """15 z E w """ 0::: <( 10 <( 01 2 c (/) '-.../

<( 5 _J o_

SHAM LESION

FIGURE 3

* D CONTROL ~ STRESS

ELECTROLYTIC PVN LESION

95

The effect of stress on plasma renin activity in rats with electrolytic PVN lesions. The data represent mean± S.E.M., n=7 for the sham/control group, n=6 for the sham/stress group, n=5 for the lesion/control group, and n=7 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 5.7 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the lesion F(l, 21) -10.789; stress F(l,21) = 14.617; and lesion*stress interaction F(l,21) -9. 031.

96

FIGURE 4

z 30 0 D CONTROL f-

25 * <( ~STRESS (l'.'. f-z "' w L 20 u _c

z "" -0 ~ 15 u

z <(

z ()> w c 10 (l'.'. \.._,./

<( 2 5 (/)

:5 0... 0

SHAM ELECTROLYTIC LESION PVN LESION

The effect of stress on plasma renin concentration in rats with electro­lytic PVN lesions. The data represent mean± S.E.M., n=7 for the sham/control group, n=6 for the sham/stress group, n=5 for the lesion/­control group, and n=7 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 4.5 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a highly significant effect of the lesion F(l,21) = 16.713; stress F(l,21) = 26.858; and lesion*stress interaction F(l,21) - 17.588.

w z 0 ()'.'. 25 w f---(j)

0 """' 20 u 1J

- "' f--- Ol ()'.'. ::t 0 'J 15 u <(

~ 10

s o_

FIGURE 5

SHAM LESION

* D CONTROL ~ STRESS

ELECTROLYTIC PVN LESION

97

The effect of stress on plasma corticosterone levels in rats with electrolytic PVN lesions. The data represent mean± S.E.M., n=7 for the sham/control group, n=6 for the sham/stress group, n=5 for the lesion/­control group, and n=7 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 7.0 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,21) = 11.06; and lesion*stress interaction F(l, 21) = 14. 245. The lesion did not produce a significant effect F(l,21) = 0.64.

98

B. Electrolyic Lesions in the Nucleus Reuniens

In a control experiment, lesions in the thalamic nucleus reuniens

were made, dorsal and caudal to the PVN. These lesions were small,

discrete and did not extend to the region of the third ventricle.and the

PVN (Figure 6). Following histological verification of placement in the

nucleus reuniens, five rats were included in the lesion/control and six

rats included in the lesion/stress group. There were six rats in the

sham/stress group, and eight rats in the sham/control group.

Electrolytic lesions in the nucleus reuniens did not prevent the

stress-induced increase in plasma renin activity (Figure 7) or plasma

renin concentration (Figure 8). The stress-induced increase in plasma

corticosterone levels was attenuated but not completely blocked by nucleus

reuniens lesions (Figure 9). It is possible that this lesion interrupted

afferent fibers to the PVN that are involved in stimulating CRF neurons

in the PVN. Stressed rats in which the lesions missed both the nucleus

reuniens and the PVN (n=S), had hormonal values similar to the sham/stress

group and the group in which the nucleus reuniens was destroyed (PRA 14.3

± 3.6; PRC 13.5 ± 1.8; and corticosterone 18.2 ± 1.7).

FIGURE 6

CORONAL SECTION OF RAT BRAIN WITH AN ELECTROLYTIC LESION IN THE THALAMIC NUCLEUS REUNIENS

99

>­f-> 20 f- ~

L u _c

<( n z ""' 15 z E w ""' 0::: <( 10 <( O'> 2 c ~ ~ 5 _J

Q_

FIGURE 7

D CONTROL

~ STRESS

SHAM LESION

*

ELECTROLYTIC LESION NUCLEUS

REUNIENS

100

The effect of stress on plasma renin activity in rats with electrolytic lesions in the nucleus reuniens. The data represent mean± S.E.M., n-8 for the sham/control group, n=6 for the sham/stress group, n=5 for the lesion/control group and n=6 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 6.4 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,21) = 11.58; but not of the lesion F(l,21) = 0.761 or lesion*stress interaction F(l,21) - 0.92.

z 0 I-<{ ~ 1-z~ W L

20

~ ~ 15 o E u """ z <{

z w 0:::

<{ 2 (/)

:5 CL

(J>

c 10

5

SHAM LESION

FIGURE 8

* D CONTROL

~STRESS

ELECTROLYTIC LESION NUCLEUS

REUNIENS

101

The effect of stress on plasma renin concentration in rats with electro­lytic lesions in the nucleus reuniens. The data represent mean± S.E.M., n=8 for the sham/control group, n=6 for the sham/stress group, n=S for the lesion/control group and n=6 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 7.4 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,21) = 10.291; but not of the lesion F(l,21) - 0.042 or lesion*stress interaction F(l,21) = 1.109.

w z 0 0::: 25 w 1-(f)

0 ~ 20 u ""O

I- "" 0::: (Jl

0 3 15 u <t

~ 10

s Q_

SHAM LESION

FIGURE 9

* D CONTROL r'Z2l STRESS

ELECTROLYTIC LESION NUCLEUS

REUNIENS

102

The effect of stress on plasma corticosterone levels in rats with electrolytic lesions in the nucleus reuniens. The data represent mean± S.E.M., n=7 for the sham/control group, n=6 for the sham/stress group, n=5 for the lesion/control group and n=6 for the lesion/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 6.5 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,20) - 10.29; but not of the lesion F(l,20) = 0.601 or lesion*stress interaction F(l,20) = 1. 83.

103

c. Injection of Ibotenic Acid into the PVN

Figure 10 is a section of an intact PVN stained with cresyl violet.

Figure 11 shows the needle tracks and destruction of neurons in the PVN

following an injection of ibotenic acid. Acceptable lesions were

restricted to the PVN and, compared with control sections, showed greater

than 50% cell loss. Following histological verification of the injection

site and evaluation of the hormonal data, there were seven rats included

in both the vehicle/control and vehicle/stress group. Four rats in the

ibotenic acid/control and six rats in the ibotenic acid/stress group were

considered accurate PVN lesions.

Ibotenic acid-induced lesions in the PVN prevented the stress­

induced increase in plasma renin activity (Figure 12) and plasma renin

concentration (Figure 13). Stress-induced increases in corticosterone

levels also were prevented by ibotenic acid lesions of neurons in the PVN

(Figure 14). By comparison, stressed rats with ibotenic acid injections

which missed or caused unnoticeable damage to neurons in the PVN (n=6) had

PRA (28.9 ± 3.8), PRC (35.5 ± 4.5) and corticosterone levels (16.9 ± 1.9)

which were indistinguishable from the vehicle/stress group. Thus far, the

data suggested that cell bodies in the PVN are involved in stress-induced

renin and corticosterone secretion.

104

FIGURE 10

INTACT PARAVENTRICULAR NUCLEUS

PARAVENTRICULAR NUCLEUS FOLLOWING BILATERAL INJECTION OF IBOTENIC ACID

105

~ 35 > - /'"""'\

0 ~ 30 <{ 6 z ..c 25 - t')

Q"' o::: E 20

<{ "' 2 <{ 15 (f) 0\ <{ c _J '-"' Q_ 10

*

VEHICLE

FIGURE 12

D CONTROL fZZJ STRESS

IBOTENIC ACID

106

The effect of stress on plasma renin activity in rats with ibotenic acid lesions in the PVN. The data represent mean ± S. E .M., n=7 for both vehicle groups, n=4 for the ibotenic acid/control group, and n=6 for the ibotenic acid/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 10.5 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the drug F(l,20) = 4.959; stress F(l,20) = 9.488; and drug*stress interaction F(l,20) = 6.063.

z 0

~ er 1-z /"-w L

40

u :J

z ~ 30

0 "" u -E

~ "" 20 w <( er (JI

c ~ '-/ 10 (/) <( _J

o_

*

VEHICLE

FIGURE 13

D CONTROL ~ STRESS

IBOTENIC ACID

107

The effect of stress on plasma renin concentration in rats with ibotenic acid lesions in the PVN. The data represent mean± S.E.M., n=7 for both vehicle/control and vehicle/stress groups, n=4 for the ibotenic acid/­control group, and n=6 for the ibotenic acid/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 13.5 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the drug F(l,20) = 5.063; and stress F(l,20) = 8.069. The drug*stress interaction F(l,20) - 3.619 was not significant.

w z 0 tr 15 w 1-(/) o; u " I- Ol 10 i:r ::i. 0 ....._.,, u <(

2 (/) <( _J

()_

5

*

VEHICLE

FIGURE 14

D CONTROL f7ZJ STRESS

IBOTENIC ACID

108

The effect of stress on plasma corticosterone levels in rats with ibotenic acid lesions in the PVN. The data represent mean± S.E.M., n=7 for both vehicle/control and vehicle/stress, n=4 for the ibotenic acid/control group, and n=6 for the ibotenic acid/stress group.

* Significant difference from the corresponding control group, p < 0.05. A minimal difference of 4.3 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the drug F(l,21) = 12.273; and stress F(l, 21) 19. 645. The drug*stress interaction was not significant F(l,21) = 3.198.

109

o. Injection of 6-0HDA into the PVN

Figure 15A shows a coronal section of the PVN following vehicle

injection and staining with dopamine ~-hydroxylase (DBH) antiserum.

Figure 15B shows a coronal section of the PVN following 6-0HDA injection

and staining with DBH antiserum, showing loss of DBH immunoreactivity.

In general, there was a 60-80% reduction in DBH staining (as compared to

vehicle rats) in a 500-1000 micron area that encompassed the PVN and

anterior hypothalamus, but did not spread as far as the mediobasal

hypothalamus. Following immunocytochemical and histological verification,

the vehicle/control and 6-0HDA/control groups consisted of seven rats

each, the vehicle/stress group consisted of six rats, and the 6-0H­

DA/stress group had five rats.

The stress (GER) procedure produced a significant increase in plasma

renin acivity and concentration as well as increasing plasma corticos­

terone levels in rats which received an injection of vehicle or non­

injected animals. Injection of 6-0HDA into the PVN prevented the effect

of stress on plasma renin activity (Figure 16) and plasma renin concentra-

ti on (Figure 17) . In addition the effect of stress on plasma cor-

ticosterone levels was also blocked (Figure 18). Since 6-0HDA injections

into the PVN prevented the stress - induced increase in corticosterone

levels, it is possible that this caused decreased secretion of the renin

substrate (angiotensinogen) from the liver. This would cause PRA values

to be low even though renin levels were unaffected. To test this, in

110

addition to measurements of PRC, plasma angiotensinogen levels were

determined (Figure 19). There were no differences observed between the

groups. This suggested that neither stress nor 6-0HDA injections affected

plasma levels of renin substrate. Thus far the data suggest that

catecholaminergic neurons or nerve terminals in the PVN play a role in the

stress-induced increase in renin and corticosterone secretion.

FIGURE lSA

CORONAL SECTION FROM RAT INJECTED WITH VEHICLE AND STAINED WITH DBH ANTISERUM

CORONAL SECTION FROM RAT INJECTED WITH 6-0HDA AND STAINED WITH DBH ANTISERUM,

SHOWING LOSS OF DBH IMMUNOREACTIVITY

111

>­I-> - /'""'\ I- Cf)

25

0 ~ 20 <( 0 z _c

z I") 15 w"" o: E <( "" 10 2 <t (f) Q)

<( c 5 _J '-.,,/

Q_

*

VEHICLE

FIGURE 16

D CONTROL ~ STRESS

6-0HDA

112

The effect of stress on plasma renin activity in rats with 6-0HDA lesions. The data represent mean ± S.E.M.; n=7 for vehicle/control, n=6 for vehicle/stress, n=7 for 6-0HDA/control and n=5 for 6-0HDA/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 5.48 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the drug F(l,21) = 13.576; stress F(l,21) = 35.595; and drug*stress interaction F(l,21) = 12.427.

z 0

~ O'.'. f-

50

z /"­w L

0 g 40 z .r::

8 ""' 30 z E - ""' z -w <{ 20 O'.'. (Jl

<{ ~ 2 (/)

s 0...

10

*

VEHICLE

FIGURE 17

D CONTROL ~ STRESS

6-0HDA

113

The effect of stress on plasma renin concentration in rats with 6-0HDA lesions. The data represent mean± S.E.M.; n=7 for vehicle/control, n-6 for vehicle/stress, n=7 for 6-0HDA/control and n-5 for 6-0HDA/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 12.9 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of the drug F(l,21) = 6.939; stress F(l,21) = 24.623; and drug*stress interaction F(l,21) = 6.059.

~ 30 0 O'.'. w 25 I-(/)

0 -; 20 u "" I- Ol

~ 3 15 u

~ 10 (/) <t: _J 5 D...

*

VEHICLE

FIGURE 18

D CONTROL f7ZJ STRESS

6-0HDA

114

The effect of stress on plasma corticosterone levels in rats with 6-0HDA lesions. The data represent mean± S.E.M.; n-7 for vehicle/control, n-6 for vehicle/stress, n=7 for 6-0HDA/control and n-5 for 6-0HDA/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 10.0 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,21) = 9.422. There were no significant effects of either the drug F(l,21) = 0.194 or drug*stress interaction F(l,21) = 3.371.

FIGURE 19

120 ~----------~

z w (.'.)

0 z,,.......

100

(/) ~ 80 z"' w-1- E 0 .-- 60 (.'.) . zo <{ ci; <{ 5 40 2 (/) s 20 Q_

VEHICLE

D CONTROL f7ZJ STRESS

6-0HDA

115

The effect of stress on plasma angiotensinogen levels in rats with 6-0HDA lesions. The data represent mean± S.E.M.; n-7 for vehicle/­control, n=6 for vehicle/stress, n=7 for 6-0HDA/control and n=5 for 6-0HDA/stress.

No significant differences found among the groups, p < 0.05. A minimal difference of 33.l would be required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates no significant effects of the drug F(l, 21) -1. 395; stress F(l, 21) = 0. 072; and drug*stress interaction F(l, 21) -0. 721.

116

E. Injection of Sotalol into the PVN

Figure 20 shows a coronal section of the PVN stained with cresyl

violet following a bilateral injection of sotalol into the PVN. Only rats

in which the tip of the cannula was immediately dorsal to the border of

the PVN, or inside the dorsal PVN, were acceptable. In several rats, the

cannulae ended inside the PVN, producing large mechanical lesions. These

rats were excluded from the statistical analysis of the hormonal data.

Following histological verification of the injection site, eight rats were

used in the vehicle/control, five rats in vehicle/stress and non­

injected/stress, and four rats in the non-injected/control group. There

were six rats used in the sotalol/control group and eight rats in the

sotalol/stress group.

Figures 21 and 22 show the effect of stress on plasma renin activity

and plasma renin concentration, respectively, in rats injected with either

vehicle or sotalol, or in non-injected rats. Stress-induced increases in

PRA and PRC were not prevented by injection of sotalol. However, the

stress-induced increase in corticosterone levels was prevented by

injections of sotalol into the PVN (Figure 23).

FIGURE 20

CORONAL SECTION OF THE PARAVENTRICULAR NUCLEUS FOLLOWING BILATERAL INJECTION OF SOTALOL

117

>­~

> ~,....._ OL <( .c

z~ ZE w"" (t <(

JO

25

20

15 <( Ol

25 10 (/) <( _J 5 ()_

*

VEHICLE

FIGURE 21

D CONTROL ~STRESS

*

SOTALOL

118

*

UNINJECTED

The effect of stress on plasma renin activity in rats injected with either vehicle or sotalol, or in non-injected rats. The data represent mean± S.E.M.; n=8 for vehicle/control, n=4 for vehicle/stress, n=6 for sotalol/control, n=7 for sotalol/stress, n=4 for non-injected/control and n=5 for non-injected/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 8.42 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,28) - 32.526; but not of the drug F(2,28) = 0.871 or drug*stress interaction F(2,28) -0.528.

z 0 ~ 25 O'.: f-,........

~ ~ 20 0 z r-0~ o E 15 z"'--<( z WOl 10 O'.: c <( 2

...._,,,,

(f) 5 ~ o_

FIGURE 22

*

VEHICLE SOTALOL

D CONTROL ~ STRESS

*

UNINJECTED

119

The effect of stress on plasma renin concentration in rats injected with either vehicle or sotalol, or in non-injected rats. The data represent mean± S.E.M.; n=8 for vehicle/control, n=4 for vehicle/stress, n=6 for sotalol/control, n=7 for sotalol/stress, n=4 for non-injected/control and n=5 for non-injected/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 9.6 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,28) - 17.357; but not of the drug F(2,28) = 0.249 or drug*stress interaction F(2,28) -0.553.

FIGURE 23

35....---------------,

~ 30 0 O'.'. w I-(/) 0 25 Q""CJ u"'-- (])

~3 20 0 u <{

2 15 (/)

s 0...

*

VEHICLE

D CONTROL

~STRESS

SOTALOL

*

UN INJECTED

120

The effect of stress on plasma corticosterone levels in rats injected with either vehicle or sotalol, or in non-injected rats. The data represent mean± S.E.M.; n=8 for vehicle/control, n=4 for vehicle/stress, n-6 for sotalol/control, n=7 for sotalol/stress, n=4 for non-injected/control and n=S for non-injected/stress.

*Significant difference from the corresponding control group, p < 0.05. A minimal difference of 8.6 is required between the means by the Student­Newman-Keuls' test.

Two-way ANOVA indicates a significant effect of stress F(l,28) = 9.823. No significant effect is seen with the drug F(2,28) - 1.724 or drug*stress interaction F(2,28) = 1.411.

121

F. Comparison of Normal Plasma Pools to Experimental Controls

Pooled plasma samples for "normal" and "high" PRA or "normal" and

"high" corticosterone levels are run routinely with each respective assay.

The pooled plasma samples are compiled from rat plasma that has been

previously assayed and classified within the range of "normal" or "high"

based on its corticosterone or PRA value. A random sample of these

values, taken from assays run between 1986 and 1989, gave the following

values (mean± standard error of the mean):

"normal" PRA 7.6 ± 0.5 (n 23) "high" PRA 23.6 ± 2.8 (n 23)

"normal" corticosterone 8.9 ± 1.1 (n 13) "high" corticosterone 20.9 ± 2.0 (n 10)

The pooled values for "normal" PRA and corticosterone are very

similar to previously published values for PRA and corticosterone in

untested control animals, shown below.

PRA Corticosterone

Untested Control Rats

7.1 ± 1.2 8.9 ± 2.1

(Van de Kar et al., 1985) (Van de Kar et al., 1985)

The normal plasma pool values were compared to the means of the

sham- or vehicle-treated groups in each of the experiments of this

dissertation. ANOVA and Newman-Keuls' Multiple Range Test were run to

determine if significant differences between the groups exist. With

respect to plasma renin activity, the data are shown on the following

page:

Group Number

4

1

5

2

6

3

Table 1.

Comparison of Mean Values: Normal PRA Pool versus Experimental

Sham- or Vehicle-Treated Groups

Description Mean + Std.Error

6-0HDA Expt. Vehicle/Control 3.6 ± 0.8

Electrolytic PVN Sham/Control 4.5 ± 0.6

Sotalol Expt. Vehicle/Control 6.3 ± 0.5 + Non-injected/Control

Electrolyt. N. Reun. Sham/Cont. 6.4 ± 0.4

Normal PRA Pool 7.6 ± 0.5

Ibotenic Expt. Vehicle/Control 9.2 ± 0.9

Standard Error of Treatment Means - 0.74

122

n

7

7

12

8

23

7

123

Table 2

Newman-Keuls' Multiple Range Test for PRA

Treatment vs. Treatment Difference Sig .OS

Group 4 Group 1 0.94

Group 4 Group 5 2. 71 * Group 4 Group 2 2.82 * Group 4 Group 6 4.06 * Group 4 Group 3 5.62 * Group 1 Group 5 1. 77

Group 1 Group 2 1. 88

Group 1 Group 6 3.11 * Group 1 Group 3 4.68 * Group 5 Group 2 0.11

Group 5 Group 6 1. 34

Group 5 Group 3 2.91 * Group 2 Group 6 1. 23

Group 2 Group 3 2.80 * Group 6 Group 3 1.56

124

The means of the experimental vehicle- or sham-treated goups were

not significantly elevated above the normal PRA plasma pool. However,

two experimental groups had PRA values lower than the normal PRA pool

levels.

With respect to corticosterone, the data are as follows:

Group Number

3

4

6

1

2

5

Table 3

Comparison of Mean Values: Normal Corticosterone Pool versus Experimental

Sham- or Vehicle-Treated Groups

Description Mean + Std.Error

Ibotenic Expt. Vehicle /Control 6.5 ± 0.5

6-0HDA Expt. Vehicle/Control 7.1 ± 0.5

Normal Corts Pool 8.9 ± 1.1

Electrolyt. PVN Sham/Control 9.3 ± 2.3

Electrolyt. N. Reun. Sham/Cont. 12.8 ± 0.9

Sotalol Expt. Vehicle/Control 12.9 ± 1.2 + Non-injected Control

Standard Error of Treatment Means= 0.74

!!

7

6

13

7

7

12

125

Table 4

Newrnan-Keuls' Multiple Range Test for Corticosterone

Treatment vs. Treatment Difference Sig .05

Group 3 Group 4 0.49

Group 3 Group 6 2.36

Group 3 Group 1 2.75

Group 3 Group 2 6.31 * Group 3 Group 5 6.34 * Group 4 Group 6 1. 86

Group 4 Group 1 2.26

Group 4 Group 2 5.81 * Group 4 Group 5 5.85 * Group 6 Group 1 0.39

Group 6 Group 2 3.95

Group 6 Group 5 3.98

Group 1 Group 2 3.55

Group 1 Group 5 3.58

Group 2 Group 5 0.03

None of the vehicle- or sham-treated experimental groups were

significantly different from the normal corticosterone plasma pool.

CHAPTER V

DISCUSSION

A. Overview

The following conclusions can be drawn from the results presented:

(1) Neurons in the PVN mediate the stress-induced increase in renin and

corticosterone secretion. (2) Catecholaminergic nerve terminals in the

PVN mediate the effect of stress on renin and corticosterone secretion.

(3) Stress-induced corticosterone secretion appears to be mediated by ~

receptors in the PVN, while (4) stress-induced renin secretion is mediated

by a receptor that is not ~-adrenergic. Stress-induced renin secretion

may be mediated by a different catecholamine receptor, a neuropeptide or

other substance(s) colocalized in catecholaminergic nerve terminals.

B. Justification of the Stress Paradigm

Abundant evidence was presented in the literature review to

demonstrate that renin and corticosterone secretion can be elevated by

many different stressors. The stress paradigm used in these experiments

was a conditioned fear paradigm that involved exposing the rats to an

initially neutral stimulus (the stress chamber) which was followed after

ten minutes by an aversive stimulus (footshock). The latter unconditioned

stimulus elicited a range of unconditioned responses, including neuroen­

docrine (increased renin and corticosterone levels) and behavioral

(defecation, urination, freezing) responses. As the pairings of stimuli

126

127

were continued, the originally neutral stimulus became a conditioned

stimulus, capable of eliciting the same responses as the unconditioned

stimulus.

1. Classical Conditioning of Fear Stimuli

Classical conditioning can account for the wide range of stimuli

that can produce fear in animals and humans (Watson, 1924). An experiment

by Hunt and Otis (1953) demonstrated that one symptom of fear, defecation,

can be classically conditioned. They presented a rat with a flashing

light for three minutes, at the end of which time the animal received an

electric foot shock. Eight conditioning trials were run. At first,

defecation occurred primarily when the shock was delivered, but it

increasingly began to occur during the presentation of the flashing light.

A similar phenomenon was observed with the conditioned fear

paradigm used in the experiments of this dissertation. Behavioral

observations were made throughout the testing procedure, and recorded on

the third and fourth (final) days. Rats were monitored for defecation

and urination, as well as exploratory, freezing or grooming behaviors.

On the first day, the rats were placed in the stress chamber for a ten

minute interval, at the end of which time they received a mild foot shock.

The rats were tested in this manner for four consecutive days, the only

exception being that on the final day, no foot shock was given. On the

first day in the stress chamber the rats did not defecate until the time

of the foot shock. However by the third day (and in some cases the

second), as soon as the rats were placed in the stress chamber they began

128

to defecate. The defecation suggested that the rats had paired the

placement in the chamber with the impending foot shock, supporting the

hypothesis that classical conditioning is an integral part of the ability

of previously neutral stimuli to elicit fear. Lorens et al. (1989)

reported the number of fecal boli as a defecation score, a quantitative

measurement of a behavioral response to conditioned fear.

In the experiments of this dissertation, the control animals

explored the test chamber and groomed themselves. The stressed animals,

after the first or second day of conditioning, would sometimes explore,

and sometimes groom, but for the most part would defecate, urinate and

exhibit freezing behavior. Freezing behavior consists of tense and silent

immobility, and is seen in many species in times of danger. The

observation of freezing behavior in response to the CS represents another

similarity between my experiments and those presented by Hunt and Otis

(1953). Normally a rat responds to an electric foot shock with a great

increase in activity - running around frantically, jumping, and trying to

escape the shock. However, in my experiments and the experiments of Hunt

and Otis (1953), the stimulus followed by shock caused a radically

different behavior - immobility. This is quite an exception to the usual

pattern of the CS (in my case, the stress chamber) eliciting a similar

(though not exact) reaction to that of the UCS (the footshock). Gray

(1971) suggested the possibility that freezing and more active forms of

escape and avoidance behavior are controlled by separate brain systems,

and represent a combination of learning and innate behavior.

The process of producing electrolytic lesions or injecting

neurotoxins is an invasive procedure that could, theoretically, interrupt

129

fiber pathways that are involved in the learning/conditioning process.

Because this surgical intervention occurs two weeks prior to the

behavioral conditioning paradigm, it is important to look for potential

differences in the behavior of sham or vehicle-injected rats compared with

rats with electrolytic or chemical lesions in the PVN. Behavioral

observations during the conditioning process indicated that the lesions

did not prevent defecation, urination and freezing in the stressed rats.

This suggests that the lesions did not interfere with the conditioning

process.

C. Electrolytic and Ibotenic Acid Lesions in the PVN

The data suggest that neurons in the PVN mediate the effect of

stress on renin secretion and confirm results by others (Rivier and Vale,

1983; Dohanics et al., 1986) that neurons in the PVN mediate the effect

of stress on corticosterone secretion. This conclusion is based on the

finding that both complete electrolytic lesions in the PVN and selective

destruction of cells in the PVN prevented the stress-induced elevation in

plasma renin activity, plasma renin concentration and plasma cort­

icosterone levels, whereas electrolytic lesions in the thalamic nucleus

reuniens were ineffective. Ibotenic acid was selected for its ability to

selectively destroy cell bodies while leaving fibers of passage intact

(Markowska et al., 1985; Winn et al., 1984). Magnocellular neurons in the

PVN have been reported to be resistant to the neurotoxic effects of kainic

acid (Zhang and Ciriello, 1985). Herman and Wiegand (1986) have reported

that ibotenic acid selectively destroyed parvocellular neurons in a dose­

dependent manner while sparing magnocellular neurons. They reported that

130

loss of parvocellular neurons occurred at 0.5, 0.75, 1.0 and 2.0 µg doses

of ibotenic acid. Loss of magnocellular neurons (mostly vasopressinergic)

was observed only after the 2 .0 µg dose. Their protocol involved

unilateral injections (compared to our protocol of bilateral injections)

directed towards the central portion of the PVN at the level of the

posterior magnocellular division. In our study, the drug volume was

higher (0.3 µl compared to 0.1 µl) and the injection site had to be at

least 0.5 mm lateral to the midsagittal suture. Injections of ibotenic

acid closer to the midline led to diffusion of the drug into the third

ventricle, resulting in death of the rats. We did not observe selective

sparing of the magnocellular neurons after injections of ibotenic acid.

The effect of stress on corticosterone secretion is most likely

propagated by stimulation of CRF neurons in the PVN. The majority of the

diencephalic CRF-neurons are located in the parvocellular neurons in the

dorsal and medial subnuclei of the PVN (Liposits et al., 1987; Liposits

and Paull, 1985). A small number of CRF-neurons also are present in the

magnocellular divisions of the nucleus (Swanson and Sawchenko, 1983;

Piekut and Joseph, 1986). Dohanics et al. (1986) reported that the full

corticosterone and ACTH response to ether or ether plus laparotomy stress

requires not only an intact PVN, but also an intact neurointermediate

lobe. Therefore, our data are consistent with the hypothesis that CRF

neurons in the parvocellular PVN are activated during the CER procedure

to stimulate ACTH and thus corticosterone secretion. Furthermore, the

data with corticosterone support the histological verification of the

accuracy of the lesions in the PVN. The finding that basal corticosterone

concentration was not altered by the PVN lesions is in agreement with

131

previous studies by Meyerhoff et al. (1987) and Makara et al. (1986), who

reasoned that in the absence of CRF input, the cells in the pituitary can

maintain a basal level of ACTH but cannot respond to stimuli which would

increase ACTH secretion. Similar findings were also reported using CRF

antisera (Nakani et al., 1985; Ono et al., 1985; Rivier et al., 1982).

Nucleus reuniens lesions inhibited the effect of stress on

corticosterone but not renin secretion. This suggests that fibers which

course through the nucleus reuniens and innervate the CRF neurons in the

PVN may have been disconnected by this lesion. Since no inhibition of

stress-induced renin secretion occurred in these rats, the data suggest

that the fibers, and possibly cells, which mediate the effect of stress

on renin secretion are different from the fibers and cells which mediate

stress-induced corticosterone secretion.

The nature of the PVN neurons which mediate the effect of stress

on renin secretion is not known. It is clear, both from the present

results and from studies by Porter (1988), that cells in the PVN play an

important role in the regulation of renin secretion. Electrolytic lesions

in the PVN prevent the increase in plasma renin activity after injection

of the serotonin releasing drug p-chloroamphetamine (Gotoh et al., 1987).

Furthermore, these lesions also prevent the increase in plasma renin

activity that follows immobilization stress, although a reduction in

plasma renin substrate (angiotensinogen) levels could account for this

effect (Gotoh et al., 1987). It is important to note that, in the present

study, a saturating concentration of exogenous renin substrate was added

in the plasma renin concentration assay to control for possible variation

of renin substrate in the plasma samples from the experimental rats. In

132

contrast, the plasma renin activity assay was a measure of the production

of angiotensin I from endogenous renin, combined with endogenous renin

substrate. Since plasma renin activity and concentration were equally

affected by the lesions, the changes were clearly due to changes in renin,

not renin substrate concentration in plasma.

D. Injection of 6-0HDA into the PVN

The main conclusion of the 6-0HDA experiment is that catecholaminer­

gic nerve terminals in the PVN mediate the effect of stress on both renin

and corticosterone secretion. This conclusion is based on the observation

that injection of 6-0HDA into the PVN inhibited the stress-induced

increase in renin and corticosterone secretion. 6-0HDA is a neurotoxin

that is taken up by the catecholaminergic nerve endings and produces

degeneration of the nerve terminals (Breese and Traylor, 1970; Uretsky and

Iversen, 1970). 6-0HDA is considered to be selective for catecholaminer­

gic neurons. Uretsky and Iversen (1970) reported that doses of 6-0HDA

that decreased brain norepinephrine and dopamine levels did not affect

brain serotonin or GABA. Shor-Posner et al. (1986) have reported that

injection of 6-0HDA into the PVN, in the same dose as in my experiment,

produced a 75% reduction in norepinephrine, 56% reduction of epinephrine

and 47% reduction in dopamine concentration in the PVN. The catechol­

amines were determined by HPLC and electrochemical detection. Histochemi­

cal analysis also revealed a significant decrease in catecholamine

fluorescence in the PVN. This study suggests that 6-0HDA can cause a

degeneration of noradrenergic, adrenergic and dopaminergic elements in the

PVN. Further support was provided by Fety et al. (1984), who injected 6-

0HDA into the fourth ventricle of rats to study the biochemical reponses

133

of the adrenergic (Cl, C2) cell groups versus the noradrenergic (Al, A2

and A6) cell groups. Their results indicated that the adrenergic cell

groups are sensitive to the neurotoxic effects of 6-0HDA since they

exhibited changes in tyrosine hydroxylase and dopamine-&-hydroxylase in

a similar manner to the noradrenergic cell groups.

In the 6-0HDA experiment of this dissertation, norepinephrine

depletion was verified immunocytochemically by using an antibody for

dopamine-&-hydroxylase. No attempt was made to differentiate between loss

of norepinephrine-, dopamine- or epinephrine-containing nerve terminals.

Consequently, these results cannot indicate which catecholamine mediates

the effects of stress on renin or corticosterone secretion.

The suggestion that catecholaminergic nerve terminals in the PVN

mediate stress-induced corticosterone secretion is supported by studies

involving the ventral noradrenergic bundle (VNAB), which conveys most of

the noradrenergic (Moore and Bloom, 1979; Palkovits, 1981; Swanson and

Sawchenko, 1980) and adrenergic (Ungerstedt, 1971) afferents to the PVN.

Injections of 6-0HDA into the VNAB inhibited the ACTH stress response

(Szafarczyk et al. , 1985), prevented the increase in corticosterone

secretion following photic, acoustic and sciatic nerve stimulation

(Feldman et al., 1988), and significantly reduced CRF levels in hypophys­

ial portal blood (Guillaume et al., 1987). In addition, Feldman et al.

(1986) demonstrated that 6-0HDA injections into the PVN significantly

inhibit the adrenal response to photic, acoustic and sciatic nerve

stimulation. Together, these results suggest a role for catecholaminergic

afferents to the PVN in the activation of adrenocortical responses to

stress.

134

The dorsal noradrenergic bundle (DNAB) may play a role as a

mediator of stress-induced renin secretion. Some afferent fibers from

the locus coeruleus travel to the PVN in the DNAB (Kobayashi et al., 1975;

Loizou, 1969). Lightman et al. (1984) demonstrated that injections of 6-

0HDA into the DNAB markedly attenuate the increase in plasma renin

activity in response to hemorrhage, while injections of 6-0HDA into the

VNAB had no effect. This study, in combination with my results, suggests

that catecholaminergic afferents to the PVN mediate the renin response to

stress.

E. Injection of Sotalol into the PVN

Since the data suggest that catecholaminergic nerve terminals in

the PVN mediate the stress-induced increase in renin and corticosterone

secretion, the next question to address was whether or not E-adrenergic

receptors mediate this effect. Previous studies demonstrated that the &­

adrenoceptor antagonist propranolol (1 mg/kg, i.p.) significantly

attenuated the stress-induced increase in plasma renin activity (Van de

Kar et al., 1985). Since propranolol is known to cross the blood-brain

barrier (Garvey and Ram, 1975; Lemmer et al., 1985), and the PVN contains

a high concentration of~ receptors (Wanaka et al., 1989), it is possible

that a E-receptor, located in the PVN, could be involved in mediating the

effect of stress on renin and corticosterone secretion. In the present

experiment, sotalol was selected instead of propranolol because it is

devoid of local anesthetic activity and is both a &1 and E2 antagonist

(Fitzgerald, 1984). Injection of sotalol into the PVN prevented the

stress-induced increase in corticosterone secretion but did not prevent

135

the stress-induced increase in renin secretion. This suggests that a &­

adrenergic receptor in the PVN plays a role in the corticosterone response

to stress. However, many B-adrenoceptor antagonists also bind to

serotonin receptors in the brain (Conway et al., 1978). Since only one

dose of sotalol was administered, it is possible that this dose (15 nmol)

could affect serotonergic as well as B-adrenergic receptors. Other

investigators have injected much higher doses of sotalol and propranolol

in the brain. Injections of sotalol (400 nmol) or propranolol (120 nmol)

into the perif ornical region of the anterior hypothalamus did not prevent

the feeding response induced by either epinephrine (20 nmol) or norepi­

nephrine (20 nmol). This feeding response was prevented by injections of

the a-adrenoceptor antagonists phentolamine (30 or 60 nmol), tolazoline

(100 nmol) or phenoxybenzamine (15 nmol) in the same brain region

(Leibowitz, 1975). Injection of propranolol (116 nmol) or timolol (95

nmol) into the cerebral ventricles prevented the effect of air stress on

renal sympathetic nerve activity in spontaneously hypertensive rats. In

addition, icv injection of propranolol (4 or 25 nmol) produced an

inhibition of epinephrine-stimulated ACTH release (Szafarczyk et al.,

1987). In my study, the fact that sotalol inhibited the stress-induced

increase in corticosterone secretion but not the stress-induced increase

in renin secretion argues for a relative selectivity in receptor binding.

Other investigators have provided evidence for the role of &-

receptors in the regulation of corticosterone secretion. Szafarczyk et

al. (1987) demonstrated a stimulatory role for epinephrine and norepineph-

rine on corticosterone release. Injection (icv) of epinephrine or

norepinehrine induced ACTH release comparable to that occuring under

136

(ether) stress conditions. Epinephrine was more efficacious than

norepinephrine on an equimolar basis. The stimulating effect of

norepinephrine was reversed by icv pretreatment with prazosin, an a 1

antagonist, whereas the effect of epinephrine was inhibited by either

prazosin or propranolol. These results suggest that norepinephrine

stimulates ACTH release via a 1 receptors, whereas epinephrine acts via both

a 1- and &-adrenoceptors. In vitro studies support the role of &-adrenergic

receptors in CRF release. Norepinephrine stimulated the release of

immunoreactive CRF from cultured neonatal rat hypothalamic cells. This

effect was blocked by propranolol but not by prazosin. In addition, the

E.-adrenoceptor agonist isoproterenol significantly increased CRF while the

a-adrenoceptor agonist phenylephrine was ineffective, except at high

concentrations (Widmaier et al., 1989). Together, these data strongly

suggest a stimulatory role for &-receptors in the regulation of cor­

ticosterone secretion.

To speculate as to which neurotransmitter could be involved in

mediating the effect of stress on corticosterone secretion, several

studies should be considered. Leibowitz et al. (1986) demonstrated that

norepinephrine injections into the PVN dose-dependently increased serum

corticosterone levels, while dopamine had no effect. Epinephrine in the

PVN was even more efficacious than norepinephrine in increasing serum

corticosterone levels. Further, a study by Spinedi et al. (1988)

demonstrated that central inhibition of phenylethanolamine-N-methyltrans­

ferase (PNMT), producing a selective decrease in hypothalamic epinephrine

levels, significantly decreased the plasma ACTH response to ether stress,

and at later times also caused a selective decrease in CRF content. These

137

results support a stimulatory role of central epinephrine-containing

neurons in CRF release. Sawchenko et al. (1987) estimated that at least

40% of the retrogradely- labeled catecholaminergic cell groups that project

to the PVN could be adrenergic (PNMT-immunoreactive). These results

suggest that stress-induced increases in corticosterone levels could be

mediated via adrenergic nerve terminals activating B-receptors in the PVN.

Stress-induced increases in plasma renin activity and concentration

are mediated by a different receptor. It is possible that an a-adrenergic

receptor could mediate the stress-induced increase in renin secretion.

However, pharmacological studies suggest that activation of brain a 2-

adrenoceptors lowers renin secretion by inhibiting sympathetic outflow to

the kidneys. Injection of clonidine (icv) prevented the air jet stress­

induced increase in renal sympathetic nerve activity in spontaneously

hypertensive rats. This effect was reversed by icv injection of the a 2-

adrenergic antagonists yohimbine and rauwolscine (Koepke and DiBona,

1986). !CV injection of clonidine lowered plasma renin activity at a dose

which is ineffective when given intravenously (Reid et al., 1975). Blair

et al. (1977) demonstrated that administration of L-DOPA to dogs

pretreated with the peripheral decarboxylase inhibitor carbidopa results

in a drop in blood pressure and an inhibition of renin secretion. Since

this treatment increases brain catecholamine content, the results suggest

that catecholamines formed from L-DOPA can act within the central nervous

system to cause a decrease in renin secretion. In this study the decrease

in plasma renin activity was dependent upon the presence of the renal

nerves. These studies suggest that central a 2-adrenoceptors are involved

in decreasing renin secretion. However, it is possible that central a1

-

138

adrenoceptors could mediate the stress-induced increase in renin

secretion. Ganong et al. (1982) suggested that the activation of central

a 1 receptors increases renin secretion in anesthetized dogs. However,

these data are not very convincing since they only reported renin values

following icv injection of the a 1-agonists methoxamine and phenylephrine,

without the appropriate vehicle controls.

The PVN has been shown to project to those segments of the thoracic

cord containing sympathetic preganglioninc neurons innervating the kidney

(Swanson and Mc Kellar, 1979). Other studies have suggested that the

sympathetic nervous system mediates the stress response from the brain to

the kidney (Golin et al., 1988). However, in our laboratory, (see STRESS:

Effect of Peripheral Sympathectomy) sympathectomized-adrenal enucleated

rats showed no inhibition of the stress-induced increase in plasma renin

activity/concentration or plasma corticosterone levels (Richardson Morton

et al., 1988). This suggests that the sympathetic nervous system and

adrenal catecholamines are not the only mediators of the stress-induced

increase in renin secretion. An alternate theory is that cell bodies in

the PVN are stimulated to secrete a renin-releasing factor to the

bloodstream to promote renin release from the kidneys (Van de Kar et al.,

1987).

If a and~ receptors in the PVN do not mediate stress-induced renin

secretion, another possibility is that a neuropeptide, coreleased at the

catecholaminergic nerve terminals in the PVN, mediates stress-induced

renin secretion. A likely candidate is neuropeptide Y, which is

coexpressed in noradrenergic and adrenergic brain stem cell groups and is

present in noradrenergic and adrenergic nerve terminals in the PVN

139

(Sawchenko et al., 1985). Intraventricular injection of NPY has been

shown to increase plasma ACTH and corticosterone secretion (Harfstrand et

al., 1987) and plasma angiotensin II levels (Harfstrand et al., 1986).

However, Fuxe et al. (1983) reported that intracisternal injection of NPY

causes hypotension, which could also increase renin secretion by

activation of the intra-renal stretch receptor (Fahri et al., 1982).

Yet another neuropeptide of interest is galanin, due to its

preferential coexistance with noradrenergic cell groups that project to

the PVN. Combined immunohistochemical-retrograde transport studies have

demonstrated galanin-immunoreactivity coexisting with dopamine-&­

hydroxylase immunoreactivity in the Al and A6 cell groups, with no

evidence for galanin coexistence in adrenergic (PNMT-immunoreactive)

neurons that project to the PVN. Galanin inputs were confined to the

anterior, periventricular, dorsal and ventral subdivisions of the

parvocellular PVN, with only a sparse projection to the oxytocinergic

neurons of the magnocellular division of the PVN (Levin et al, 1987).

F. Hypothetical Organization of Brain Pathways Mediating

Stress-Induced Renin and Corticosterone Secretion

The neural pathways mediating stress-induced renin and cort­

icosterone secretion are likely to be complex and multi-synaptic, due to

the many inputs which must be integrated. First, there is a sensory

component that arises during behavioral conditioning. When the rat

receives the foot shock as part of the conditioned fear paradigm, the

sensory message is transmitted from the nerve fibers in the skin to the

spinal cord, where it travels in the lateral spinothalamic tract to the

140

posteroventral nuclei of the thalamus. The spinothalamic fibers are

accompanied by fibers that terminate in the reticular formation (spino­

reticular) and the periaqueductal gray and tectum (spinotectal). The

periaqueductal grey has been shown to play a role in flight or defensive

behavior (de Molina and Hunsperger, 1962).

The pain input route most likely continues from the midbrain

reticular formation, through the thalamus, to reach the septum and

hippocampus. The septo-hippocampal system plays a major role in pain

reception. The hippocampus emits a theta wave (4-9 cps rhythm) as a

response to a painful stimulus in the rat, and in the emotional states of

disappointment and frustration in humans. Lesions in the medial nuclei

of the septum and the thalamic nucleus centrum medianum abolish the theta

wave emitted by the hippocampus in response to a painful stimulus (Gray,

1971).

Damage to the hippocampus has been shown to facilitate acquisition

of shuttle box avoidance (Green et al., 1967; Van Hoesen et al., 1972)

possibly due to a deficit in a behavioral inhibitory system (Douglas,

1967; Gray, 1982). These studies, and others (Squire and Zola-Morgan,

1983), suggest a critical role of the hippocampal formation in learning

and memory. The hippocampal formation and septa! region share extensive

bidirectional connections. The lateral, medial and posterior divisions

of the septa! regions are associated with the hippocampal formation while

the ventral division (bed nuclei) is primarily related to the amygdala

(Swanson et al., 1987). The subicular complex and the entorhinal area of

the hippocampus receive an extensive sensory input from the cortex,

including olfactory, visual, auditory and somatosensory. Additional

141

sensory information is relayed through two routes, one via the thalamus,

extrahippocampal cortex and amygdala and the other via ascending pathways

from the brainstem and basal forebrain. The hippocampus also receives

considerable serotonergic input from the raphe nuclei (Azmitia, 1978;

Lorens and Guldberg, 1974; Steinbusch, 1981). Immunoreactive-CRF neurons

have been demonstrated in the hippocampus (Swanson, 1983). However, the

axonal projections of these cells are relatively short, suggesting that

they might function as interneurons (Kohler, 1986).

The hippocampus has many interconnections to another limbic

structure, the amygdala (Krettek and Price, 1978). The amygdaloid complex

has been shown to be involved in the regulation of fear-motivated behavior

(Blanchard and Blanchard, 1972; Spevack et al., 1975). The amygdala

appears to play a role in the stress response via the pituitary-adrenal

cortical axis . The central nucleus of the amygdala, in particular,

contains large numbers of CRF-containing cell bodies and terminals

(Swanson et al., 1983; Moga and Gray, 1985). Both the central and medial

amygdaloid nuclei project to the thalamic nucleus reuniens (Aggleton and

Mishkin, 1984). This helps to explain why lesions in the nucleus reuniens

prevented the effect of stress on corticosterone secretion.

Most amygdaloid nuclei project to the bed nucleus of the stria

terminalis (BNST) and the hypothalamus. Since the hypothalamus also

receives a substantial projection from the BNST, there exists a di­

synaptic route for the amygdala to modulate hypothalamic function (Krettek

and Price, 1978). The BNST also receives fibers from the dorsal and

median raphe nucleus (Azmitia and Segal, 1978), as well as a massive

projection from the hippocampus (Swanson and Cowan, 1977). To summarize,

142

the incoming sensory information and "learning"/associative processes that

accompany the conditioned fear paradigm are likely conveyed to the PVN via

limbic regions of the telencephalon since direct neocortical projections

to the hypothalamus cannot be demonstrated (Swanson et al., 1987). The

hippocampus, amygdala and ventral septal region project heavily to the

BNST, which in turn projects to the CRF regions of the PVN. The question

that remains is: how does this input to the PVN become integrated with the

direct nordrenergic (A2) and adrenergic (Cl-C3) input via the VNAB (brain

stem to the CRF neurons in the PVN)?

Swanson et al. (1983) suggested that the nucleus of the solitary

tract, which receives direct vagal and glossopharyngeal afferents

(Beckstead and Norgren, 1979) could serve as a major relay point of

visceral sensory information to CRF cell bodies in the basal forebrain.

Likewise the parabrachial nucleus, which receives a large input from the

NTS (Norgren, 1978), and the locus coeruleus also project to many of the

sites that contain CRF-immunoreactive cell bodies, including the

parvocellular PVN, the bed nucleus of the stria terminalis and the central

nucleus of the amygdala. As mentioned above, these nuclei may also be

involved in the relay of visceral sensory information to CRF cell bodies

(Swanson et al., 1983). This hypothetical pathway could provide a direct

stimulus for corticosterone secretion.

Another possibility is that the noradrenergic input to the PVN is

serving as a neuromodulator to "set a bias" to amplify or buffer incoming

stimuli to the cells. Even though norepinephrine might have little effect

on the resting excitability of the CRF neurons, it could have profound

143

effects on the amount of CRF released following an action potential

(Nicoll, 1982). For example, pretreatment with norepinephrine (5 µ.M)

before iontophoretically-applied glutamate (50 nA) in a pyramidal

(hippocampal) cell has been shown to evoke a higher frequency and longer

duration train of action potentials than following glutamate alone

(Madison and Nicoll, 1986).

Wallace et al. (1989) demonstrated that the amygdala innervates the

locus coeruleus (A6) and NTS (A2) noradrenergic cells, and the C2

adrenergic cells. Most of the rostral A6 and A2 neurons had amygdaloid

terminal contacts. This pathway could be involved in integrating the

sensory/associative processes with the stimulatory catecholaminergic

afferents to the PVN.

The possibility exists that serotonergic neurons play a role in

transmitting the stress stimulus. The dorsal raphe nucleus (B7) is one

of the nine 5-HT cell groups originally described by Dahlstrom and Fuxe

(1965), which innervates the forebrain. Studies from our laboratory have

shown that electrolytic lesions in the dorsal raphe nucleus prevent the

stress-induced increase in plasma renin activity using a three-minute CER

(Van de Kar et al., 1984). This experiment was repeated, extending the

CER to ten minutes so that corticosterone levels could be evaluated.

Confirming the previous results, we found that electrolytic lesions in the

dorsal raphe nucleus prevented the stress - induced increase in plasma renin

activity and concentration and corticosterone levels (Richardson Morton

et al., 1986). There could be a connection between the fact that both

dorsal raphe lesions and PVN lesions prevent the stress-induced increase

in plasma renin activity, plasma renin concentration and corticosterone

levels.

144

Sawchenko et al. (1983) have demonstrated that serotonergic

neurons in the midbrain (especially the dorsal raphe) innervate the PVN.

Although the serotonergic fiber density in the PVN is low compared with

immediately surrounding areas, there is a distinct projection to the

medial parts of the parvocellular division. Liposits et al. (1987) have

demonstrated by electron microscopic methods that serotonergic nerve

terminals innervate CRF-immunoreactive parvocellular cells in the PVN.

Injections (i.p.) of the 5-HT2 antagonist, LY53857, did not prevent

the stress-induced increase in renin or corticosterone secretion

(Richardson Morton et al., 1986). However, injection of the 5-HT1A

agonists buspirone and ipsapirone significantly attenuated the stress­

induced increase in plasma renin activity, plasma renin concentration and

corticosterone levels (Urban et al., 1986; Van de Kar et al., 1985; Lorens

et al. , 1989). Both buspirone and ipsapirone inhibit the firing of

serotonergic neurons in the dorsal raphe (Basse-Tomusk, 1986; VanderMaelen

et al., 1986). Thus, serotonergic neurons in the dorsal raphe nucleus

could be involved in stress-induced renin and corticosterone secretion.

There are fiber projections from the locus coeruleus to the dorsal raphe

nucleus (Cowan and Park, 1986), with direct noradrenergic nerve terminal

innervation of serotonin nerve cells in the dorsal raphe (Baraban and

Aghajanian, 1981), and projections from the dorsal raphe to the locus

coeruleus (Conrad et al., 1974). These interconnections could provide a

relay for the noradrenergic bundles to influence hypothalamic neuroen­

docrine responses.

Another possibility is that the dorsal raphe serotonergic neurons,

which also project to the hippocampus (Swanson et al., 1987), provide a

145

tonic inhibition of the "behavioral inhibition" system of the hippocampus.

In this case, administration of 5-HT1A anxiolytic agents or electrolytic

lesions in the dorsal raphe nucleus would decrease the input to the septo­

hippocampal system, thus reducing the activity of the septo-hippocampal

behavioral system, ultimately decreasing the behavioral and possibly

neuroendocrine stress response (Kuhar, 1986).

Clearly, the neural pathways mediating stress-induced renin and

corticosterone secretion are complex and multi-synaptic. From this

dissertation, it can be concluded that stress-induced corticosterone

secretion appears to be mediated by Preceptors in the PVN, while stress­

induced renin secretion is mediated by a receptor that is not P-adrener­

gic. Stress- induced renin secretion may be mediated by a different

catecholamine receptor, a neuropeptide or other substance(s) colocalized

in catecholaminergic nerve terminals.

The following figure is a schematic diagram of the hypothetical

brain pathways mediating stress-induced renin and corticosterone

secretion.

FIGURE 24

Hypothetical Brain Pathways Mediating Stress-Induced Renin and Corticosterone Secretion

~ AG~

146

LITERATURE CITED

Aggleton, J. P. and Mishkin, M.: Projections of the amygdala to the thalamus in the cynomolgus monkey. J. Comp. Neurol. 222: 56-68, 1984.

Aghajanian, G. K., Kuhar, M. J., and Roth, R.H.: Serotonin-containing neuronal perikarya and terminals: Differential effects of p-chlorophenyl­alanine. Brain Research 54: 85-101, 1973.

Allen, J. P. and Allen, C. F.: Role of the amygdaloid complexes in the stress-induced release of ACTH in the rat. Neuroendocrinology 15: 220-230, 1974.

Allen, J. P. and Allen, C. F.: Amygdalar participation in tonic ACTH secretion in the rat. Neuroendocrinology 19: 115-125, 1975.

Allen, L. G., Kalra, P. S., Crowley, W. R., and Kalra, S. P.: Comparison of the effects of neuropeptide Y and adrenergic transmitters on LH release and food intake in male rats. Life Sciences 37: 617-623, 1985.

Alonso, G., Szafarczyk, A., Balmefrezol, M., and Assenmacher, I.: Immunocytochemical evidence for stimulatory control by the ventral noradrenergic bundle of parvocellular neurons of the paraventricular nucleus secreting corticotropin releasing hormone and vasopressin in rats. Brain Research 397: 297-307, 1986.

Alper, R. H. and Ganong, W. F.: Pharmacological evidence that the sympathetic nervous system mediates the increase in secretion of renin produced by p-chloroamphetamine. Neuropharmacology 11: 1237-1240, 1984.

Anderson, D. E. and Brady, J. V.: Preavoidance blood pressure elevations accompanied by heart rate decreases in the dog. Science 172: 595-597, 1971.

Anderson, D. E. and Brady, J. V.: Differential preparatory cardiovas­cular responses to amine and operative behavioral conditioning. Condit. Refl. 7: 82-96, 1972.

Anderson, D. E. and Brady, J. V.: Cardiovascular responses to avoidance conditioning in the dog: effects of beta adrenergic blockade. Psychosom. Med. 38: 181-189, 1976.

Anderson, D. E., Yingling, J. E., and Brady, J. V.: Cardiovascular responses to avoidance conditioning in the dog: effect of alpha adrenergic blockade. Pav. J. Biol. Sci. 11: 150-161, 1976.

Armario, A., Montero, J. L., and Balasch, J.: Sensitivity of corticosterone and some metabolic variables to graded levels of low intensity stresses in adult male rats. Physiology and Behavior 37: 559-561, 1986.

147

148

Armario, A., Restrepo, C., Castellanos, J. M., and Balasch, J.: Dissociation between Adrenocorticotropin and Corticosterone Responses to Restraint after previous chronic exposure to stress. Life Sciences 36: 2085-2092, 1985.

Ayers, C. R.: Plasma renin activity and renin-substrate concentration in patients with liver disease. Gire. Res. 20: 594-598, 1967.

Azmitia, E. C. The serotonin-producing neurons in the midbrain median and dorsal raphe nuclei. In L. L. Iversen, S. D. Iversen and S. H. Snyder (Eds.), Handbook of Psychopharmacology: Chemical Pathways in the Brain. New York: Plenum Press, 1978.

Azmitia, E. C. and Segal, M.: An autoradiographic analysis of differential ascending projections of the dorsal and median raphe nuclei in the rat. J. Comp. Neurol. 179: 641-668, 1978.

Baertschi, A. J., Gahwiler, B., Antoni, F., Holmes, M. C., and Makara, G. B.: No role of vasopressin in stress-induced ACTH secretion? Nature 308: 85-86, 1984.

Baraban, J. M. and Aghajanian, G. K.: Noradrenergic innervation of serotonergic neurons in the dorsal raphe: Demonstration by electron microscopic autoradiography. Brain Research 204: 1-11, 1981.

Bartter, F. C., Casper, A. C., Delea, C. S., and Slater, J. D. H.: On the role of the kidney in control of adrenal steroid products. Metabolism 10: 1006-1020, 1961.

Basse-Tomusk, A. and Rebec, G. V.: Ipsapirone depresses neuronal activity in the dorsal raphe nucleus and the hippocampal formation. Europ. J. Pharmacol. 130: 141-143, 1986.

Basso, N., Gripson, D., Ruiz, P., and Taquini, A. C.: Regional distributions of angiotensin in the central nervous system of the rat: effect of DOCA-salt treatment. Clin. Sci. 63: 1495-1525, 1982.

Bayliss, P. H. and Robertson, G. L.: Rat vasopressin response to insulin-induced hypoglycemia. Endocrinology 107: 1975-1979, 1980.

Beaty, 0., Sloop, C. H., Schmid, H. E., and Buckalew, V. M.: Renin response and angiotensinogen control during graded hemorrhage and shock in the dog. Am. J. Physiol. 231: 1300-1307, 1976.

Beaulieu, S., Di Paolo, T., and Barden, N.: Control of ACTH secretion by the central nucleus of the amygdala: implication of the serotoninergic system and its relevance to the glucocorticoid delayed negative feedback mechanism. Neuroendocrinology 44: 247-254, 1986.

149

Beaulieu, S., Di Paolo, T., Cote, J., and Barden, N.: Participation of the central amygdaloid nucleus in the response of adrenocorticotropin secretion to immobilization stress: Opposing roles of the noradrenergic and dopaminergic systems. Neuroendocrinology 45: 37-46, 1987.

Beckstead, R. M. and Norgren, R.: An autoradiographic examination of the central distribution of the trigeminal, facial, glossopharyngeal, and vagus nerve in the monkey. J. Comp. Neurol. 184: 455-472, 1979.

Bellin, S. I., Bhatnagar, R. K., and Johnson, A. K.: Periventricular noradrenergic systems are critical for angiotensin-induced drinking and blood pressure responses. Brain Research 403: 105-112, 1987.

Benetos, A., Gavras, I., and Gavras, H.: Norepinephrine applied in the paraventricular hypothalamic nucleus stimulates vasopressin release. Brain Res. 381: 322-326, 1986.

Beny, J. L. and Baertschi, A. J.: Oxytocin: major tropin-releasing factor secreted from diabetes insipidus rat pituitary in vitro. Neuroendocrinology 31: 261-264, 1980.

cortico­posterior

Berkenbosch, F., Vermes, I., Buijs, R. M., and Tilders, F. J. H.: Vasopressin is not involved in the catecholamine-induced release of ACTH, alpha-MSH and beta-endorphin from the rat pituitary gland. Neuroendocrinology 37: 117-121, 1983.

Bertand, M. The Behavioral Repertoire of the Stump Tail Macque. New York: Karger, 1969.

Bing, J. and Poulsen, K.: Experimentally-induced changes in plasma angiotensinogen and plasma renin. Acta Pathol. Microbiol. Scand. 77: 389-398, 1969.

Biron, P., Koiw, E., Nowaczynski, W., Broulet, J., and Genest, J.: The effects of intravenous infusion of valine-5-angiotensin II and other pressor agents on urinary electrolytes and corticosteroids including aldosterone. J. Clin. Invest. 40: 338-347, 1961.

Bjorklund, A., Falck, B., and Rosengren, E.: Monoarnines in the pituitary gland of the pig. Life Sciences 6: 2103-2110, 1967.

Blair, M. L. , Chen, Y. H. , and Izzo, J. L. : Influence of renal perfusion pressure on alpha- and beta-adrenergic stimulation of renin release. Arn. J. Physiol. 248: E317-E326, 1985.

Blair, M. L., Feigl, E. 0., and Smith, 0. A.: Elevation of plasma renin activity during avoidance performance in baboons. Arn. J. Physiol. 231: 772-776, 1976.

Blair, M. L., Reid, I. A., and Ganong, W. F.: Effect of L-DOPA on plasma renin activity with and without inhibition of extracerebral DOPA decarboxylase in dogs. J. Pharmacol. Exp. Ther. 202: 209-215, 1977.

150

Blair-West, J. R., Reid, I. A., and Ganong, W. F.: Stimulation of angiotensinogen release by raised blood angiotensin concentration in the dog. Clin. Sci. Mol. Med. 46: 665-669, 1974.

Blanchard, D. C. and Blanchard, R. J.: Innate and conditioned reactions to threat in rats with amygdaloid lesions. J. Comp. Physiol. Psychol. 81: 281-290, 1972.

Bliss, E. L., Ailion, J., and Zwanziger, J.: Metabolism of norepineph­rine, serotonin and dopamine in rat brain with stress. J. Pharmacol. Exp. Ther. 168: 258-263, 1968.

Bloom, F. E., Battenberg, E., Rossier, J., Ling, N., andGuillemin, R.: Neurons containing beta-endorphin exist separately from those containing enkephalin: Immunocytochemical studies. Proc. Natl. Acad. Sci. (USA) 75: 1591-1595, 1978.

Blumberg, A. L., Ackerly, J. A., and Peach, M. J.: Differentiation of neurogenic and myocardial angiotensin II receptors in isolated rabbit atria. Circ. Res. 36: 719-726, 1975.

Bodian, D. and Maren, T. H.: Effect of neuro- and adenohypophysectomy on retrograde degeneration in the hypothalamic nuclei of the rat. J. Comp. Neurol. 94: 485-511, 1951.

Bonjour, J. P. and Malvin, R. L.: Stimulation of ADH release by the renin-angiotensin system. Am.J. Physiol. 218: 1555-1559, 1970.

Breese, G. R. and Traylor, T. D.: Effect of 6-hydroxydopamine on brain norepinephrine and dopamine: Evidence for selective degeneration of catecholamine neurons. Journal of Pharmacology and Experimental Therapeut­ics 174(3): 413-420, 1970.

Britton, D. R., Koob, G. F., Rivier, J., and Vale, W.: Intraventricular corticotropin-releasing factor enhances behavioral effects of novelty. Life Sciences 31: 363, 1982.

Britton, S. W. and Silvette, H.: Some effects of cortico-adrenal extract and other substances on adrenalectomized animals. Am. J. Physiol. 99: 15-32' 1931.

Brod, J. Haemodynamics and emotional stress. In M. Koster,H. Musaph and P. Visser (Eds.), Psychosomatics in Essential Hypertension. Bihl. Psychiat., 1970.

Brod, J., Fencl, V., Hejl, Z., and Jirka, J.: Circulatory changes underlying blood pressure elevation during acute emotional stress (mental arithmetic) in normotensive and hypertensive subjects. Clin. Sci. 18: 269-279, 1959.

151

Brod, J., Fencl, V., Hejl, Z., Jirka, J., and Ulrych, M.: General and regional haemodynamic patterns underlying essential hypertension. Clin. Sci. 23: 339-349, 1962.

Brody, M. J., Alper, R.H., O'Neill, T. P., and Porter, J.P. 1. Central neural control of the cardiovascular system. In A. Zanchetti and R. C. Tarazi (Eds.), Handbook of hypertension: Pathophysiology of hypertension- regulatory mechanisms. : Elsevier Science Publishers, 1986.

Brody, M. J. and Johnson, A. K. Role of the anteroventral third ventricle region in fluid and electrolyte balance, arterial pressure regulation, and hypertension. In L. Martini and W. F. Ganong (Eds.), Frontiers in Neuroendocrinology. New York: Raven Press, 1980.

Brown, M. R., Fischer, L.A., Spiess, J., Rivier, C., Rivier, J., and Vale, W.: Corticotropin-releasing factor: actions on sympathetic nervous system and metabolism. Endocrinology 111: 928, 1982.

Brown, R. J. Mammalian social odors: A critical review. In J. S. Rosenblatt,R. A. Hinde, L. Beer and M. Busnel (Eds.), Advances in the study of behavior. New York: Academic Press, 1979.

Brown-Sequard, C. E.: Recherches experimentales sur la physiologie et la pathologie des capsules surrenales. C.R. Acad. Sci. (Paris) 43: 422-425, 1856.

Brownfield, M. S., Reid, I. A., Ganten, D., and Ganong, W. F.: Differential Distribution of Immunoreactive Angiotensin and Angioten­sin-Converting Enzyme in Rat Brain. Neuroscience 7: 1759-1769, 1982.

Bruni, J. F., Hawkins, R. L., and Yen, S. S. C.: Serotonergic mechanism in the control of beta-endorphin and ACTH release in male rats. Life Sciences 30: 1247-1254, 1982.

Buckingham, J.C. and Hodges, J. R.: Hypothalamic receptors influencing the secretion of corticotropin-releasing hormone in the rat. J. Physiol. (London) 290: 421-431, 1979.

Bugnon, C., Hadji, Y., Iassemis, M., Fellman, D., and Cardot, J.: Reserpine-induced depletion of corticoliberin (CRF)-like immunoreactivity in the zona externa of the rat median eminence. Brain Research 275: 198-201, 1983.

Bunag, R. D. and Inoue, A. : Differentiated pressor and sympathetic response to dual brain stimulation: Ventromedial hypothalamus versus locus coeruleus. Proc. Soc. Exp. Biol. Med. 178: 90-100, 1985.

Caldwell, P. R. B., Seegal, B. C., and Hsu, K. C.: Angioten­sin-Converting Enzyme: Vascular Endothelial Localization. Science 191: 1050-1051 .• 1976.

152

Campbell, D. J., Skinner, S. L., and Day, A. J.: Cellophane perinephr­itis hypertension and its reversal in rabbits. Circ. Res. 33: 105-112, 1973.

Cannon, W. B.: The emergency function of the adrenal medulla in pain and the major emotions. Am. J. Physiol. 33: 356-372, 1914.

Carlsson, A., Falck, B., and Hillarp, N.: Cellular localization of brain monoamines. Acta Physiol. Scand. 56: 1-27, 1962.

Carretero, 0. and Gross, F.: Renin substrate in plasma under various experimental conditions in the rat. Am J Physiol 213(3): 695-700, 1967.

Castren, E. and Saavedra, J.M.: Repeated stress increases the density of angiotensin II binding sites in rat paraventricular nucleus and subfornical organ. Endocrinology 122(1): 370-372, 1988.

Changaris, D., Demers, L., Keil, L., and Severs, W. Immunopharmacology of angiotensin I in brain. In J. Buckley and C. Ferrario (Eds.), Central Actions of Angiotensin and Related Hormones. New York: Pergamon Press, 1977.

Changaris, D. G., Keil, L., and Severs, W. B.: Angiotensin II immunohis­tochemistry of the rat brain. Neuroendocrinology 25: 257-274, 1978.

Charney, D. S. and Redmond, D. E.: Neurobiological mechanisms in human anxiety--Evidence supporting central noradrenergic hyperactivity. Neuropharmacology 22: 1531-1536, 1983.

Churchill, P. C., Churchill, M. C., and McDonald, F. D.: Renin secretion and distal tubule Na+ in rats. Am. J. Physiol. 235(6): F611-F616, 1978.

Cintra, A., Fuxe, K., Harfstrand, A., Agnati, L. F., Wikstrom, A. C., Okret, S., Vale, W., and Gustafsson, J. A.: Presence of glucocorticoid receptor immunoreactivity in corticotropin releasing factor and in growth hormone releasing factor immunoreactive neurons of the rat di- and telencephalon. Neuroscience Letters 77: 25-30, 1987.

Ciriello, J. and Calaresu, F.: Lateral reticular nucleus: a site of somatic and cardiovascular integration in the cat. Am. J. Physiol. 233: RlOO, 1977.

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 Research 310: 355-359, 1984.

Clamage, D. M., Sanford, C. S., Vander, A. J., and Mouw, D.R.: Effect of psychosocial stimuli on plasma renin activity in rats. Am J Physiol 231: 1290-1294, 1976.

Cohen, D. H. and Randall, D. C.: Classical conditioning of cardiovas­cular responses. Ann. Rev. Physiol. 46: 187-197, 1984.

153

Conrad, L. C. A., Leonard, C. M., and Pfaff, D. W.: Connections of the median and dorsal raphe nuclei in the rat: An autoradiographic and degeneration study. J. Comp. Neurol. 156: 179-206, 1974.

Conway, J., Greenwood, D. T., and Middlemiss, D. N.: Central nervous actions of beta-adrenoceptor antagonists. Clin. Sci. Mol. Med. 54: 119-124, 1978.

Cori, C. F. and Cori, G. T.: The fate of sugar in the animal body. VII. The carbohydrate metabolism of adrenalectomized rats and mice. J. Biol. Chem. 74: 473-494, 1927.

Correa, F. M., Innis, R. B., Uhl, G. R., and Snyder, S. H.: Bradykinin-­like immunoreactive neuronal systems localized histochemically in rat brain. Proc. Natl. Acad. Sci. (USA) 76: 1489-1493, 1979.

Corrodi, H., Fuxe, K., and Hokfelt, T.: The effect of immobilization stress on the activity of central monoamine neurons. Life Sciences 7: 107-112, 1968.

Cowan, R. L. and Park, M. L.: PHAL Mapping of the Locus Coeruleus to Dorsal Raphe Nucleus Projection in the Rat. Neuroscience Abstracts 12: 139' 1986.

Cowan, W. M., Raisman, G., and Powell, R. P. S.: The connections of the amygdala. J. Neurol. Neurosurg. Psychiat. 28: 137-151, 1965.

Cuello, A. C., Scapagnini, U., Licko, V., Preziosi, P., and Ganong, W. F. : Effect of dihydroxyphenylserine on the increase in plasma cor­ticosterone in rats treated with alpha-methyl-para-tyrosine. Neuroen­docrinology 13: 115-122, 1973.

Cunningham, E. T. and Sawchenko, P. E.: Anatomical specificity of noradrenergic inputs to the paraventricular and supraoptic nuclei of the rat hypothalamus. The Journal of Comparative Neurology 274: 60-76, 1988.

Dabsys, S. M., Balda, M. S., Pirola, C. J., Finkielman, S., and Nahmod, V. E.: Angiotensin converting enzyme activity in the amygdaloid complex in a neurogenic hypertensive model. Clin. and Exper. Hyper. Al0(4): 605-615, 1988.

Dahlstrom, A. and Fuxe, K.: Evidence for the existence of monoamine-con­taining neurons in the central nervous system IV. Acta Physiol. Scand. 64(247): 1-36, 1965.

Darlington, D. N., Shinsako, J., and Dallman, M. F.: Paraventricular lesions: hormonal and cardiovascular responses to hemorrhage. Brain Research 439: 289-301, 1988.

Davis, J. O. and Freeman, R.H.: The use of angiotensin II blockade to study adrenal steroid secretion. Fed. Proc. 35: 2508-2511, 1976.

154

De Kloet, E. R., Sybesma, H., and Reul, M. H. M.: Selective control by corticosterone of serotonin-1 receptor capacity in raphe-hippocampal system. Neuroendocrinology 42: 513-521, 1986.

de Molina, A. F. and Hunsperger, R. W.: Organization of the subcortical system governing defence and flight reactions in the cat. J. Physiol. 160: 200-213, 1962.

De Vito, E., Koninckx, E., Cabrerea, R., and Nolly, H.: Half-life of circulating renin under different experimental conditions. Mayo Clin. Proc. 52: 424-426, 1977.

Decavel, C., Geffard, M., and Calas, A.: Comparative study of dopamine­and noradrenaline-immunoreactive terminals in the paraventricular and supraoptic nuclei of the rat. Neuroscience Letters 77: 149-154, 1987.

DiBona, G. F.: Neural regulation of renal tubular sodium reabsorption and renin secretion. Fed. Proc. 44: 2816-2822, 1985.

Dierickx, K. and somatostatin neurons 349-359, 1979.

Vandesande, F.: Immunochemical in the rat hypothalamus. Cell

localization of Tiss. Res. 201:

Dohanics, J., Kapocs, G., Janaky, T., Kiss, J. Z., Rappay, G., Laszlo, F. A., and Makara, G. B.: Mechanism of restoration of ACTH release in rats with long-term lesions of the paraventricular nuclei. J. Endocr. 111: 75-82, 1986.

Douglas, R. J.: The hippocampus and behavior. Psychological Bulletin 67: 416-442, 1967.

Dua, J. K. and Dobson, M. J.: Role of olfactory cues in acquisition and extinction of avoidance. Journal of Experimental Psychology 103: 461-465, 1974.

Dunn, A. J. and Berridge, C. W.: Corticotropin-releasing factor administration elicits a stress-like activation of cerebral catecholamin­ergic systems. Pharmacol. Biochem. Behav. 27(4): 685-691, 1987.

Dykman, R. A. and Gantt, W. H.: Relation of experimental tachycardia to amplitude of motor activity and intensity of the motivating stimulus. Am. J. Physiol. 185: 495-498, 1956.

Dzau, V. J. and Pratt, R. R. Renin-Angiotensin System: Biology, Physiology and Pharmacology. In H. A. Fozzard (Ed.), The Heart and Cardiovascular System. New York: Raven Press, 1986.

Eckland, D. J. A., Todd, K., Jessop, D. S., Biswas, S., and Lightman, S. L.: Differential effects of hypothalamic catecholamine depletion on the release of arginine vasopressin and CRF-41 into hypothalamo­hypophyseal portal blood. Neuroscience Letters 90: 292-296, 1988.

155

Eggena, P. and Barrett, J. D.: Renin substrate release in response to perturbations of renin-angiotensin system. Am. J. Physiol. 254: E389-E393, 1988.

Eisenberg, R. M.: Further evidence of a central alpha-adrenergic inhibitory influence on the hypothalamo-pituitary-adrenal axis in the rat. Neuroendocrinology 17: 154-166, 1975.

Eliasson, K.: Stress and Catecholamines. Acta Med. Scand. 215: 197-204, 1984.

Epstein, S. and Hamilton, S.: Cyproheptadine inhibition of stimulated plasma renin activity. J. Clin. Endocrinol. Metab. 45: 1235-1237, 1977.

Fahri, E. R., Cant, J. R., and Barger, A. C.: Interactions between intrarenal epinephrine receptors and the renal baroreceptor in the control of PRA in conscious dogs. Circ. Res. 50: 477-485, 1982.

Fanselow, M. S.: Odors released by stressed rats produce opioid analgesia in unstressed rats. Behavioral Neuroscience 99(3): 589-592, 1985.

Feldman, S. and Conforti, N.: Amygdalectomy inhibits adrenocortical responses to somatosensory and olfactory stimulation. Neuroendocrinology 32: 330-334, 1981.

Feldman, S., Conforti, N., and Chowers, I.: Adrenocortical responses following sciatic nerve stimulation in rats with partial hypothalamic deafferentations. Acta endocr., Copenh. 80: 625-629, 1975.

Feldman, S., Conforti, N., and Melamed, E.: Norepinephrine depletion in the paraventricular nucleus inhibits the adrenocortical responses to neural stimuli. Neuroscience Letters 64: 191-195, 1986.

Feldman, S., Conforti, N., and Melamed, E.: Paraventricular nucleus serotonin mediates neurally stimulated adrenocortical secretion. Brain Res. Bull. 18: 165-168, 1987.

Feldman, S., Conforti, N., and Melamed, E.: Involvement of ventral noradrenergic bundle in corticosterone secretion following neural stimuli. Neuropharmacology 27(2): 129-133, 1988.

Fety, R., Lambas-Senas, L., Chamba, G., and Renaud, B.: Changes in tyrosine hydroxylase and dopamine-beta-hydroxylase activities but not in phenylethanolamine-N-methyltransferase activity within central adrenaline neurons after 6-hydroxydopamine administration. Biochemical Pharmacology 33(12): 1887-1891, 1984.

Finley, J. C. W., Maderdrut, J. L., and Petrusz, P.: The immuno­cytochemical localization of enkephalin in the central nervous system of the rat. J. Comp. Neurol. 198: 541-565, 1981.

156

Fischer-Ferraro, C., Nahmod, V. E., Goldstein, D. J., and Finkielman, S.: Angiotensin and renin in rat and dog brain. J. Exp. Med. 133: 353-361, 1971.

Fitzgerald, J. D. Beta-adrenoceptor antagonists. In P. A. Van Zwieten (Ed.), Handbook of Hypertension. Pharmacology of antihypertensive drugs. Amsterdam: Elsevier Science Publishers, 1984.

Fitzsimons, C. L. and Ciriello, J.: Selective Destruction of Parvocel­lular Neurons in the Paraventricular Nucleus Decreases Norepinephrine Turnover Rates in Kidney of Spontaneously Hypertensive Rats. Fed. Proc. 45: 875, 1986.

Fitzsimons, J. T.: The effect on drinking of peptide precursors and of shorter chain peptide fragments of angiotensin II injected into the rat's diencephalon. J. Physiol. 214: 295-303, 1971.

Freeman, R.H. and Rostorfer, H. H.: Hepatic changes in renin substrate biosynthesis and alkaline phosphatase activity in the rat. Am. J. Physiol. 223: 364-370, 1972.

Fujii, A. M. and Vatner, S. F.: Direct versus indirect pressor and vasoconstrictor actions of angiotensin in conscious dogs. Hypertension 7: 253-261, 1985.

Fuller, R. W.: Serotonergic stimulation of pituitary-adrenocortical function in rats. Neuroendocrinology 32: 118-127, 1981.

Fuller, R. W. and Snoddy, H. D.: Elevation of plasma corticosterone by swim stress and insulin-induced hypoglycemia in control and fluoxetine-­pretreated rats. Endocrine Research Communications 4(1): 11-23, 1977.

Fuller, R. W. and Snoddy, H. D.: Effect of serotonin-releasing drugs on serum corticosterone concentration in rats. Neuroendocrinology 31: 96-100, 1980.

Fuller, R. W. and Snoddy, H. D.: Elevation of serum corticosterone concentrations in rats by pergolide and other dopamine agonists. Endocrinology 109: 1026-1032, 1981.

Furness, J. B., Heath, J. W., and Costa, M.: Aqueous aldehyde (Faglu) methods for the fluorescence histochemical localization of catecholamines and for ultrastructural studies of central nervous tissue. Histochemistry 57: 285-295, 1978.

Fuxe, K. , Agnati, L. MerloPich, E., Hokfelt, 189-192, 1983.

F.' T.'

Harfstrand, A., Zini, I., Tatemoto, and Mutt, V.: Acta Physiol. Scand.

K.' 118:

157

Fuxe, K., Ganten, D., Hokfelt, T., and Bohne, P.: Immunohistochemical evidence for the existence of angiotensin II containing nerve terminals in the brain and spinal cord of the rat. Neuroscience Letters 2: 229-234, 1976.

Fuxe, K., Ganten, D., Hokfelt, T., Locatelli, V., Poulsen, K., Stock, G., Rix, E., and Taugner, R.: Renin-like immunocytochemical activity in the rat and mouse brain. Neuroscience Letters 18: 245-250, 1980.

Fuxe, K., Hokfelt, T., Eneroth, P., Gustafsson, J. A., and Skett, P.: Prolactin: Localization in nerve terminals of the rat hypothalamus. Science 196: 899-900, 1977.

Galen, F. X., Devaux, C., Guyenne, T., Menard, J., and Corvol, P.: Multiple Forms of Human Renin- Purification and Characterization. The Journal of Biological Chemistry 254(11): 4848-4855, 1979.

Galosy, R. A., Crawford, I. L., and Thompson, M. E. Behavioral stress and cardiovascular regulation: Neural mechanisms. In Smith, Galosy and Weiss (Eds.), Circulation, Neurobiology, and Behavior. : Elsevier Science Publishing Co., 1982.

Ganong, W. F.: Neurotransmitters and pituitary function: regulation of ACTH secretion. Fed. Proc. 39: 2923-2930, 1980.

Ganong, W. F. and Barbieri, C. Neruoendocrine components in the regulation of renin secretion. In: Frontiers Neuroendocrinology, 1982.

Ganong, W. F., Chalett, J., Jones, H., Kaplan, S. L., Karteszi, M., Stith, R. D., and Van de Kar, L. D.: Further characterization of putative alpha-adrenergic receptors in brain that affect blood pressure and the secretion of ACTH, GH and renin in dogs. Endocrinologia Experimentalis 16: 191-205, 1982.

Ganong, W. F. , Rudo 1 ph, C. conponents in the regulation Hypertension 1: 207-218, 1979.

D., and Zimmermann, H.: Neuroendocrine of blood pressure and renin secretion.

Garvey, H. L. and Ram, N.: Comparative antihypertensive effects and tissue distribution of beta adrenergic blocking drugs. J. Pharmacol. Exp. Ther. 194: 220-233, 1975.

Giarcovich, S. and Enero, M.A.: Decreased brain serotonergic activity after acute propranolol. Eur. J. Pharmacol. 100: 123-125, 1984.

Gibbs, D. M.: Dissociation of oxytocin, vasopressin and corticotropin secretion during different types of stress. Life Sciences 35: 487-491, 1984.

158

Gibbs, D. M. and Vale, W.: Effect of the serotonin reuptake inhibitor fluoxetine on corticotropin-releasing factor and vasopressin secretion into hypophysial portal blood. Brain Research 280: 176-179, 1983.

Gibson, A., Hart, S. L., and Patel, S.: Effects of 6-hydroxydopamine-­induced lesions of the paraventricular nucleus, and of prazosin, on the corticosterone response to restraint in rats. Neuropharmacology 25(3): 257-260, 1986.

Gilbert, F., Brazell, C., Tricklebank, M. D., and Stahl, S. M.: Activation of the 5-HTlA receptor subtype increases rat plasma ACTH concentration. European Journal of Pharmacology 147: 431-439, 1988.

Gillies, G.E., Linton, E.A., Lowry, P.J.: Corticotropin releasing activity of the new CRF is potentiated several times by vasopressin. Nature 299: 355-357, 1982.

Glavin, G. B.: Selective noradrenaline depletion markedly alters stress responses in rats. Life Sciences 37: 461-465, 1985.

Goldblatt, H., Lynch, J., Hanzal, R. F., and Summerville, W.W.: Studies on experimental hypertension. I. The production of persistent elevation of systolic blood pressure by means of renal ischemia. J. Exp. Med. 59: 347-379, 1934.

Goldman, C. K., Marino, L., and Leibowitz, S. F.: Postsynaptic alpha-2 noradrenergic receptors mediate feeding induced by paraventricular nucleus injection of norepinephrine and clonidine. European Journal of Pharmacol­ogy 115: 11-19, 1985.

Golin, R. M. A., Gotoh, E., Said, S. I., and Ganong, W. F.: Phar­macological evidence that the sympathetic nervous system mediates the increase in renin secretion produced by immobilization and head-up tilt in rats. Neuropharmacology 27(12): 1209-1213, 1988.

Goodwin, F. J. , Kirshman, J. D. , Sealey, J. E. , and Laragh, J. H. : Influence of the pituitary gland on sodium conservation, plasma renin and renin substrate concentrations in the rat. Endocrinology 86: 824-834, 1970.

Gordon, D. B. and Sullivan, P. H.: Acute effect of over transfusion on plasma renin activity in the rat. Endocrinology 85: 341-343, 1969.

Gotoh, E., Bahnson, T. D., Alper, R. H., and Ganong, W. F.: Role of the hypothalamus in the regulation of renin secretion in rats. The Physio­logist 28(4): 41.14, 1985.

Gotoh, E., Golin, R. M. A., and Ganong, W. F.: Relation of the ventromedial nuclei of the hypothalamus to the regulation of renin secretion. Neuroendocrinology 47: 518-522, 1988.

159

Gotoh, E., Murakami, K., Bahnson, T. D., and Ganong, W. F.: Role of brain serotonergic pathways and hypothalamus in regulation of renin secretion. Am. J. Physiol. 253: Rl79-Rl85, 1987.

Gray, J. A. The Psychology of Fear and Stress. London: Weidenfeld and Nicolson, 1971.

Gray, J. A. The Neuropsychology of Anxiety. Oxford: Clarendon Press, 1982.

Green, R. H., Beatty, W. S., and Schwartzbaum, J. S.: Comparative effects of septo-hippocampal and caudate lesions on avoidance behavior in rats. Journal of Comparative and Physiological Psychology 64: 444-452, 1967.

Greenwood, F. C., Hunter, W. M., and Glover, J. S.: The preparation of 131I-labelled human growth hormone of high specific radioactivity. Biochem. J. 89: 114-123, 1963.

Gregory, T. J., Wallis, C. J., and Printz, M. P.: Regional changes in rat brain angiotensin following bilateral nephrectomy. Hypertension 4: 827-838, 1982.

Grossman, S. P., Grossman, L., and Walsh, L.: Functional organization of the rat amygdala with respect to avoidance behavior. J. Comp. Physiol. Psychol. 88: 829-850, 1975.

Guillaume, V., Conte-Devolx, B., bute, N. , Grino, M. , Alonso, tropin-releasing factor release mediated by brain catecholamines. 143-146, 1987.

Szafarczyk, A., Malaval, F., Pares-Her­G., and Oliver, C.: The cortico­in rat hypophysial portal blood is Neuroendocrinology 46:

Gustafsson, J. A., Okret, S., Wikstrom, A. C., Anderson, B., Radojcic, M., Wrange, 0., Sachs, W., and Doupe, A. J. On the use of poly- and monoclonal antibodies in studies on the structure and function of the glucocorticoid receptor. In H. Eriksson and J. A. Gustafsson (Eds.), Steroid hormone receptors: structure and function. Amsterdam: Elsevier, 1983.

Gutman, F. D. , Tagawa, H. , Haber, E. , and Barger, A. C. : Renal arterial pressure, renin secretion and blood pressure control in trained dogs. Amer. J. Physiol. 224: 66-72, 1973.

Guyton, A. C. The autonomic nervous system: the adrenal medulla. In (Ed.), Medical Physiology. Philadelphia: W.B. Saunders Co., 1981.

Haber, C., Loerner, D., Page, L. B., Kliman, B., and Purnode, A.: Application of a radioimmunoassay for angiotensin I to the physiologic measurements of plasma renin activity in normal human subjects. Journal of Clinical Endocrinology 29: 1329-1355, 1969.

160

Hall, C. S.: Temperament: a survey of animal studies. Psycho!. Bull. 38: 909-943, 1941.

Harfstrand, A., Eneroth, P., Agnati, L., and Fuxe, K.: Further studies on the effects of central administration of neuropeptide Y on neuroen­docrine function in the male rat: relationship to hypothalamic catechol­amines. Regulatory Peptides 17: 167-179, 1987.

Harfstrand, A., Fuxe, K., Agnati, L. F., Eneroth, P., Zini, I., Zoli, M., Andersson, K., and von Euler, G.: Studies on neuropeptide Y-catechol­amine interactions in the hypothalamus and in the forebrain of the male rat. Relationship to neuroendocrine function. Neurochem. Int. 8(3): 355-376, 1986.

Harfstrand, A., Fuxe, K., Cintra, A., Agnati, L. F., Zini, I., 'Wikstrom, A. C., Okret, S., and Yu, Z. Y.: Glucocorticoid receptor immunoreactivity in monoaminergic neurons of rat brain. Proc. Natl. Acad. Sci. (USA) 83: 9779-9783, 1986.

Harris, G. 'W.: Neural Control of the Pituitary Gland. Physiol. Rev. 28: 139, 1948.

Harrop, G. A., Soffer, L. J., Ellsworth, R., and Trescher, J. H.: Studies on the suprarenal cortex. III. Plasma electrolytes and electrolyte excretion during suprarenal insufficiency in the dog. J. Exp. Med. 58: 17-38, 1933.

Hatton, G. I., Hutton, U. E., Hoblitzell, E. R., and Armstrong, 'W. E.: Morphological evidence for two populations of magnocellular elements in the rat paraventricular nucleus. Brain Research 108: 187-193, 1976.

Hauger, R. L., Aguilera, G., Baukal, A., and Catt, K. J.: Characteriza­tion of angiotensin II receptors in the anterior pituitary gland. Mol. Cell. Endocrinol. 25: 203-212, 1982.

Healy, D. P., Munzel, P. A., and Insel, P. A.: Localization of the beta-1 and beta-2-adrenergic receptors in rat kidney by autoradiography. Gire. Res. 57: 278-284, 1985.

Herman, J. P. and 'Wiegand, S. J.: Ibotenate-induced cell death in the hypothalamic paraventricular nucleus: differential susceptibility of magnocellular and parvicellular neurons. Brain Res. 383: 367-372, 1986.

Hetherington, A. 'W. and Ranson, S. 'W.: The spontaneous activity and food intake of rats with hypothalamic lesions. Am. J. Physiol. 136: 609-617, 1942.

Hiwada, K., Tanaka, H., and Kokubu, T.: The influence of nephrectomy, uretal ligation, and of estradiol on plasma renin substrate in unilateral­ly nephrectomized rats. Pflugers Arch. 365: 177-182, 1976.

161

Hokfelt, T., Fuxe, K., and Goldstein, M.: Immunohistochemical studies on monoamine-containing cell systems. Brain Research 62: 461-469, 1973.

Hokfelt, T., Fuxe, K., Goldstein, M., and Johansson, O.: Immunohis­tochemical evidence for the existence of adrenaline neurons in the rat brain. Brain Research 66: 239-251, 1974.

Hokfelt, T., Fuxe, K., Johansson, 0., Jeffcoate, S., Distribution of thyrotropin-releasing hormone (TRH) in the system as revealed with immunohistochemistry. Eur. J. 389-392, 1975.

and White, N.: central nervous Pharmacol. 34:

Hokfelt, T., Johansson, 0., and Goldstein, M.: Chemical anatomy of the brain. Science 225: 1326-1334, 1984.

Hollenberg, N. K., Williams, G. H., and Adams, D. F.: Essential hypertension: Abnormal renal vascular and endocrine responses to a mild psychologic stimulus. Hypertension 3: 11-17, 1981.

Holmberg, G. and Gershon, S. : Autonomic and psychiatric effects of yohimbine hydrochloride. Psychopharmacologia 2: 93-106, 1961.

Horky, K., Rojo-Ortega, J. M., Rodriguez, J., Boucher, R., and Genest, J.: Renin, renin substrate and angiotenin I - converting enzyme in the lymph of rats. Am. J. Physiol. 220(2): 307-311, 1971.

Hosoya, Y. and Matsushita, M.: Identification and distribution of the spinal and hypophysial projection neurons in the paraventricular nucleus of the rat. A light and electron microscopic study with the horseradish peroxidase method. Exp. Brain Res. 35: 315-331, 1979.

Hunt, H. F. and Otis, L. S.: Conditioned and unconditioned emotional defecation in the rat. J. Comp. Physiol. Psychol. 46: 378-382, 1953.

Husain, M. K., Manger, W. M., Rock, T. W., Weiss, R. J., and Frantz, A. G.: Vasopressin release due to manual restraint in the rat: Role of body compression and comparison with other stressful stimuli. Endocrinology 104: 641-644, 1979.

Iimori, K., Tanaka, M., Kohno, Y., Ida, Y., Nakagawa, R., Hoaki, Y., Tsuda, A., and Nagasaki, N.: Psychological stress enhances noradrenaline turnover in specific brain regions in rats. Pharmacol. Biochem. Behav. 16(4): 637-640, 1982.

Inagami, T. and Murakami, K.: Prorenin. Biomed. Res. 1: 456-475, 1980. Ingelfinger, J. R., Pratt, R. E., Ellison, K., and Dzau, V. J.: Sodium regulation of angiotensinogen mRNA expression in rat kidney cortex and medulla. J. Clin. Invest. 78: 1311-1315, 1986.

162

Ingle, D. J., Higgins, G. M., and Kendall, E. C.: Atrophy of the adrenal cortex in the rat produced by administration of large amounts of cortin. Anat. Rec. 71: 363-372, 1938.

Itoh, S., Carretero, A., and Murray, R. D.: Possible role of adenosine in the macula densa mechanism of renin release in rabbits. J. Clin. Invest. 76: 1412-1417, 1985.

Ixart, G. , Alonso, G. , Szafarczyk, A. , Malaval, F. , Nouguier- Soule, J. , and Assenmacher,!.: Adrenocorticotropic Regulations after Bilateral Lesions of the Paraventricular or Supraoptic Nuclei and in Brattleboro Rats. Neuroendocrinology 35: 270-276, 1982.

Ixart, G., Barbanel, G., Conte-Devoix, B., Grino, M., Oliver, C., and Assenmacher, I.: Evidence for basal and stress-induced release of corticotropin releasing factor in the push-pull cannulated median eminence of conscious free-moving rats. Neuroscience Letters 74: 85-89, 1987.

Jacobowitz, D. M. and O'Donohue, T.: Alpha-melanocyte stimulating hormone: Immunohistochemical identification and mapping in neurons of rat brain. Proc. Natl. Acad. Sci. (USA) 75: 6300-6304, 1978.

James, W. The Principles of Psychology. New York: Henry Holt, 1890.

Jennes, L. and Stumpf, W. E.: LHRH-systems in the brain of the golden hamster. Cell Tiss. Res. 209: 239-256, 1980.

Jezova, D., Jurcovicova, J., Vigas, M., Murgas, K., and Labrie, F.: Increase in plasma ACTH after dopaminergic stimulation in rats. Psycho­pharmacology 85: 201-203, 1985.

Jindra, A., Kvetnansky, R., Belova, T. I., andSudakov, K. V. Effect of acute and repeated immobilization stress on plasma renin activity, catecholamines and corticosteroids in Wistar and August Rats. In E. Usdin and I. Kopin (Eds.), Catecholamines and Stress. North Holland: Elsevier, 1980.

Johnson, J. A., Davis, J. 0., and Witty, R. T.: Effects of catechol­amines and renal nerve stimulation on renin release in the nonfiltering kidney. Circ. Res. 29: 646-653, 1971.

Johnson, M. D.: Circulating epinephrine stimulates renin secretion in anesthetized dogs by activation of extrarenal adrenoceptors. Am. J. Physiol. 246: F676-F684, 1984.

Johnson, M. D., Shier, D. N., and Barger, A. C.: Circulating catechol­amines and control of plasma renin activity in conscious dogs. Am. J. Physiol. 236(3): H463-H470, 1979.

163

Johnston, C. A., Spinedi, E. J., and Negro-Vilar, A.: Effect of acute ether stress on monoamine metabolism in median eminence and discrete hypothalamic nuclei of the rat brain and on anterior pituitary hormone secretion. Neuroendocrinology 41: 83-88, 1985.

Jones, M. T., Millhouse, E. W., and Burden, J.: Effects of various putative neurotransmitters on the secretion of corticotropin-releasing hormone from the rat hypothalamus in vitro -a model of the neurotransmit­ters involved. J. Endocr. 69: 1-10, 1976.

Joseph, S. A.: Immunoreactive adrenocorticotropin in rat brain: A neuroanatomical study using antiserum generated against synthetic ACTH. Am. J. Anat. 158: 533-548, 1980.

Joseph, S. A. and Sternberger, L. A.: The unlabeled antibody method. Contrasting color staining of beta-lipotropin and ACTH-associated hypothalamic peptides without antibody removal. J. Histochem. Cytochem. 27: 1430-1437, 1979.

Kahn, D., Abrams, G. M., Zimmerman, E. A., Carraway, R., and Leeman, S. E.: Neurotensin neurons in the rat hypothalamus: An immunohistochemical study. Endocrinology 107: 47- 54, 1980.

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 Research 329: 205-212, 1985.

Kant, G. J., Mougey, E. H., Pennington, L. L., and Meyerhoff, J. L.: Graded Footshock Stress Elevates Pituitary Cyclic AMP and Plasma B­endorphin, B-LPH, Corticosterone and Prolactin. Life Sciences 33: 2657-2663, 1983.

Kaplan, N. M.: The biosynthesis angiotensin II, adrenocorticotropin, 2029-2039, 1965.

of adrenal steroids: effects of and potassium. J. Clin. Invest. 44:

Karteszi, M., Van de Kar, L. D., Makara, G., Stark, E., and Ganong, W.F.: Evidence that the mediobasal hypothalamus is involved in serotoner­gic stimulation of renin secretion. Neuroendocrinology 34: 323-326, 1982.

Keeton, T. K. and Campbell, W. B.: The pharmacologic alteration of renin release. Pharmacological Reviews 32(2): 81-227, 1980.

Keil, L. C., Summy-Long, J., and Severs, W. B.: Release of vasopressin by angiotensin II. Endocrinology 96: 1063-1065, 1975.

Keller-Wood, M., Kimura, B., Shinsako, J., and Phillips, M. I.: Interaction between CRF and angiotensin II in control of ACTH and adrenal steroids. Am. J. Physiol. 250: R396- R402, 1986.

164

Kety, S. S. The biogenic amines in the central nervous system: Their possible roles in arousal, emotion and learning. In F. 0. Schmitt (Ed.), The Neurosciences Second Study Program. New York: Rockefeller University Press, 1970.

Khayyall, M., MacGregor, J., Brown, J. J., Lever, A. F., and Robertson, J.I.S.: Increase of plasma renin-substrate concentration after infusion of angiotensin in the rat. Clin. Sci. 44: 87-90, 1973.

Kidman, A.: Stress, cognition and the nervous system. Neurochem. Int. 6(6): 715-720, 1984.

Kimura, H., McGeer, P. L., and McGeer, E. G.: The central cholinergic system studied by choline acetyltransferase immunohistochemistry in the cat. J. Comp. Neurol. 200: 151-201, 1981.

Kiss, J. Z., Mezey, E., and Skirboll, L.: Corticotropin-releasing factor -immunoreactive neurons of the paraventricular nucleus become vasopressin positive after adrenalectomy. Proc. Natl. Acad. Sci. (USA) 81: 1854-1858, 1984.

Klett, C. and Heckenthal, E.: and secretion by angiotensin 2027-2047, 1987.

Induction of angiotensinogen synthesis II. Clin. and Exper. Hyper. A9(12):

Kling, A. and Hutt, P. J.: Effect of hypothalamic lesions on the amygdala syndrome in the cat. A.M.A. Arch. Neurol. Psychiat. 79: 511-517, 1958.

Knepel, W., Homolka, L., Vlaskovska, M., and Nutto, D.: Stimulation of adrenocorticotropinjbeta-endorphin release by synthetic ovine cor­ticotropin-releasing factor in vitro. Neuroendocrinology 38: 344-350, 1984.

Knepel, W. and Meyer, D. K.: Role of the renin-angiotensin system in isoprenaline-induced vasopressin release. J. Cardiovasc. Pharmacol. 2: 815-824, 1980.

Knigge, K. M.: Adrenocortical response to stress in rats with lesions in hippocampus and amygdala. Proc. Soc. Exp. Biol. Med. 108: 18-21, 1961.

Kobayashi, R. M., Palkovits, M., Jacobowitz, D. M., and Kopin, I. J.: Biochemical Mapping of the Noradrenergic Projection from the Locus Coeruleus. Neurology 25: 223-233, 1975.

Koenig, J. I. , Gudelsky, G. A. , and Meltzer, H. Y. : Stimulation of corticosterone and beta-endorphin secretion in the rat by selective 5-HT receptor subtype activation. Europ. J. Pharmacol. 137: 1-8, 1987.

Koepke, J. P. and DiBona, G. F.: Central adrenergic receptor control of renal function in conscious hypertensive rats. Hypertension 8: 133-141, 1986.

16S

Kohno, Y., Tanaka, M., Nakagawa, R., Toshima, N., and Nagasaki, N.: Regional distribution and production rate of 3-methoxy-4- hydroxyphenyl­ethyleneglycol sulphate (MHPG-S04) in rat brain. J. Neurochem. 36: 286-289, 1981.

Koizumi, K. and Yamashita, H.: Studies of antidromically identified neurosecretory cells of the hypothalamus by intracellular and ex­tra-cellular recordings. J. Physiol. (London) 221: 683-70S, 1972.

Korf, J., Aghajanian, G. K., and Roth, R.H.: Stimulation and destruc­tion of the locus coeruleus: Opposite effects on 3-methoxy-4-hydroxy­phenylglycol sulfate levels in the rat cerebral cortex. Eur. J. Pharmacol. 21: 30S-310, 1973.

Krettek, J. E. and Price, J. L.: Amygdaloid projections to subcortical structure within the basal forebrain and the brainstem in the rat and the cat. J. Comp. Neurol. 178: 2SS-280, 1978.

Krieg, W. J. S.: The hypothalamus of the albino rat. J. Comp. Neurol. SS: 19-89, 1932.

Krieger, D. T. and Hauser, H.: Comparison of synchronization of circadian corticosteroid rhythms by photoperiod and food. Proc. Natl. Acad. Sci. (USA) 1975: 1577-1581, 1978.

Kuhar, M. J.: Neuroanatomical substrates of anxiety: a brief survey. T.I.N.S. July: 307-311, 1986.

Kuypers, H. G. J. M. and Maisky, V. A.: Retrograde axonal transport of horseradish peroxidase from spinal cord to brainstem cell groups in the cat. Neuroscience Letters 1: 9, 1973.

Kvetnansky, R., Kopin, I. J., and Saavedra, J.M.: Changes in epineph­rine in individual hypothalamic nuclei after immobilization stress. Brain Research lSS: 387-390, 1978.

Leckie, B. J.: Inactive renin: an attempt at a perspective. Clinical Science 60: 119-130, 1981.

Leenin, F. H. and Shapiro, A. P.: Effects of intermittent electric shock on plasma renin activity in rats. Proc. Soc. Exp. Biol. Med. 146: 534-S38, 1974.

Leibowitz, S. F.: Pattern of drinking and feeding produced by hypothal­amic norepinephrine injection in the satiated rat. Physiol. Behav. 14: 731-742, 1975.

Leibowitz, S. F., Diaz, S., and Spencer, L.: Adrenergic and neuropeptide Y systems in the hypothalamic paraventricular nucleus affect circulating levels of corticosterone and glucose. Neuroscience Abstracts 12(2): 782, 1986.

166

Leibowitz, S. F., Hammer, N. J., and Chang, K.: Hypothalamic paraventri­cular nucleus lesions produce overeating and obesity in the rat. Physiol. Behav. 27: 1031-1040, 1981.

Leibowitz, S. F., Roland, C. R., Hor, L., and Squillari, V.: Noradrener­gic feeding elicited via the paraventricular nucleus is dependent upon circulating corticosterone. Physiol. Behav. 32: 8S7-864, 1984.

Lemmer, B., Winkler, H., Ohm, T., and Fink, M.: Chronopharmacokinetics of beta-receptor blocking drugs of different lipophilicity (propranolol, metoprolol, sotalol, atenolol) in plasma and tissues after single and multiple dosing in the rat. Naunyn-Schmiedeberg's Arch Pharmacol 330: 42-49, 198S.

Lenz, H. J., Raedler, A., Greten, H., and Brown, M. R.: CRF initiates biological actions within the brain that are observed in response to stress. Am. J. Physiol. 2S2: R34-R39, 1987.

Leranth, C., Antoni, F. A., and Palkovits, M.: Ultrastructural demonstration of ovine CRF-like immuno-reactivity (oCRF-LI) in the rat hypothalamus: processes of magnocellular neurons establish membrane specializations with parvocellular neurons containing oCRF-LI. Regul. Peptides 6: 179-188, 1983.

Leroux, P. and Pelletier, G.: Radioautographic study of binding and internalization of corticotropin-releasing factor by rat anterior pituitary corticotrophs. Endocrinology 114: 14-21, 1984.

Levin, M. C., Sawchenko, P. E., Howe, P. R. C., Bloom, S. R., and Polak, J.M.: Organization of galanin-immunoreactive inputs to the paraventric­ular nucleus with special reference to their relationship to catechol­aminergic afferents. The Journal of Comparative Neurology 261: 562-S82, 1987.

Li, C. H. , Evans, H. M. , and Simpson, M. E. : Adrenocorticotropin hormone. J. Biol. Chem. 149: 413-424, 1943.

Lightman, S. L., Todd, K., and Everitt, B. J.: Ascending noradrenergic projections from the brainstem: Evidence for a major role in the regulation of blood pressure and vasopressin secretion. Experimental Brain Research SS: 14S-151, 1984.

Lind, R. W., Swanson, L. W., Bruhn, T. 0., and Ganten, D.: The distribution of angiotensin II- immunoreactive cells and fibers in the paraventriculo-hypophysial system of the rat. Brain Research 338: 81-89, 198S.

Lindvall, O. and Bjorklund, A.: The organization of ascending catechol­amine neuron systems in the rat brain. Acta Physiol. Scand. 412: 1-48, 1974.

167

Liposits, Z., Gores, T., Setalo, G., Lengvari, I., Flerko, B., Vigh, S., and Schally, A.V.: Ultrastructural characteristics of immunolabelled, corticotropin releasing factor (CRF)-synthesizing neurons in the rat brain. Cell Tissue Res. 229: 191-196, 1983.

Lipos its, Z. , Lengvari, I. , Vigh, S. , Gores, T. , Setalo, G. , and Flerko, B. The paraventriculo- infundibular corticotropin releasing factor (CRF)-Immunoreactive System of the Rat- I. Morphological Aspects. In K. W. McKerns and V. Pantie (Eds.), Neuroendocrine Correlates of Stress. New York: Plenum Press, 1985.

Liposits, Z. and Paull, W. K.: Ultrastructural alterations of the paraventriculo-infundibular corticotropin releasing factor (CRF)-immuno­reactive neuronal system in long term adrenalectomized rats. Peptides 6: 1021-1036, 1985.

Liposits, Z., Phelix, C., and Paull, W. K.: Electron microscopic analysis of tyrosine hydroxylase, dopamine-beta-hydroxylase and phenyl­ethanolamine-N-methyl transferase immunoreactive innervation of the hypothalamic paraventricular nucleus in the rat. Histochemistry 84: 105-120, 1986.

Liposits, Z., Phelix, C., and Paull, W. K.: Adrenergic innervation of corticotropin releasing factor (CRF)- synthesizing neurons in the hypothalamic paraventricular nucleus of the rat. Histochemistry 84: 201-205, 1986.

Liposits, Z., Phelix, C., and Paull, W. K.: Synaptic interaction of serotonergic axons and corticotropin releasing factor (CRF) synthesizing neurons in the hypothalamic paraventricular nucleus of the rat. A light and electron microscopic immunocytochemical study. Histochemistry 86: 541-549, 1987.

Liposits, Z., Uht, R. M., Harrison, R. W., Gibbs, F. P., Paull, W. K., and Bohn, M. C.: Ultrastructural localization of glucocorticoid receptor (GR) in hypothalamic paraventricular neurons synthesizing corticotropin releasing factor (CRF). Histochemistry 87: 407-412, 1987.

Ljungdahl, A., Hokfelt, T., and Nilsson, G.: Distribution of substance P-like immunoreactivity in the central nervous system of the rat. I. Cell bodies and nerve terminals. Neuroscience 3: 861-944, 1978.

Loeb, R. F., Atchley, D. W., Benedict, E. M., and Leland, J.: Electro­lyte balance studies in adrenalectomized dogs with particular reference to the excretion of sodium. J. Exp. Med. 57: 775-792, 1933.

Loeffler, J. R., Stockigt, J. R., and Ganong, W. F.: Effect of alpha­and beta-adrenergic blocking agents on the increase in renin secretion produced by stimulation of the renal nerves. Neuroendocrinology 10: 129-138, 1972.

168

Loizou, L. A.: Projections of the nucleus locus coeruleus in the albino rat. Brain Research 15: 563-566, 1969.

Long, C. N. H., Katzin, B., and Fry, E. G.: Adrenal cortex and carbohydrate metabolism. Endocrinology 26: 309-344, 1940.

Lorens, S. A. and Guldberg, H. C.: Regional 5-hydroxytryptamine following selective midbrain raphe lesions. Brain Research 78: 45-56, 1974.

Lorens, S. A., Mitsushio, H., and Van de Kar, L. D. Effects of the 5-HTlA agonist ipsapirone on the behavioral, endocrine and neurochemical responses to conditioned fear. In T. Archer,P. Bevan,C. Broekkamp and A. Cools (Eds.), Behavioral Pharmacology of 5-HT. : In Press, 1989.

Lorens, S. A. and Van de Kar, L. D.: Differential effects of serotonin (5-HTlA and 5-HT2) agonists and antagonistson renin and corticosterone secretion. Neuroendocrinology 45: 305-310, 1987.

Luiten, P. G. M., ter Horst, G. J., Karst, H., and Steffens, A. B.: The course of paraventricular efferents to autonomic structures in medulla and spinal cord. Brain Research 329: 374-378, 1985.

Madison, D. V. and Nicoll, R. A.: Actions of noradrenaline recorded intracellularly in rat hippocampal CAl pyramidal neurons, in vitro. J. Physiol. 372: 221-244, 1986.

Makara, G. B., Stark, E., Kapocs, G., and Antoni, F. A.: Long-term effects of hypothalamic paraventricular lesion on CRF content and stimulated ACTH secretion. Am. J. Physiol. 250: E319-E324, 1986.

Makara, G. B., Stark, E., Karteszi, M., Palkovits, M., and Rappay, G.: Effects of paraventricular lesions on stimulated ACTH release and CRF in stalk-median eminence of the rat. Am. J. Physiol. 240: E441-E446, 1981.

Manuck, S. B., Giordani, B., McQuaid, K. J., and Garritz, S. J.: Behaviorally-induced cardiovascular reactivity among sons of reported hypertensive normotensive patients. J. Psychosom. Res. 25: 261, 1981.

Maran, J. W. and Yates, F. E.: Cortisol secretion during intrapituitary infusion of angiotensin II in conscious dogs. Am. J. Physiol. 233: E273-E285, 1977.

Marino, L.A., DeBellis, M. D., and Leibowitz, S. F.: Alpha-2 adrenergic receptors in the paraventricular nucleus mediate feeding induced by norepinephrine and clonidine. Soc. Neurosci. Abstr. 9: 467, 1983.

Markowska, A., Bakke, H.K., Walther, B., and Ursin, H.: Comparison of electrolytic and ibotenic acid lesions in the lateral hypothalamus. Brain Research 328: 313-323, 1985.

Mason, J. W.: Plasma 17-corticosteroid stimulation of the amygdaloid complex in Physiol. 196: 44-48, 1959.

169

levels during electrical conscious monkeys. Am. J.

Mason, J. W., Maher, J. T., Hartley, L. H., Mougey, E. H., Perlow, M. J., and Jones, L.G. Selectivity of corticosteroid and catecholamine response to various natural stimuli. In G. Serban (Ed.), Psychopathology of Human Adaptation. New York: Plenum Press, 1976.

Mcintyre, M. and Stein, D. G.: Differential effects of one- vs. two-stage amygdaloid lesions on activity, exploratory, and avoidance behavior in the albino rat. Behav. Biol. 9: 451-465, 1973.

McKellar, S. and Loewy, A. D.: Organization of some brain stem afferents to the paraventricular nucleus of the hypothalamus in the rat. Brain Res 217: 351-357, 1981.

Meister, B., Hokfelt, T., and Staines, W.: Presence of glutamic acid decarboxylase in some corticotropin-releasing factor- (CRF-) containing neurons of the paraventricular nucleus. Acta Physiol. Scand. 131: 471-473, 1987.

Meldrum, M. J., Xue, C. S., Badino, L., and Westfall, T. C.: Angiotensin facilitation of noradrenergic neurotransmission in central tissues of the rat: effects of sodium restriction. Journal of Cardiovascular Pharmacology 6: 989-995, 1984.

Menard, J. and Catt, K. J.: Measurement of renin activity, concentration and substrate in rat plasma by radioimmunoassay of angiotensin I. Endocrinology 90: 422-430, 1972.

Merchenthaler, I., Vigh, S., Petrusz, P., and Schally, A. V.: Im­munocytochemical localization of corticotropin releasing factor (CRF) in the rat brain. Am. J. Anat. 165: 385, 1982.

Merchenthaler, I., Vigh, S., Petrusz, P., and Schally, A. V.: The paraventriculo-infundibular corticotropin releasing factor (CRF) pathway as revealed by immunocytochemistry in long-term hypophysectomized or adrenalectomized rats. Regulatory Peptides 5: 295-305, 1983.

Mermet, C. C. and Gonon, F. G.: Ether stress stimulates noradrenaline release in the hypothalamic paraventricular nucleus. Neuroendocrinology 47: 75-82, 1988.

Meyer, D. K., Phillips, M. I., and Eiden, L.: Studies on the presence of angiotensin II in rat brain. J. Neurochem. 38: 816-820, 1982.

Meyerhoff, J. L., Mougey, E. H., and Kant, G. J.: Paraventricular lesions abolish the stress-induced rise in pituitary cyclic adenosine monophosphate and attenuate the increase in plasma levels of proopio­melanocortin-derived peptides and prolactin. Neuroendocrinology 46: 222-230, 1987.

170

Michelakis, A. M. andMizukoshi, H.: Distribution and disappearance rate of renin in man and dog. J. Clin. Endocrinol. Metab. 33: 27-34, 1971.

Middlemiss, D. N., Blakeborough, L., and Leather, S. R.: Direct evidence for an interaction of beta-adrenergic blockers with the serotonin receptor. Nature 267: 289-290, 1977.

Mills, F. J.: The endocrinology of stress. Aviat. Space Environ. Med. 56: 642-650, 1985.

Modlinger, R. S., Schonmuller, J.M., and Arora, S. P.: Stimuation of aldosterone, renin and cortisol by tryptophan. J. Clin. Endocrinol. Metab. 48: 599-603, 1979.

Moga, M. M. and Gray, T. S.: Evidence for corticotropin-releasing factor, neurotensin and somatostatin from the central nucleus of the amygdala to the parabrachial nucleus in the rat. J. Comp. Neurol. 241: 275, 1985.

Moore, R. Y. and Bloom, F. E.: Central catecholamine neuron systems: anatomy and physiology of the norepinephrine and epinephrine systems. Ann. Rev. Neurosci. 2: 113-168, 1979.

Morris, B. J., Davis, J. 0., Zatzman, M. L., and Williams, G. M.: The renin-angiotensin-aldosterone system in rabbits with congestive heart failure produced by aortic constriction. Gire. Res. 40: 275-282, 1977.

Mukherjee, A., Kulkarni, P., Mccann, S. M., and Negro-Vilar, A.: Evidence for the presence and characterization of angiotensin II receptors in rat anterior pituitary membranes. Endocrinology 110: 665-667, 1982.

Mulder, G. H. and Smelik, P. G.: A perfusion technique for the study of sites of action of glucocorticoids in rat hypothalamus-pituitary-adrenal system in vitro. Endocrinology 100: 1143-1152, 1977.

Nakani, T., Audhya, T., Kani, N., and Hollander, C. S.: Evidence for a role of endogenous corticotropin-releasing factor in cold, ether, immobilization and traumatic stress. Proc. Natl. Acad. Sci. (USA) 82: 1247-1251, 1985.

Nasjletti, A. and Masson, G. M. C.: Effects of corticosteroids on plasma angiotensinogen and renin activity. Am. J. Physiol. 217(5): 1396-1400, 1969.

Nasjletti, A. and Masson, G. M. C.: Stimulation of angiotensinogen formation by renin and angiotensin. Proc. Soc. Exp. Biol. Med. 142: 307-310, 1973.

Natcheff, N., Logofeton, A., and Tzaneva, N.: Hypothalamic control of plasma renin activity. Pflugers Archiv. 371: 279-285, 1977.

171

Nicoll, R. A.: Neurotransmitters can say more than just "yes" or "no". Trends in NeuroSciences 5(11): 269-374, 1982.

Nielsen, A. H. and Poulsen, K.: Evidence for a predominant renal secretion and clearance of inactive plasma renin, studied by in vivo inhibition of protein synthesis. Journal of Hypertension 6: 129-137, 1988.

Norgren, R.: Projections from the nucleus of the solitary tract in the rat. Neuroscience 3: 207-218, 1978.

Oates, H. F., Fretten, J. A., and Stokes, G. S.: Disappearance rate of circulating renin after bilateral nephrectomy in the rat. Clin. Exp. Pharmacol. Physiol. 1: 547-549, 1974.

Olds, J. and Olds, M. In F. Barron {Ed.), New Directions in Psychology. New York: Holt, Rinehart and Winston, 1965.

Olivecrona, H.: Paraventricular nucleus and pituitary gland. Acta Physiol. Scand. 40(136): 1-178, 1957.

Olschowka, J. A., O'Donohue, T. L., Mueller, G. P., and Jacobowitz, D. M.: Hypothalamic and extrahypothalamic distribution of CRF-like im­munoreactive neurons in the rat brain. Neuroendocrinology 35: 305-308, 1982.

Ono, N., Samson, W. K., Mc Donald, J. K., Lumpkin, M. D., Bedran de Castro, I. C., and Mc Cann, S. M.: Effect of intravenous and intra­ventricular injection of antisera directed against corticotropin-releasing factor in the secretion of anterior pituitary hormones. Proc. Natl. Acad. Sci. (USA) 88: 7787-7790, 1985.

Oparil, S., Vassaux, C., Sander, C. A., and Haber, E.: Role of renin in acute postual homeostasis. Circulation 41: 89-95, 1970.

Osborn, J. L., DiBona, G. F., and Thames, M. D.: Beta-1 receptor mediation of renin secretion elicited by low-frequency renal nerve stimulation. J. Pharmacol. Exp. Ther. 216: 265-269, 1981.

Osmond, D. H., Wilczynski, E. A., and Loh, A. Y.: Peculiarities of plasma prorenin measurements in man, dog, and rat, and their theoretical implications. Clin. Exp. Hypertension 4: 2213-2230, 1982.

Padfield, P. L. and Morton, J. J.: Effects of angiotensin II on arginine-vasopressin in physiological and pathological situations in man. J. Endocr. 74: 251-259, 1977.

Palkovi ts, M. : Catecholamines in the hypothalamus: An anatomical review. Neuroendocrinology 33: 123-128, 1981.

Palkovits, M., Brownstein, M., Saavedra, J. M., and Axelrod, J.: Norepinephrine and dopamine content of hypothalamic nuclei of the rat. Brain Research 77: 137-149, 1974.

172

Palkovits, M., Eskay, R. L., and Antoni, F. A.: Atrial natriuretic peptide in the median eminence is of paraventricular nucleus origin. Neuroendocrinology 46: 542-544, 1987.

Palkovits, M., Saavedra, J. M., Jacobowitz, D. M., Kizer, J. S., Zaborszky, I., and Brownstein, M. J.: Serotonergic innervation of the forebrain: effect of lesions on serotonin and tryptophan hydroxylase levels. Brain Research 130: 121-134, 1977.

Parent, A., Descarries, L., and Beaudet, A.: Organization of ascending serotonin systems in the adult rat brain. A radioautographic study after intraventricular administration of 3H 5-hydroxytryptamine. Neuroscience 6: 45-138, 1981.

Parfrey, P. S., Markandu, N. D., Roulston, J. E., Jones, B. E., Jones, J. C., and MacGregor, G. A.: Relation between arterial pressure, dietary sodium intake, and renin system in essential hypertension. Br. Med. J. 283: 94-97, 1981.

Paris, J. M., Lorens, S. A., Van de Kar, L. D., Urban, J. H., Rich­ardson-Morton, K. D., and Bethea, C. L.: A comparison of acute stress paradigms: Hormonal responses and hypothalamic serotonin. Physiol Behav 39(1): 33-43, 1987.

Passo, S. S., Assaykeen, T. A., Goldfien, A., and Ganong, W. F.: Effect of alpha- and beta-adrenergic blocking agents on the increase in renin secretion produced by stimulation of the medulla oblongata in dogs. Neuroendocrinology 7: 97-104, 1971.

Paull, W. K. and Gibbs, F. P.: The corticotropin releasing factor (CRF) neurosecretory system in intact, adrenalectomized, and adrenalec­tomized-dexamethasone treated rats. Histochemistry 78: 303-316, 1983.

Pavlov, I. P. Conditioned Reflexes, translated and edited by G.V. Anrep. Oxford: Oxford University Press, 1927.

Peach, M. J. Adrenal Medulla. In I. M. Page and F. M. Bumpus (Eds.), Angiotensin. New York: Springer-Verlag, 1974.

Peach, M. J., Bumpus, F. M., and Khairallah, P.A.: Release of adrenal catecholamines by angiotensin I. J. Pharmacol. Exp. Ther. 176(2): 366-376, 1971.

Pearlmutter, A. F., Rapino, E., and Saffran, M.: A semiautomated in vitro assay for CRF: activities of peptides related to oxytocin and vasopressin. Neuroendocrinology 15: 106-119, 1974.

Perez de la Mora, M., Possani, L. D., Tapia, R., Teran, L., Palacios, R., Fuxe, K., Hokfelt, T., and Ljungdahl, A.: Demonstration of central gamma-aminobutyrate-containing nerve terminals by means of antibodies against glutamate decarboxylase. Neuroscience 6: 875-895, 1981.

173

Petrusz, P., Merchenthaler, I., Maderdrut, J. L., and Heitz, P. U.: Central and peripheral distribution of corticotropin-releasing factor. Fed. Proc. 44: 229-23S, 198S.

Pettinger, W. A., Augusto, L., and Leon, A. S. Alteration of renin release by stress and adrenergic receptor and related drugs in unanes­thetized rats. In C. M. Bloor (Ed.), Comparative Pathophysiology of Circulatory Disturbances. New York: Plenum Press, 1972.

Pettinger, W. A. and Keeton, T. K.: Altered renin release and propran­olol potentiation of vasodilatory drug hypotension. J. Clin. Invest. SS: 236-243, 197S.

Pettinger, W. A., Keeton, T. K., Campbell, W. B., and Harper, D. C.: Evidence for a renal alpha-adrenergic receptor inhibiting renin release. Circ. Res. 38: 338-346, 1976.

Phillips, M. I. and Stenstrom, B.: An analysis of angiotensin II and III in the brain with high pressure liquid chromatography. Physiologist 26: A47, 1983.

Piekut, D. T. and Joseph, S. A.: Relationship of CRF-immunostained cells and magnocellular neurons in the paraventricular nucleus of rat hypothal­amus. Peptides 6: 873-882, 198S.

Piekut, D. T. and Joseph, S. A.: Co-existance of CRF and vasopressin immunoreactivity in parvocellular paraventricular neurons of rat hypothalamus. Peptides 7: 891-898, 1986.

Pittman, Q. J., Blume, H. W., and Renaud, L. P.: Connections of the hypothalamic paraventricular nucleus with the neurohypophysis, median eminence, amygdala, lateral septum and midbrain periaqueductal gray: An electrophysiological study in the rat. Brain Research 21S: lS-28, 1981.

Plotsky, P. M.: Hypophyseotropic regulation of adenohypophyseal adrenocorticotropin secretion. Fed.Proc. 44: 207-213, 198S.

Plotsky, P. M.: Facilitation of immunoreactive corticotropin-releasing factor secretion into the hypophysial-portal circulation after activation of catecholaminergic pathways or central norepinephrine injection. Endocrinology 121: 924-930, 1987.

Pohl, M., Carayon, A., Cesselin, F., and Hamon, M.: Angiotensin II-like material extracted from the rat brain is distinct from authentic angiotensin II. J. Neurochem. Sl: 1407-1413, 1988.

Porter, J. P.: Low-Level Electrical Stimulation of the Paraventricular Nucleus Increases Plasma Renin Activity in Conscious Rats. Neuroscience Abst. 12: S39, 1986.

174

Porter, J. P.: Electrical stimulation of the paraventricular nucleus increases plasma renin activity. Am. J. Physiol. 254: R325-R330, 1988.

Porter, J. P.: The renin response to aortic occlusion is enhanced by stimulation of the hypothalamus. Hypertension 12: 52-58, 1988.

Powell, T. P. S., Cowan, W. M., and Raisman, G.: The central olfactory connections. J. Anat. 99: 791-813, 1965.

Printz, M. P., Ganten, D., Unger, T., and Phillips, M. I.: Minireview: The brain renin-angiotensin system 1982. Exp. Brain. Res. 4(supp): 4-52, 1982.

Privitera, P. J. , Webb, J. G. , and Walle, T. : Effects of centrally administered propranolol on plasma renin activity, plasma norepinephrine and arterial pressure. Eur J Pharmacol 54: 51-60, 1979.

Ramsay, D. J., Keil, L. C., Sharpe, M. C., and Shinsako, J.: Angiotensin II infusion increases vasopressin, ACTH and 11-hydroxy-corticosteroid secretion. Am. J. Physiol. 234: R66-R71, 1978.

Rasmussen, A. T.: Innervation of the hypophysis. Endocrinology 23: 263-278, 1938.

Rasmussen, K. and Jacobs, B. L.: Single unit activity of locus coeruleus neurons in the freely moving cat. II. Conditioning and pharmacologic studies. Brain Research 371: 335-344, 1986.

Redgate, E. S.: ACTH release evoked by electrical stimulation of brain stem and limbic system sites in the cat: the absence of ACTH release upon infundibular area stimulation. Endocrinology 4: 806-823, 1970.

Redgate, E. S. and Fahringer, E. E.: A comparison of the pituitary adrenal activity elicited by electrical stimulation of preoptic, amygdaloid and hypothalamic sites in the rat brain. Neuroendocrinology 12: 334-343, 1973.

Reid, I. A., Brooks, V. L., Rudolph, C. D., and Keil, L. C.: Analysis of the actions of angiotensin on the central nervous system of conscious dogs. Am. J. Physiol. 243: R82-R91, 1982.

Reid, I. A., MacDonald, D. M., Pachnis, B., and Ganong, W. F.: Studies concerning the mechanism of suppression of renin secretion by clonidine. J. Pharmacol. Exp. Ther. 192: 713-721, 1975.

Reid, I. A., Morris, B. J., and Ganong, W. F.: The renin-angiotensin system. Ann. Rev. Physiol. 40: 377-410, 1978.

Reid, I. A. and Ramsay, D. J.: The effect of intracerebroventricular administration of renin on drinking and blood pressure. Endocrinology 97: 536-542, 1975.

175

Reid, I. A., Schrier, R. W'., and Earley, L. E.: An effect of extrarenal beta adrenergic stimulation on the release of renin. J. Clin. Invest. 51: 1861-1869, 1972.

Reinstein, D. K., Lehnert, H., and W'urtman, R. J.: Dietary tyrosine suppresses the rise in plasma corticosterone following acute stress in rats. Life Sciences 37: 2157-2163, 1985.

Renaud, L. P. and Arnauld, E.: Supraoptic phasic neurosecretory neurons: response to stimulation of amygdala and dorsal hippocampus and sensitivity to microiontophoresis of amino acids and noradrenaline. Soc. Neurosci. Abstr. 5: 233, 1979.

Ricardo, J. A. and Koh, E.T.: Anatomical evidence of direct projections from the nucleus of the solitary tract to the hypothalamus, amygdala and other forebrain structures in the rat. Brain Research 153: 1-26, 1978.

Richardson, D., Stella, A., Keonetti, G., Barotorelli, A., and Zanchetti, A.: Mechanisms of renal release of renin by electrical stimulation of the brain stem of the cat. Circ. Res. 34: 425-434, 1974.

Richardson Morton, K. D., Johnson, M. D., and Van de Kar, L. D.: Effect of stress on renin secretion is not blocked by adrenal enucleation combined with chemical sympathectomy. Fed. Proc. 47: 3877, 1988.

Richardson Morton, K. D. , Van de Kar, L. D. , and Lorens , S . A. : Stress-induced renin and corticosterone release: effect of dorsal raphe lesions and 5-HT2 antagonist. Fed. Proc. 45: 3019, 1986.

Richardson Morton, K. D., Van de Kar, L. D., Lorens, S. A., Paris, J. M., Urban, J. H., and Kunimoto, K.: The effect of stress on renin and corticosterone secretion is blocked by electrolytic lesions in the mesencephalic dorsal raphe and hypothalamic paraventricular nuclei. Neurosci. Abst. 12: 290.15, 1986.

Richoux, J. P., Cordonnier, J. L., Bouhnik, J., Clauser, E., Corvol, P., Menard, J., and Grignon, G.: Imrnunocytochemical localization of angiotensinogen in rat liver and kidney. Cell Tissue Res. 233: 439-451, 1983.

Rivier, C., Rivier, J., and Vale, W.: Inhibition of adrenocorticotropic hormone secretion in the rat by imrnunoneutralization of cor­ticotropin-releasing factor. Science 218: 377-378, 1982.

Rivier, C. and Vale, W'.: Interaction (CRF) and arginine vasopressin (AVP) docrinology 113: 939-942, 1983.

of corticotropin-releasing factor on ACTH secretion in vivo. En-

Rivier, C. and Vale, W.: Modulation of stress-induced ACTH release by corticotropin-releasing factor, catecholamines and vasopressin. Nature 305: 325-327, 1983.

176

Rivier, J., Speiss, J., and Vale, W.: Characterization of rat hypothal­ainic corticotropin-releasing factor. Proc. Natl. Acad. Sci. 80: 4851-4855, 1983.

Robertson, G. L.: The regulation of vasopressin function in health and disease. Rec. Prog. Horm. Res. 33: 333-374, 1977.

Romero, J.C., Lazar, J. D., and Hoobler, S. W.: Effects of renal artery constriction and subsequent contralateral nephrectomy on the blood pressure, plasma renin activity, and plasma renin substrate concentration in rabbits. Lab. Invest. 22: 581-587, 1970.

Rossier, J., Battenberg, E., Pittman, Q., Bayon, A., Koda, L., Miller, R., Guilleman, R., and Bloom, F.: Hypothalainic enkephalin neurons may regulate the neurohypophysis. Nature 277: 653-655, 1979.

Rostand, S. G., Work, J., and Luke, R. G.: Effect of reduced chloride reabsorption on renin release in the isolated rat kidney. Pflugers Arch. 405: 46-51, 1985.

Russell, D. M., Dhariwal, A. P. S., Mc Cann, S. M., and Yates, F. E.: Inhibition by dexamethasone of the in vivo pituitary response to corticotropin releasing factor (CRF). Endocrinology 85: 512-521, 1969.

Saavedra, J. M., Israel, A., Plunkett, L. M., Kurihara, M., Shigematsu, K., and Correa, F.M.A.: Quantitative distribution of angiotensin II binding sites in rat brain by autoradiography. Peptides 7: 679, 1986.

Saavedra, J. M., Palkovits, M., Kizer, J. S., Brownstein, M., and Zivin, J. A.: Distribution of biogenic amines and related enzymes in the rat pituitary gland. Journal of Neurochemistry 25: 257-260, 1975.

Saffran, M. and Schally, A. V.: The status of the corticotropin releasing factor (CRF). Neuroendocrinology 24: 359-375, 1977.

Sahu, A., Kalra, S. P., Crowley, W. R., and Kalra, P. S.: Evidence that NPY-containing neurons in the brainstem project inot selected hypothalainic nuclei: implication in feeding behavior. Brain Research 457: 376-378, 1988.

Saper, C. B., Loewy, A. D., Swanson, L. W., and Cowan, W. M.: Direct hypothalamo-autonomic connections. Brain Research 117: 305-312, 1976.

Saper, C. B., Reis, D. J., and Joh, T.: Medullary catecholamine inputs to the anteroventral third ventricular cardiovascular regulatory region in the rat. Neuroscience Letters 42: 285-291, 1983.

Sapolsky, R. M., Krey, L. C., and Mc Ewen, B. S.: Stress down-regulates corticosterone receptors in a site-specific manner in the brain. Endocrinology 114: 287-292, 1984.

177

Sawchenko, P. E.: Adrenalectomy-induced enhancement of CRF and vasopressin immunoreactivity in parvocellular neurosecretory neurons: anatomic, peptide and steroid specificity. The Journal of Neuroscience 7(4): 1093-1106, 1987.

Sawchenko P.E.: Evidence for differential regulation of corticotropin­releasing factor and vasopressin immunoreactivities in parvocellular neurosecretory and autonomic-related projections of the paraventricular nucleus. Brain Research 437: 253-263, 1987.

Sawchenko, P. E. and Swanson, L. W.: A method for tracing biochemically defined pathways in the central nervous system using combined fluorescence retrograde transport and immunohistochemical techniques. Brain Research 210: 31-51, 1981.

Sawchenko, P. E. and Swanson, L. W.: Central noradrenergic pathways for the integration of hypothalamic neuroendocrine and autonomic responses. Science 214(6): 685-687, 1981.

Sawchenko, P. E. and Swanson, L. W.: The distribution and cells of or1g1n of some afferent projections to the paraventricular and supraoptic nuclei in the rat. Soc Neurosci 7: 325, 1981.

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 Research Reviews 4: 275-325, 1982.

Sawchenko, P. E. and Swanson, L. W.: Immunohistochemical identification of neurons in the paraventricular nucleus of the hypothalamus that project to the medulla or to the spinal cord in the rat. J. Comp. Neurol. 205: 260-272, 1982.

Sawchenko, P. E. and Swanson, L. W.: Localization, colocalization and plasticity of corticotropin-releasing factor immunoreactivity in rat brain. Fed Proc 44: 221-227, 1985.

Sawchenko, P. E., Swanson, L. W., Grzanna, R., Howe, P. R. C., Bloom, S. R., and Polak, J.M.: Colocalization of neuropeptide Y immunoreactivity in brainstem catecholaminergic neurons that project to the paraventricular nucleus of the hypothalamus. The Journal of Comparative Neurology 241: 138-153, 1985.

Sawchenko, P. E., Swanson, L. W., and Joseph, S. A.: The distribution and cells of origin of ACTH (1-39)-stained varicosities in the paraven­tricular and supraoptic nuclei. Brain Res. 232: 265-274, 1982.

Sawchenko, P. E., Swanson, L. W., Steinbusch, H.W.M., and Verhofstad, A.A.J.: The distribution and cells of origin of serotonergic inputs to the paraventricular and supraoptic nuclei of the rat. Brain Res. 277: 355-360, 1983.

178

Sawchenko, P. E., Swanson, L. W., and Vale, W. W.: Co-expression of corticotropin-releasing factor and vasopressin immunoreactivity in parvocellular neurosecretory neurons of the adrenalectomized rat. Proc. Natl. Acad. Sci. (USA) 81: 1883-1887, 1984.

Sayers, G., White, A., and Long, C. N. H.: Preparation and properties of pituitary adrenotropic hormone. J. Biol. Chem. 149: 425-436, 1943.

Scapagnini, U., Moberg, G. P., Van Loon, G. R., De Groot, J., and Ganong, W. F.: Relation of brain 5-HT content to the diurnal variation in plasma corticosterone in the rat. Neuroendocrinology 7: 90-96, 1971.

Scapagnini, U., Van Loon, G. R., Moberg, G. P., and Ganong, W. F.: Effect of alpha-methyl-para-tyrosine on the circadian variation of plasma corticosterone in rats. European J. Pharmacol. 11: 266-268, 1970.

Schanberg, S. M., Schildkraut, J. J., Breese, G. R., and Kopin, I. J.: Metabolism of norepinephrine-H3 in rat brain- -Identification of conjugated 3-methoxy-4-hydroxyphenylglycol as the major metabolite. Biochem. Pharmac. 17: 247-254, 1968.

Schulte, W., Neus, H., Thones, M., and von Eiff, A. W.: Basal blood pressure variability and reactivity of blood pressure to emotional stress in essential hypertension. Basic Res. Cardiol. 79: 9, 1984. Scroop, G. C., Katie, F., Joy, M. D., and Lowe, R. D.: Importance of central vasomotor effects in angiotensin-induced hypertension. Brit. Med. J. 1: 324-326, 1971.

Sealey, J. E., Atlas, S. A., Laragh, J. H., Silverberg, M., and Kaplan, A. P.: Initiation of plasma prorenin activation by Hageman factor­dependent conversion of plasma prekallikrein to kallikrein. Proc. Natl. Acad. Sci. 76: 5914-5918, 1979.

Selye, H.: A syndrome produced by diverse nocuous agents. Nature 138: 32, 1936.

Selye, H.: An attempt at a natural classification of the steroids. Nature 151: ]662-663, 1943.

Selye, H.: The general adaptation syndrome and the diseases of adaptation. Journal of Clinical Endocrinology 6: 117-231, 1946.

Selye, H. The Physiology and Pathology of Exposure to Stress. Montreal: Medical Publishers, Inc., 1950.

Selye, H.: Stress and aerospace medicine. Aerospace Med. 44(2): 190-193, 1973.

Selye, H. Stress in Health and Disease. London: Butterworths, 1976.

179

Shade, R. E., Davis, J. 0., Johnson, J. A., and Witty, R. T.: Effects of arterial infusion of sodium and potassium on renin secretion in the dog. Circ. Res. 31: 719-727, 1972.

Share, L.: Interrelations between vasopressin and the renin-angiotensin system. Fed. Proc. 38: 2267-2271, 1979.

Sherlock, D. A., Field, P. M., and Raisman, G.: Retrograde transport of horseradish peroxidase in the magnocellular neurosecretory system of the rat. Brain Research 88: 403-414, 1975.

Shor-Posner, G., Azar, A. P., Jhanwar-Uniyal, M., Filart, R., and Leibowitz, S. F.: Destruction of noradrenergic innervation to the paraventricular nucleus: Deficits in food intake, macronutrient selection, and compensatory eating after food deprivation. Pharmacology Biochemistry and Behavior 25: 381-392, 1986.

Sidman, M.: Avoidance conditioning with brief shock and no extroceptive warning signal. Science 118: 157-158, 1953.

Sigg, E. B., Keim, K. L., and Sigg, T. D.: On the Mechanism of Renin Release by restraint stress in rats. Pharmac. Biochem. Behav. 8(1): 47-50, 1978.

Silverman, A. J., Hoffman, D. L., and Zimmerman, E. A.: The descending afferent connections of the paraventricular nucleus of the hypothalamus. Brain Res. Bull. 6: 47-61, 1981.

Silverman, A. J. and Zimmerman, E. A.: Magnocellular neurosecretory system. Ann. Rev. Neurosci. 6: 357-380, 1983.

Sims, K. B., Hoffman, D. L., Said, S. I., and Zimmerman, E. A.: Vasoactive intestinal polypeptide (VIP) in mouse and rat brain: An immunocytochemical study. Brain Research 186: 165-183, 1980.

Skeggs, L. T., Kahn, J. R., and Shumway, N. P.: The preparation and function of the hypertensin-converting enzyme. J. Exp. Med. 103: 295-299, 1956.

Skofitsch, G., Insel, T. R., and Jacobowitz, D. M.: Binding sites for corticotropin releasing factor in sensory areas of the rat hindbrain and spinal cord. Brain Research Bulletin 15: 519-522, 1985.

Smith, 0. A., DeVito, J. L., and Astley, C. A.: Organization of central nervous system pathways influencing blood pressure responses during emotional behavior. Clin. and Exper. Hyper. A6(1&2): 185-204, 1984.

Smythe, G. A. , Bradshaw, J. E. , and Vining, R. F. : Hypothalamic Monoamine Control of Stress-Induced Adrenocorticotropin Release in the Rat. Endocrinology 113(3): 1062-1071, 1983.

180

Soffer, A.: Regatine and benadine in the diagnosis of pheochromocytoma. Med. Clin. North Am. 38: 375-384, 1954.

Spevack, A. A., Campbell, C. T., and Drake, L.: Effect of amygdalectomy on habituation and CER in rats. Physiol. Behav. 15: 199-207, 1975.

Spinedi, E., Johnston, C. A., Chisari, A., and Negro-Vilar, A.: Role of central epinephrine on the regulation of corticotropin-releasing factor and adrenocorticotropin secretion. Endocrinology 122(5): 1977-1983, 1988.

Spinedi, E. and Negro-Vilar, A.: Angiotensin II and ACTH release: Site of action and potency relative to corticotropin releasing factor and vasopressin. Neuroendocrinology 37: 446-453, 1983.

Squire, L. R. and Zola-Morgan, S. The neurology of memory: the case for correspondence between the findings for man and non-human primate. In J. A. Deutsch (Ed.), The Physiological Basis of Memory. New York: Academic Press, 1983.

Sripairojthikoon, W. and Wyss, J.M.: Cells of origin of the sympathetic renal innervation in rat. Am. J. Physiol. 252: F957-F963, 1987.

Steel, R. G. D. and Torrie, J. H. Principles and Procedures of Statistics with Special Reference to the Biological Sciences. New York: Mc Graw Hill, 1960.

Steele, M. K., Negro-Vilar, A., and Mccann, S. M.: Effect of angiotensin II on in vivo and in vitro release of anterior pituitary hormones in the female rat. Endocrinology 109: 893-899, 1981.

Steinbusch, H. W. M.: Distribution of serotonin-immunoreactivity in the central nervous system of the rat--cell bodies and terminals. Neuroscience 6: 557-618, 1981.

Stellar, E.: The psychology of motivation. Psychol. Rev. 61: 5-22, 1954.

Stockigt, J. R., Collins, R. D., and Biglieri, E.G.: Determination of Plasma renin concentration by angiotensin I radioimmunoassay. Circulation Research 28/29: 175-180, 1971.

Stone, E. A.: Effect of stress on sulfated glycol metabolites of brain norepinephrine. Life Sciences 16: 1725-1730, 1975.

Stone, E. A.: Central noradrenergic activity and the formation of glycol sulfate metabolites of norepinephrine. Life Sciences 19: 1491-1498, 1976.

Suemaru, S., Hashimoto, K., and Ota, Z.: Brain corticotropin-releasing factor (CRF) and catecholamine responses in acutely stressed rats. Endocrinol. Japon. 32(5): 709-718, 1985.

181

Sugden, R. F. and Eccleston, D.: Glycol sulphate ester formation from [14C]noradrenaline in brain and the influence of a COMT inhibitor. J. Neurochem. 18: 2461-2468, 1971.

Swaab, D. F., Nijveldt, F., and Pool, C. W.: Distribution of oxytocin and vasopressin cells in the rat supraoptic and paraventricular nucleus. J. Endocrinol. 67: 461-462, 1975.

Swanson, L. W. The hippocampus and the concept of the limbic system. In W. Siefert (Ed.), Neurobiology of the hippocampus. New York: Academic Press, 1983.

Swanson, L. W. The hypothalamus. In A. Bjorklund,T. Hokfelt and L. W. Swanson (Eds.), Handbook of Chemical Neuroanatomy: Integrated systems of the CNS. Amsterdam: Elsevier, 1987.

Swanson, L. W. and Cowan, W. M.: An autoradiographic study of the organization of the efferent connections of the hippocampal formation in the rat. J. Comp. Neural. 172:49-84, 1977.

Swanson, L. W. and Cowan, W. M.: The connections of the septal region in the rat. J. Comp. Neural. 186: 621-656, 1979.

Swanson, L. W., Kohler, C., and Bjorklund, A. The limbic region. I: The septohippocampal system. In A. Bjorklund,T. Hokfelt and L. W. Swanson (Eds.), Handbook of Chemical Neuroanatomy. Integrated Systems of the CNS. Amsterdam: Elsevier Science Publishers B.V., 1987.

Swanson, L. W. and Kuypers, H. G. J.M.: The paraventricular nucleus of the hypothalamus: Cytoarchitectonic subdivisions and the organization of projections to the pituitary, dorsal vagal complex and spinal cord as demonstrated by retrograde fluorescence double-labeling methods. J. Comp. Neural. 194: 555-570, 1980.

Swanson, L. W. and McKellar, S.: The distribution of oxytocin- and neurophysin-stained fibers in the spinal cord of the rat and monkey. J. Comp. Neurol. 188: 87-106, 1979.

Swanson, L. W. and Sawchenko, P. E.: Paraventricular nucleus: a site for the integration of neuroendocrine and autonomic mechanisms. Neuroen­docrinology 31: 410-417, 1980.

Swanson, L. W. and Sawchenko, P. E.: Hypothalamic Integration: Organization of the paraventricular and supraoptic nuclei. Ann. Rev. Neurosci. 6: 269-324, 1983.

Swanson, L. W., Sawchenko, P. E., Berod, A., Hartman, B. K., Helle, K. B., and Vanorden, D. E.: An immunohistochemical study of the organization of catecholaminergic cells and terminal fields in the paraventricular and supraoptic nuclei of the hypothalamus. Journal of Comparative Neurology 196: 271-285, 1981.

182

Swanson, L. W., Sawchenko, P. E., Rivier, J., and Vale, W. W.: Organization of ovine corticotropin-releasing factor immunoreactive cells and fibers in the rat brain: An immunohistochemical study. Neuroen­docrinology 36: 165-186, 1983.

Swanson, L. W., Sawchenko, P. E., Wiegand, S. J., and Price, J. L.: Separate neurons in the paraventricular nucleus project to the median eminence and to the medulla or spinal cord. Brain Research 197: 207-212, 1980.

Szafarczyk, A., Alonso, G., Ixart, G., Malaval, F., and Assenmacher, I.: Diurnal-stimulated and stress-induced ACTH release in rats is mediated by ventral noradrenergic bundle. Am. J. Physiol. 249: E219-E226, 1985.

Szafarczyk, A., Malaval, F., Laurent, A., Gibaud, R., and Assenmacher, I.: Further evidence for a central stimulatory action of catecholamines on adrenocorticotropin release in the rat. Endocrinology 121(3): 883-892, 1987.

Tager, H., Hohenboken, M., Markese, J., and Dinerstein, R. J.: Identification and localization of glucagon-related peptides in rat brain. Proc. Natl. Acad. Sci. (USA) 77: 6229-6233, 1980.

Taher, M. S., McLain, L. G., McDonald, K. M., and Schrier, R. W.: Effects of beta-adrenergic blockade on renin response to renal nerve stimulation. J. Clin. Invest. 57:459-465, 1976.

Tanaka, M., Kohno, Y., Nakagawa, R., Ida, Y., Takeda, S., and Nagasaki, N.: Time-related differences in noradrenaline turnover in rat brain regions by stress. Pharmacol. Biochem. Behav. 16(2): 315-319, 1982.

Taugner, C., Poulsen, K., Hackenthal, E., and Taugner, R.: Immuno­cytochemical localization of renin in mouse kidney. Histochemistry 62: 19-27, 1979.

Taugner, R., Buhrle, C. P., and Nobiling, R.: Ultrastructural changes associated with renin secretion from the juxtaglomerular apparatus of mice. Cell Tiss. Res. 237: 459-472, 1984.

Thierry, A. M., Jovey, F., Glowinski, J., and Kety, S. S.: Effects of stress on the metabolism of norepinephrine, dopamine and serotonin in the central nervous system in the rat. I. Modifications of norepinephrine turnover. J. Pharmacol. Exp. Ther. 163: 163-171, 1968.

Tilde rs, F. J. H. , Berkenbosch, F. , Vermes, I. , Linton, E. A. , and Smelik, P. G.: Role of epinephrine and vasopresson in the control of the pituitary-adrenal response to stress. Fed. Proc. 44(1): 155-160, 1985.

Tobian, L., Tomboulian, A., and Janacek, J.: The effect of high perfusion pressures on the granulation of juxtaglomerular cells in an isolated kidney. J. Clin. Invest. 38: 605-610, 1959.

183

Tramu, G., Croix, C., and Pillez, A.: Ability of the CRF immunoreactive neurons of the paraventricular nucleus to produce a vasopressin-like material. Neuroendocrinology 37: 467-469, 1983.

Tribollet, E. and Dreifuss, J. J.: Localization of neurones projecting to the hypothalamic paraventricular nucleus area of the rat: A horseradish peroxidase study. Neuroscience 6: 1315-1328, 1981.

Tsagarakis, S., Holly, J. M. P., Rees, L. H., Besser, G. M., and Grossman, A.: Acetylcholine and norepinephrine stimulate the release of corticotropin-releasing factor-41 from the rat hypothalamus in vitro. Endocrinology 123: 1962-1969, 1988.

Turkkan, J. S., Brady, J. V., and Harris, A. H. Animal studies of stressful interactions: A behavioral-physiological overview. In L. Goldberger and S. Bregnitz (Eds.), Handbook of Stress-Theoretical and Clinical Aspects. New York: Macmillan Publishing Co., 1982.

Ueda, H., Yasuda, H., Takabatake, Y., Tizuka, M., Tizuka, T., Ihori, M., Yammoto, M., and Sakamoto, Y.: Increased renin release evoked by mesencephalic stimulation in the dog. Jpn. Heart J. 8: 498-506, 1967.

Uhlich, E., Weber, P., Eigler, J., and Groschel-Stewart, U.: Angioten­sin-stimulated AVP release in humans. Klin. Wschr. 53: 177-180, 1975.

Ungerstedt, U.: Stereotaxic mapping of the monoamine pathways in the rat brain. Acta Physiol. Scand. 82(367): 1, 1971.

Urban, J. H., Van de Kar, L. D., Lorens, S. A., and Bethea, C. L.: Effect of the anxiolytic drug buspirone on prolactin and corticosterone secretion in stressed and unstressed rats. Pharmacology Biochemistry & Behavior 25(2): 457-462, 1986.

Uretsky, N. J. and Iversen, L. L.: Effects of 6-Hydroxydopamine on Catecholamine containing neurones in the rat brain. Journal of Neurochem­istry 17: 269-278, 1970.

Vale, W., Spiess, J., Rivier, C., and Rivier, J.: Characterization of a 41-residue ovine hypothalamic peptide that stimulates secretion of corticotropin and B-endorphin. Science 213: 1394, 1981.

Vale, W., Vaughan, J., Smith, M., Yamamoto, G., Rivier, J., and Rivier, C.: Effects of synthetic ovine corticotropin-releasing factor, glucocor­ticoids, catecholamines, neurohypophysial peptides, and other substances on cultured corticotropic cells. Endocrinology 113: 1121-1131, 1983.

Valenta, J. G. and Rigby, M. K.: Discrimination of the odor of stressed rats. Science 161: 599-601, 1968.

184

Van de Kar, L. D., Karteszi, M., Bethea, C. L., and Ganong, W. F.: Serotonergic stimulation of prolactin and corticosterone secretion is mediated by different pathways from the mediobasal hypothalamus. Neuroendocrinology 41: 380-384, 1985.

Van de Kar, L. D. and Lorens, S. A.: Differential serotonergic innervation of individual hypothalamic nuclei and other forebrain regions by the dorsal and median raphe nuclei. Brain Res 162: 45-54, 1979.

Van de Kar, L. D., Lorens, S. A., McWilliams, C., Kunimoto, K., Urban, J. H., and Bethea, C. L.: Role of midbrain raphe in stress-induced renin and prolactin secretion. Brain Res. 311: 333-341, 1984.

Van de Kar, L. D., Lorens, S. A., Urban, J. H., Richardson, K. D., Paris, J., and Bethea, C. L.: Pharmacological studies on stress-induced renin and pro lac tin secretion: effects of benzodiazepines, naloxone, propranolol and diisopropylfluorophosphate (DFP). Brain Res. 345: 257-263, 1985.

Van de Kar, L. D., Lorens, S. A., Vodraska, A., Allers, G., Green, M., Van Orden, D. E., and Van Orden, L. S.: Effect of selective midbrain and diencephalic 5, 7-dihydroxytryptamine lesions on serotonin content in individual preoptico-hypothalamic nuclei and on serum luteinizing hormone level. Neuroendocrinology 31: 309-315, 1980.

Van de Kar, L. D., Morton, K. D. R., and Gulati, A.: Fenfluramine-in­duced suppression of renin secretion is not mediated by peripheral catecholamines or by alpha-2 receptors. Fed. Proc. 47: 3876, 1988.

Van de Kar, L. D., Richardson, K. D., and Urban, J. H.: Serotonin and norepinephrine-dependent effects of fenfluramine on plasma renin activity in conscious rats. Neuropharmacol. 24: 487-494, 1985.

Van de Kar, L. D. and Richardson-Morton, K. D.: Serotonergic regulation of the release of renin is not mediated by the autonomic nervous system but involves beta adrenoceptors. Neuropharmacology 25(5): 487-492, 1986.

Van de Kar, L. D., Urban, J. H., Brownfield, M. S., and Simmons, W. H.: Partial characterization of a renin-releasing factor from plasma and hypothalamus. Hypertension 9(6): 598-606, 1987.

Van de Kar, L. D., Urban, J. H., Lorens, S. A., and Richardson, K. D.: The non-benzodiazepine anxiolytic buspirone inhibits stress-induced renin secretion and lowers heart rate. Life Sci. 36: 1149-1155, 1985.

Van de Kar, L. D., Wilkinson, C. W., and Ganong, W. F.: Pharmacological evidence for a role of brain serotonin in the maintenance of plasma renin activity in unanesthetized rats. J. Pharmacol. Exp. Ther. 219: 85-94, 1981.

185

Van de Kar, L. D., Wilkinson, C. W., Skrobik, Y., Brownfield, M. S., and Ganong, W. F.: Evidence that serotonergic neurons in the dorsal raphe nucleus exert a stimulatory effect on the secretion of renin but not of corticosterone. Brain Research 235: 233-243, 1982.

Van Hoesen, G. W., Wilson, L. M., Mac Dougall, J.M., and Mitchell, J. C.: Selective hippocampal complex deafferentation and de-efferentation and avoidance behavior in rats. Physiology and Behavior 8: 873-879, 1972.

Van Loon, G. R., Scapagnini, U., Moberg, G. P., and Ganong, W. F.: Evidence for central adrenergic neural inhibition of ACTH secretion in the rat. Endocrinology 89: 1464-1469, 1971.

Vander, A. J.: Effect of catecholamines and the renal nerves on renin secretion in anesthetized dogs. Am. J. Physiol. 209(3): 659-662, 1965.

Vander, A. J. and Geelhoed, G. W.: Inhibition of renin secretion by angiotensin II. Proc. Soc. Exp. Biol. Med. 120: 339-403, 1965.

VanderMaelen, C. P., Matheson, G. K., Wilderman, R. C., and Patterson, L. A.: Inhibition of serotonergic dorsal raphe neurons by systemic and iontophoretic administration of buspirone, a non-benzodiazepine anxiolytic drug. Eur. J. Pharmacol. 129: 123-130, 1986.

Vandesande, F. and Dierickx, K.: Identification of the vasopressin producing and of the oxytocin producing neurons in the hypothalamic neurosecretory system of the rat. Cell. Tiss. Res. 164: 153-162, 1975.

Vandongen, R., Peart, W. S., and Boyd, G. W.: Adrenergic stimulation of renin secretion in the isolated perfused kidney. Circulation Research 32: 290-296, 1973.

Velley, L., Mormede, P., and Kempf, E.: Neurochemical lesion of the nucleus locus coeruleus increases neophobia in a specific exploration task but does not modify endocrine response to moderate stress. Pharmacol. Biochem. Behav. 29: 1-7, 1988.

Vellucci, S. V.: The effects of ether stress of betamethasone treatment on the concentrations of noradrenaline and dopamine in various regions of the rat brain. Br. J. Pharmacol. 60: 601-605, 1977.

Vernikos-Danellis, J., Berger, P.A., and Barchas, J. D.: Brain serotonin and pituitary-adrenal function. Prog. Brain Res. 39: 301-309, 1973.

Vernikos-Danellis, J., Kellar, K. J., Kent, D., Gonzales, C., Berger, P. A., and Barchas, J. D.: Serotonin involvement in pituitary-adrenal function. Annals of the New York Academy of Sciences 297: 518-526, 1977.

Wallace, D. M., Magnuson, D. J., and Gray, T. S.: The amygdalo-brainstem pathway: selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat. Neuroscience Letters 97: 252-258, 1989.

186

Wanaka, A., Kiyama, H., Murakami, T., Matsumoto, M., Kamada, T., Malbon, C. C., and Tohyama, M.: Immunocytochemical localization of beta-adrenergic receptors in the rat brain. Brain Research 485: 125-140, 1989.

Watabe, T., Tanaka, K., Kumagae, M., Itoh, S., Hasegawa, M., Horiuchi, T., Miyabe, S., and Ohno, H.: Diurnal rhythm of plasma immunoreactive corticotropin-releasing factor in normal subjects. Life Sciences 40: 1651-1655, 1987.

Watkins, W. B. : Presence of adrenocorticotropin and beta-endorphin immunoreactivities in the magnocellular neurosecretory system of the rat hypothalamus. Cell. Tiss. Res. 207: 65-80, 1980.

Watson, J. B. Behaviorism. New York: Norton, 1924.

Watson, S. J., Akil, H., Fischli, W., Goldstein, A., Zimmerman, E., Nilaver, G., and van Wiersma Greidanus, T.: Dynorphin and vasopressin: Common localization in magnocellular neurons. Science 216: 85-87, 1982.

Weinberger, M. H., Aoi, W., and Henry, D. P.: Direct effects of beta-adrenergic stimulation on renin release by the rat kidney slice in vitro. Circ Res 37: 318-324, 1975.

Weiner, R. I. and Ganong, W. F.: Role of brain monoamines and histamine in regulation of anterior pituitary secretion. Physiol. Rev. 58: 905-975, 1978.

Weingarten, H. P., Chang, P., and McDonald, T. J.: Comparison of the metabolic and behavioral disturbances following paraventricular- and ventromedial-hypothalamic lesions. Brain Research Bulletin 14: 551-559, 1985.

Werka, T., Skar, J., and Ursin, H.: Exploration and avoidance in rats with lesions in amygdala and piriform cortex. J. Comp. Physiol. Psycho!. 92: 672-681, 1978.

Widmaier, E. P., Lim, A. T., and Vale, W.: Secretion of cor­ticotropin-releasing factor from cultured rat hypothalamic cells: Effects of catecholamines. Endocrinology 124: 583-590, 1989.

Wiegand, S. J. and Price, J. L.: The cells of origin of the afferent fibers to the median eminence in the rat. J. Comp. Neurol. 192: 1-19, 1980.

Wilczynski, E. A. and Osmond, D. H.: Plasma prorenin in humans and dogs species differences and further evidence of a systemic activation cascade. Hypertension 5: 277-285, 1983.

Wilczynski, E. A. and Osmond, D. H.: Extrarenal prorenin in plasma requires an activator of renal origin. Hypertension 8: Suppl II,85-88, 1986.

187

Wilczynski, E. A. and Osmond, D. H.: Evidence for beta-adrenergic regulation of renal and extrarenal plasma prorenin and renin in dogs. Proc. Soc. Exp. Bio.and Med. 182: 208-214, 1986.

Wilkinson, C. W., Shinsako, J., and Dallman, M. F.: Daily rhythms in adrenal responsiveness to adrenocorticotropin are determined primarily by the time of feeding in the rat. Endocrinology 104: 350-359, 1979.

Winn, P., Tarbuck, A., and Dunnett, S. B.: Ibotenic acid lesions of the lateral hypothalamus: comparison with the electrolytic lesion syndrome. Neuroscience 12: 225-240, 1984.

Wynn, P. C., Hauger, R. L., Holmes, M. C., Millan, M. A., Catt, K. J., and Aguilera, G.: Brain and pituitary receptors for corticotropin releasing factor: localization and differential regulation after adrenalectomy. Peptides 5: 1077-1084, 1984.

Yates, F. E. and Urkuhart, J. : Control of plasma concentration of adrenocortical hormones. Physiol. Rev. 42: 359-443, 1962.

Zanchetti, A. and Stella, A.: Neural control of renin release. Clin. Sci. Mol. Med. 48: 215-223, 1975.

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 Research 341: 101-109, 1985.

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 Research 358: 334-338, 1985.

Zimmerman, E. A., Stillman, M. A., Recht, L. D., Antunes, J. L., and Carmel, P. W.: Vasopressin and corticotropin releasing factor: an axonal pathway to portal capillaries in the zona externa of the median eminence containing vasopressin and its interaction with adrenal corticoids. Ann. N.Y. Acad. Sci. 297: 405-417, 1977.

Zimmerman, H. and Ganong, W. F.: Pharmacological evidence that stimulation of central serotonergic pathways increases renin secretion. Neuroendocrinology 30: 101-107, 1980.

APPROVAL SHEET

The dissertation submitted by Kathy Richardson Morton has been read and approved by the following committee:

Louis D. Van de Kar, Ph.D. Associate Professor Department of Pharmacology Loyola University Stritch School of Medicine

Mark S. Brownfield, Ph.D. Associate Professor Department of Comparative Biosciences University of Wisconsin School of Veterinary Medicine

Stephen B. Jones, Ph.D. Associate Professor Department of Physiology Loyola University Stritch School of Medicine

Alexander G. Karczmar, M.D., Ph.D. Professor Emeritus Department of Pharmacology Loyola University Stritch School of Medicine

Stanley A. Lorens, Ph.D. Professor Department of Pharmacology Loyola University Stritch School of Medicine

T. Celeste Napier, Ph.D. Assistant Professor Department of Pharmacology Loyola University Stritch School of Medicine

The final copies have been examined by the director of the dissertation and the signature which appears below verifies the fact that any necessary changes have been incorporated and that the dissertation is now given final approval by the Committee with reference to content and form.

The dissertation is therefore accepted requirements for the degree of Doctor of

in partial fulfillment Philosophy.

,/ h~ {,

Director's Signature

of the