© 2016 justin andrew smithufdcimages.uflib.ufl.edu/uf/e0/05/01/86/00001/smith_j.pdf · 1-1...

170
NEUROPEPTIDE PLASTICITY IN THE HYPOTHALAMIC PARAVENTRICULAR NUCLEUS OF SALT-LOADED MALE MUS MUSCULUS By JUSTIN ANDREW SMITH A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2016

Upload: others

Post on 15-Mar-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

NEUROPEPTIDE PLASTICITY IN THE HYPOTHALAMIC PARAVENTRICULAR

NUCLEUS OF SALT-LOADED MALE MUS MUSCULUS

By

JUSTIN ANDREW SMITH

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL

OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2016

Page 2: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

© 2016 Justin Andrew Smith

Page 3: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

To those that struggle and persevere

Page 4: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

4

ACKNOWLEDGMENTS

I would like to thank my mentor Dr. Eric Krause for providing the guidance and

resources necessary to perform this research. His drive and dedication to scientific investigation

are an inspiration. I would also like to thank my committee members Drs. Maureen Keller-

Wood, Joanna Peris, Jason Frazier and Debbie Scheuer, for their support during my graduate

training. I am also grateful to my lab mates Helmut Hiller, Lei Wang, and Dr. Annette de Kloet

who all helped with this project and many others. Finally, I am especially thankful for the love

and support of my mother Marcia, grandmother Virginia, brothers Shawn and Glenn, and long-

time friend Walter.

During my training I was supported by a Graduate Student Fellowship. This work was

funded in part by NIH grants HL096830 (EGK), HL122494 (EGK), and HL116074 (ADdK).

Page 5: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS ...............................................................................................................4

LIST OF TABLES ...........................................................................................................................8

LIST OF FIGURES .........................................................................................................................9

LIST OF ABBREVIATIONS ........................................................................................................11

ABSTRACT ...................................................................................................................................12

CHAPTER

1 INTRODUCTION ..................................................................................................................14

Paraventricular Nucleus of the Hypothalamus .......................................................................16

Central and Systemic Maintenance of Hydromineral Homeostasis .......................................21

Stress-Responsiveness During Osmotic Challenge ................................................................25

Chronic Salt-Loading ..............................................................................................................32

Aims ........................................................................................................................................39

Acute Salt-Loading ..........................................................................................................39

Chronic Salt-Loading ......................................................................................................40

2 MATERIALS AND METHODS ...........................................................................................46

Generation of CRH Reporter Mice .........................................................................................46

Validation of CRH Reporter ...................................................................................................46

Antibody Characterization ......................................................................................................47

Immunofluorescence ...............................................................................................................48

Image Capture and Analysis ...................................................................................................49

3 CENTRAL AND BEHAVIORAL EFFECTS OF ACUTE SALT-LOADING .....................53

Background .............................................................................................................................53

Materials and Methods ...........................................................................................................54

Animals ............................................................................................................................54

Restraint Stress and Blood Sampling ..............................................................................55

Immunohistochemistry ....................................................................................................55

Image Capture and Analysis ............................................................................................57

Behavioral Testing ...........................................................................................................57

Statistics ...........................................................................................................................58

Results .....................................................................................................................................58

Subcutaneous Delivery of 2.0 M NaCl Modestly Increases pNa+ but Blunts

Restraint-Induced Elevations in Plasma CORT ...........................................................58

Page 6: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

6

Acute Hypernatremia Inversely Affects Restraint-Induced Activation of CRH and

Oxytocin Neurons in the PVN .....................................................................................59

Slight Elevations in the pNa+ are Associated with Decreased Anxiety-Like

Behavior in the EPM ....................................................................................................59

Discussion ...............................................................................................................................59

4 CENTRAL AND BEHAVIORAL EFFECTS OF CHRONIC SALT-LOADING ................69

Background .............................................................................................................................69

Methods ..................................................................................................................................70

Animals ............................................................................................................................70

Procedure .........................................................................................................................70

Restraint Stress and CORT ..............................................................................................71

Image Capture and Analysis ............................................................................................72

Statistics ...........................................................................................................................73

Results .....................................................................................................................................73

Salt-Loaded Mice Exhibit Increased Fluid Intakes and pNa+ Concentrations, but

Maintain Normal Indices of Blood Volume ................................................................73

Salt-Loaded Mice Maintain Food Intake and Body Weight ...........................................74

Salt-Loading Does Not Increase Neuronal Activation in PVN Vasopressin Neurons

or the SON ...................................................................................................................74

Salt-Loading Increases Fos Expression in the Posterior PVN ........................................74

Salt-Loading does not Affect Restraint-Induced Fos Induction in tdTomato

Neurons ........................................................................................................................75

Salt-Loading Blunts Restraint-Induced Elevations in Plasma CORT During

Recovery ......................................................................................................................75

Discussion ...............................................................................................................................75

5 CHRONIC SALT-LOADING-ASSOCIATED NEUROPEPTIDE PLASTICITY ...............86

Background .............................................................................................................................86

Methods ..................................................................................................................................88

Animals ............................................................................................................................88

In situ Hybridization ........................................................................................................88

Image Capture and Analysis ............................................................................................89

Statistics ...........................................................................................................................92

Results .....................................................................................................................................92

Salt-Loading Reduces CRH mRNA Expression in the PVN ..........................................92

Fluorescent Tracking of CRH mRNA by tdTomato .......................................................93

Total tdTomato Unaffected by Salt-Loading or Restraint ...............................................93

Salt-Loading is not Associated with Alterations in the Extent of Oxytocin-

NP/tdTomato Colocalizations in the PVN ...................................................................94

Salt-Loading is Associated with an Increase in Vasopressin-NP/tdTomato

Colocalizations in the PVN ..........................................................................................95

Salt-Loading does not Affect CRH Expression in Vasopressin-NP or Oxytocin-NP

Expressing Neurons of the SON ..................................................................................95

Page 7: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

7

Fluorogold Injections were Localized to the Nucleus Ambiguous, Pre-Bötzinger

Complex, and Rostroventrolateral Medulla .................................................................96

Retrogradely-Labeled Neurons Increase Expression of tdTomato and Vasopressin-

NP in PVN ...................................................................................................................97

Retrogradely-Labeled Neurons in the Amygdala are Distinct from tdTomato-

Expressing Neurons .....................................................................................................98

Discussion ...............................................................................................................................98

6 GENERAL DISCUSSION AND FUTURE DIRECTIONS ................................................119

Closing Remarks ...................................................................................................................137

Future Directions ..................................................................................................................140

LIST OF REFERENCES .............................................................................................................144

BIOGRAPHICAL SKETCH .......................................................................................................170

Page 8: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

8

LIST OF TABLES

Table page

2-1 Antibody information.........................................................................................................51

Page 9: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

9

LIST OF FIGURES

Figure page

1-1 Schematic of the major systems that interact with neurons in the paraventricular

nucleus to maintain hydromineral homeostasis .................................................................43

1-2 Schematic of CRH pathways projecting from the PVN ....................................................44

1-3 Osmosensitive and stress-responsive circuits ....................................................................45

2-1 Representative images validating CRH-Cre induced tdTomato expression. .....................52

3-1 Plasma measurements from mice 60 min after 2.0 M or 0.15 M NaCl injections .............64

3-2 2.0 M NaCl administration reduces the CORT response to restraint stress relative to

0.15 M NaCl.......................................................................................................................65

3-3 2.0 M NaCl injection attenuates restraint-induced activation of CRH neurons ................66

3-4 2.0 M NaCl administration and restraint increases activation of oxytocin-producing

cells in the PVN .................................................................................................................67

3-5 Acute hypernatremia attenuates anxiety-like behavior ......................................................68

4-1 Fluid consumption and the effects on indices of hydration ...............................................80

4-2 Individual fluid and food intake values .............................................................................81

4-3 Correlation graphs for food and fluid intake ......................................................................82

4-4 Daily food intake and body mass measurements ...............................................................83

4-5 Stress induced neuronal activation .....................................................................................84

4-6 Chronic salt-loading reduces the CORT response to restraint ...........................................85

5-1 Representative images of PVN sections processed for in situ hybridization of

digoxigenin-conjugated probe and CRH mRNA .............................................................104

5-2 Analysis of tdTomato reporting for CRH mRNA expression in the PVN ......................105

5-3 Analysis of total tdTomato in the PVN ............................................................................106

5-4 Representative unilateral images and quantification of oxytocin-NP/tdTomato

colocalizations in the “neurosecretory” region of the PVN .............................................107

5-5 Representative unilateral images and quantification of oxytocin-NP/tdTomato

colocalizations in the “preautonomic” region of the PVN ...............................................108

Page 10: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

10

5-6 Representative unilateral images and quantification of vasopressin-NP/tdTomato

colocalizations in the “neurosecretory” region of the PVN .............................................109

5-7 Representative unilateral images and quantification of vasopressin-NP/tdTomato

colocalizations in the “preautonomic” region of the PVN ...............................................110

5-8 Average total colocalizations in the PVN ........................................................................111

5-9 Representative unilateral images used for quantification of colocalizations in the

supraoptic nucleus ............................................................................................................112

5-10 A summary of FG injections targeting the RVLM ..........................................................113

5-11 Representative images and quantification of PVN neurons labeled with FG. .................115

5-12 Representative images and quantification of FG/vasopressin-NP colocalization ...........116

5-13 Representative images and quantification of FG/tdTomato colocalization .....................117

5-14 Representative image of FG and tdTomato in the amygdala ...........................................118

6-1 Increased vesicular glutamate transporter 2 appositions on tdTomato neurons in the

PVN with salt-loading......................................................................................................142

6-2 Decreased labeling for the astrocytic marker glial fibrillary acidic protein with salt-

loading..............................................................................................................................143

Page 11: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

11

LIST OF ABBREVIATIONS

ACTH Adrenocorticotrophin hormone

ASD Autism spectrum disorder

AT1 Angiotensin type 1

BNST Bed nucleus of the stria terminalis

CORT Corticosterone

CRH Corticotrophin releasing hormone

DAPI 4',6-Diamidino-2-phenylindole, dihydrochloride

EDTA Ethylenediaminetetraacetic acid

FG Fluorogold

HSD Hydroxysteroid dehydrogenase

HPA Hypothalamus-pituitary-adrenal

mRNA Messenger ribonucleic acid

NaCl Sodium chloride

NP Neurophysin

OVLT Organum vasculosum of the lamina terminalis

PVN Paraventricular nucleus of the hypothalamus

pNa+ Plasma sodium

RAAS Renin-angiotensin-aldosterone system

RVLM Rostral ventrolateral medulla

SEM Standard error of the mean

SFO Subfornical organ

SON Supraoptic nucleus

Page 12: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

12

Abstract of Dissertation Presented to the Graduate School

of the University of Florida

Requirements for the Degree of Doctor of Philosophy

NEUROPEPTIDE PLASTICITY IN THE HYPOTHALAMIC PARAVENTRICULAR

NUCLEUS OF SALT-LOADED MALE MUS MUSCULUS

By

Justin Andrew Smith

August 2016

Chair: Eric G. Krause

Major: Pharmaceutical Sciences – Pharmacodynamics

The paraventricular nucleus of the hypothalamus (PVN) orchestrates neurohumoral

responses to dehydration that maintain body fluid homeostasis through the coordinated actions of

neuroendocrine and preautonomic neurons producing corticotrophin-releasing hormone (CRH),

vasopressin, and oxytocin. To evaluate the neuropeptide plasticity and neuronal activation

within the PVN resulting from acute and chronic salt-loading, we analyzed the central expression

of red fluorescent protein (tdTomato) driven by CRH transcription together with

immunohistochemistry. Acute salt-loading was accomplished via subcutaneous 2.0 M NaCl

injections while chronic salt-loading took place by replacing drinking water with 2% NaCl for 5

days. Acute salt-loading resulted in an attenuated neuroendocrine response to a psychogenic

stressor, as both restraint-induced Fos in tdTomato neurons of the PVN and plasma

corticosterone were reduced compared with controls. Moreover, acute salt-loading increased the

number of activated neurons producing oxytocin in the PVN and decreased anxiety-like behavior

assessed on the elevated plus maze. Chronic salt-loaded and control mice exhibited similar Fos

induction within vasopressin neurons and in response to restraint; however, chronic salt-loading

was associated with a blunted corticosterone response to restraint, but an increased activation in

the posterior preautonomic region of the PVN. While the number of oxytocin and tdTomato

Page 13: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

13

colocalizations did not differ between groups, chronic salt-loading did significantly increase the

colocalization of vasopressin and tdTomato in putative magnocellular PVN neurons. To

evaluate neuropeptide expression in preautonomic neurons, mice were injected with the

retrograde tracer Fluorogold targeting the rostroventrolateral medulla and then subjected to salt-

loading. Interestingly, salt-loading significantly increased the colocalization of vasopressin and

Fluorogold as well as tdTomato and Fluorogold in PVN neurons. These results indicate that the

effects of acute and chronic salt-loading include central effects mediated by diverse neuronal

phenotypes in the PVN that likely act in concert to maintain homeostasis as well as influence

behavior and stress-responsiveness.

Page 14: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

14

CHAPTER 1

INTRODUCTION

Maintenance of hydromineral homeostasis is a biological imperative observed across all

mammalian species representing fundamental and well conserved homeostatic mechanisms

(Mecawi Ade et al., 2015). These mechanisms ensure adequate hydration as a result of a

complex set of endocrine, neural, and behavior-based interactions with the external environment

that are well suited to experimental manipulation (Bernard, 1949). Experimental paradigms

involving exposure to water deprivation or excessive sodium chloride serve to provide a better

understanding of normal and pathological functions guarding against hypertonicity. The

extensive investigation of these functions has led to the better understanding of multiple central

and systemic signals that converge on the hypothalamus, are integrated there, and produce well-

characterized responses (Stellar et al., 1954; Teitelbaum and Stellar, 1954). Specifically,

challenges to osmoregulation include responses governed by specialized neurons of the

paraventricular nucleus of the hypothalamus (PVN) which, as shown in Figure 1-1, coordinate

neuroendocrine secretion of vasopressin and oxytocin (Bourque, 2008), stress-responsiveness of

the hypothalamic-pituitary-adrenal (HPA) axis (Roberts et al., 2011), aspects of the renin-

angiotensin-aldosterone system (RAAS) (Hwang et al., 1986), and neural control of autonomic

renal and cardiovascular regulation (Stocker et al., 2004; Bardgett et al., 2014). Experiments that

challenge hydration and sodium handling have been especially fruitful in a somewhat isolated

sense; that is, osmoregulatory responses controlled by neurons in the PVN can be induced by

manipulating water or salt intake with reliable results. However, discovering how the individual

subtypes of neurons in the PVN coordinate, interact, and alter their transcriptional products and

signaling modalities in relation to one another requires assessment of multiple systems.

Unraveling the complexities of these relationships may lead to a better understanding of affective

Page 15: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

15

and cardiovascular disorders involving the PVN as well as the frequent comorbidity of these

disorders with dysregulated systems that normally serve to regulate hydromineral balance.

The RAAS and the neuropeptides oxytocin and vasopressin have physiological functions

that are coupled with stress-responsiveness and behavior. The physiological stressors of

hypotension and hypovolemia stimulate renin release to catalyze angiotensin-II formation and

aldosterone availability (Fitzsimons, 1998), as well as vasopressin release (Yao et al., 2011).

These hormones promote fluid retention to stabilize blood volume and pressure (Fitzsimons,

1998), but also act in the brain to modulate behavior and potentiate stress-responsiveness

(Krause et al., 2011b) through direct and indirect action in brain regions that innervate the PVN

(Englemann et al., 2000; Albrecht, 2010). Low pNa+ is alleviated by eliminating excess water

through diuresis and vasopressin inhibition (Wang et al., 2013b) as well as by increasing salt

intake and reducing salt excretion. Hypernatremia increases the pituitary release of vasopressin

and oxytocin to cause renal water reabsorption and salt excretion (Masaharu and Hiraki, 2013),

but the central release of these peptides is known to affect distant forebrain and hindbrain targets

(Antunes-Rodrigues et al., 2013) that also influence anxiety (Knobloch et al., 2012; Neumann

and Landgraf, 2012b; Benarroch, 2013b; Frazier et al., 2013). Together these signals represent a

constantly fluctuating background upon which the stress response may be potentiated or

inhibited depending on the magnitude, polarity, and duration of an osmotic challenge.

The neuroanatomical pathways and interactions between PVN neuronal subtypes that

govern stress-responsiveness, autonomic function, and adequate hydration are complex and

represent ongoing lines of research (Antunes-Rodrigues et al., 2013). While many studies have

focused on these interactions separately, a cohesive picture of the specific central pathways that

are modulated on a circuit, cellular, and molecular basis is lacking. The pathways outlined in

Page 16: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

16

Figure 1-1 represent the major systems that are potently activated during an osmotic challenge.

To provide a foundational understanding of these systems and how they are controlled, an

introduction to PVN anatomy and physiology, systemic systems that interact with the PVN, and

the salt-loading methods that are the basis of this dissertation are discussed.

Paraventricular Nucleus of the Hypothalamus

Neurons in general can be classified according to function by determining connectivity

and the types of signaling molecules they utilize, and neurons in the PVN are especially

amenable to this classification. The PVN is a bilateral nucleus situated on either side of the

anterior or rostral portion of the third ventricle (Swanson and Sawchenko, 1983). Its location at

the base of the brain is ideal for efficient connectivity with the pituitary gland, which acts as an

intermediary between the convergence of neuronal input to the hypothalamus and the dispersive

property of the blood (Swanson and Kuypers, 1980; Swanson and Sawchenko, 1980). The

neurons in the PVN that influence the pituitary are further divided into neurons that directly

release neuropeptide into the systemic circulation via the posterior pituitary (Vandesande and

Dierickx, 1975) and those that release into the portal circulation to act on neurons of the anterior

pituitary which then dispense their contents systemically (Vale et al., 1977; Vale et al., 1981;

Rivier et al., 1982; Vale et al., 1986). Neurons that project to the posterior pituitary are of two

major types, those that produce vasopressin and those that produce oxytocin (Poulain and

Wakerley, 1982); both of which are classified as magnocellular because of their large soma

compared with other the types of cells in the PVN. Magnocellular neurons are generally grouped

together although the degree to which they are segregated from other types of neurons in the

PVN is species-specific (Biag et al., 2012). Notably, the magnocellular population of neurons in

the rat is highly segregated while magnocellular neurons in the mouse are much more diffuse -

an important subject that will be revisited several times in this text. Typical grouping of

Page 17: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

17

magnocellular neurons in rats is found in the dorsal and lateral aspects of the PVN, with axonal

projections extending laterally around the fornix and then ventrally before coursing medially

towards and through the internal zone of the median eminence (Swanson and Kuypers, 1980;

Swanson and Sawchenko, 1980; 1983). Projections of magnocellular neurons are not limited to

the pituitary, as axonal collaterals and neurons with dedicated projections innervate various areas

of the brain (Sofroniew, 1980; Knobloch et al., 2012). Additionally, magnocellular vasopressin

and oxytocin neurons are found in the second highest concentration within the supraoptic

nucleus, a bilateral nucleus located just above the laterodorsal aspects of the optic chiasm and

optic tracts (Swanson and Sawchenko, 1983). SON magnocellular neurons project to the

pituitary and are intimately involved in hydromineral balance (Swanson and Sawchenko, 1983;

Bourque, 2008). Magnocellular dendrites are capable of bi-directional communication as in

addition to postsynaptic reception, dendrites have the additional ability to release neuropeptide

into the extracellular fluid to signal presynaptically and via passive diffusion (Ludwig and Leng,

2006).

In contrast to magnocellular neurons, neurons that signal to the anterior pituitary do so

via release into the hypothalamic-hypophyseal portal circulation along the external zone of the

median eminence (Akmayev, 1971; Akmayev and Popov, 1977). These neurons are classified as

parvocellular due to a smaller and more evenly contoured soma profile. Parvocellular neurons

contain and release a variety of neuropeptides that signal corresponding release (or inhibition of

the release) of paired hormones (Schally et al., 1968; Schally et al., 1973; Schally, 1978). The

primary parvocellular neuroendocrine hormones with their respective systemically-released

hormones are: corticotrophin-releasing hormone (CRH) and adrenocorticotropic hormone

(ACTH) (Vale et al., 1981); thyrotropin-releasing hormone and thyroid-stimulating hormone

Page 18: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

18

(Boler et al., 1969; Burgus et al., 1970); gonadotropin-releasing hormone/gonadotropin-

inhibiting hormone and luteinizing hormone or follicle-stimulating hormone (Clarke and

Cummins, 1982; Vale et al., 1986); prolactin-releasing peptide/prolactin-inhibiting factor and

prolactin (Freeman et al., 2000); as well as somatocrinin/somatostatin and growth hormone

(Burgus et al., 1973; Brazeau et al., 1982). Similar to magnocellular neurons, parvocellular

neurons individually project to regions of the brain other than the median eminence (Wittmann et

al., 2009). Furthermore, both magnocellular and parvocellular neuropeptides are found in

neurons residing in areas other than the PVN (Valentino et al., 1995; Rodaros et al., 2007; Guo et

al., 2013; Bosch et al., 2016), suggesting that each systemic function that is controlled by a

particular neuropeptide likely has a series of circuits in the brain that operates via similar

receptors; however, these coincidental actions are still an active area of research. Indeed, the

overall projection map of PVN neurons implicates that many central functions might be

coordinated precisely with the release of neuropeptide and suggests that the naming of certain

neuropeptides as “releasing hormones” or “releasing factors” may be unfortunate as it describes

only a fraction of what they do. In this sense, PVN neurons project directly to a multitude of

hindbrain (Geerling et al., 2010), midbrain (Rodaros et al., 2007), and forebrain targets

(Rivalland et al., 2006; Tilbrook et al., 2006) that mediate diverse functions including ingestive

behavior (Kirchgessner et al., 1988), sexual function (Tang and McKenna, 2001; Bancila et al.,

2002), anxiety-like behavior (Busnardo et al., 2013), memory processes (Wiedenmayer et al.,

2005; Lucas et al., 2013), as well as autonomic functions related to body temperature (Menendez

et al., 1990; Madden and Morrison, 2009) and cardiovascular (Loewy and McKellar, 1980;

Byrum and Guyenet, 1987; Coote, 1995; Benarroch, 2005; Coote, 2005; Li et al., 2006; Ferguson

et al., 2008; Kawabe et al., 2008; Pyner, 2009) regulation. Especially pertinent here, PVN

Page 19: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

19

connectivity includes hindbrain regions controlling and interacting with cardiovascular and renal

circuits that promote normal maintenance of water and salt handling (Guyenet, 2006; Bourque,

2008). These “preautonomic” neurons are largely grouped in the posterior or caudal region of

the PVN although there is a small degree of overlap with pituitary-projecting or “neurosecretory”

regions (Biag et al., 2012). Thus, the anatomy of the PVN is indicative of a diverse population

of neurons; however, each has a potential for increased complexity above and beyond basic

connectivity and primary phenotypical marker, especially when considering the functional

implications of colocalized neuropeptide.

The anatomical and neuropeptide designations outlined above for neurons in the PVN are

important for an overall understanding of major functions controlled by the PVN, but are also

oversimplified as multiple neuropeptides have been localized to each neuronal subtype. It should

be noted that identification of a particular neuronal substance is dependent on the validity of the

method used. Magnocellular vasopressin-producing neurons additionally express galanin

(Melander et al., 1986a; Melander et al., 1986b; Rokaeus et al., 1988; Gaymann and Martin,

1989), dynorphin (Watson et al., 1982; Whitnall et al., 1983), [Leu]enkephalin (Martin and

Voigt, 1981; Martin et al., 1983), and angiotensin-II (Kilcoyne et al., 1980). Similarly,

magnocellular oxytocin-producing neurons colocalize with markers for cholecystokinin

(Beinfeld et al., 1980; Vanderhaeghen et al., 1981; Martin et al., 1983), galanin (Gaymann and

Martin, 1989), thyrotropin-releasing hormone (Tsuruo et al., 1988), renin (Fuxe et al., 1982),

and [Met]enkephalin (Martin et al., 1983; Vanderhaeghen et al., 1983; Adachi et al., 1985).

Furthermore, magnocellular vasopressin and oxytocin neurons in both the PVN and SON

colocalize with tyrosine hydroxylase (Swanson et al., 1981; Chan-Palay et al., 1984; Spencer et

al., 1985) and magnocellular oxytocin neurons in these same regions colocalized with CRH.

Page 20: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

20

There is less diversity of colocalized neuropeptides in the parvocellular neurons of the PVN

compared with the magnocellular neurons, although the functionality of parvocellular

colocalization of CRH and vasopressin is perhaps better understood.

The PVN contains a heterogeneous population of neurons that respond to stressful stimuli

in part through activation of parvocellular cells producing CRH, vasopressin, and a subset of

cells containing both CRH and vasopressin (Lightman, 2008). A combination of steroid and

neurotransmitter influences relay information related to stressor type, novelty, and magnitude to

these neurons to induce corticotroph production and ACTH release (Armario, 2006). Several

factors are known to stimulate ACTH release from the anterior pituitary (Abousamra et al.,

1987). These factors include: oxytocin (Makara et al., 1996), norepinephrine and epinephrine

(Rivier and Vale, 1983; Labrie et al., 1984), and angiotensin-II (Gaillard et al., 1981) as well as

the better studied and more potently acting CRH and vasopressin (Gillies et al., 1982). While

CRH is the dominant ACTH secretagogue (Vale et al., 1981), CRH and vasopressin synergize to

greatly potentiate the release of ACTH (Gillies et al., 1982), thus attributing special interest to

neurons that colocalize these neuropeptides. CRH and/or vasopressin release from parvocellular

neurons initiates activation of the HPA axis (Arborelius et al., 1998; Jacobson, 2014) which

normally serves to mobilize energy reserves through adrenal glucocorticoid release in

cooperation with the autonomic nervous system (Ulrich-Lai and Herman, 2009). The efficiency

of the stress response is governed by 1) the ability to successfully integrate multiple lines of

central and peripheral information to induce a response (Ziegler and Herman, 2002) and 2) the

behavioral adaptation and glucocorticoid feedback mechanisms responsible for effecting a timely

resolution (Keller-Wood and Dallman, 1984). Moreover, CRH and vasopressin act as

neurotransmitters and neuromodulators in various regions of the brain to coordinate behavioral

Page 21: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

21

and autonomic stress responses (Dunn and Berridge, 1990), and in this sense can be localized to

forebrain- and hindbrain-projecting neurons of the PVN as well (Fig. 1-2).

Central and Systemic Maintenance of Hydromineral Homeostasis

The RAAS plays a major role in the regulation of systemic blood pressure, urinary

sodium excretion, and renal hemodynamics. In experimental models, stimuli such as

pharmacologically-induced vasodilation (Stocker et al., 2000), hypovolemia induced by

polyethylene glycol administration (Stricker and Verbalis, 1986), or hemorrhage (Ken'ichi and

Hitoshi, 2010) are potent activators of the RAAS and elicit thirst. Furthermore, volume

depletion associated with sodium loss provokes a rise in angiotensin-II and aldosterone levels

(Badaue-Passos et al., 2007). These conditions will be briefly addressed followed by an account

of their relevance to the potentiation of stress-responsiveness.

Blood pressure is intrinsically bound to hydromineral balance and deficiency compels

several homeostatic mechanisms to action. Carotid sinus and aortic baroreceptors decrease tonic

firing frequency in response to a drop in pressure, activating neurons in the nucleus of the

solitary tract that increase sympathetic outflow while decreasing parasympathetic tone (Fisher

and Paton, 2012). If the drop in pressure is severe, afferent signals from the nucleus of the

solitary tract and forebrain induce vasopressin release which acts on vasopressin type 1 receptors

of the vasculature (Aoyagi et al., 2009) and vasopressin type 2 receptors in the kidney (Sampey

et al., 1999) to promote vasoconstriction and water reabsorption, respectively. Cardiopulmonary

baroreceptors also activate nucleus of the solitary tract neurons that increase sympathetic nervous

system activity. These neurons cause rostroventrolateral medulla and intermediolateral cell

column efferents to release norepinephrine onto β1 receptors of the juxtaglomerular apparatus,

thereby increasing renin release and plasma renin activity to promote the synthesis of

angiotensin-II (Hainsworth, 2014). Additionally, reduced perfusion pressure is sensed by

Page 22: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

22

baroreceptors in the renal vasculature to increase plasma renin activity via prostaglandin

signaling (Schweda and Kurtz, 2010). Prostaglandin-mediated renin release is furthermore

induced by a reduction of NaCl circulation in the renal medulla (Schweda and Kurtz, 2010).

Thus, reduced extracellular fluid volume is monitored by the systemic vasculature and kidneys to

initiate activation of the RAAS as well as hypothalamic nuclei controlling endocrine release and

autonomic balance.

Increased circulating angiotensin-II affects both systemic and centrally-mediated

compensatory measures. Angiotensin-II binding to angiotensin type 1 (AT1) receptors is a

potent vasoconstrictive stimulus (Fitzsimons, 1998), and activation of AT1 receptors expressed

on the efferent arteriole of the kidney increases glomerular filtration rate (Hiroyuki et al., 2013)

during instances of low blood pressure. Furthermore, elevated levels of angiotensin-II

upregulate the synthesis and release of aldosterone in the zona glomerulosa of the adrenal gland

(Hattangady et al., 2012). Aldosterone binds to cytosolic mineralocorticoid receptors within

renal tubular cells of the cortical collecting ducts, increasing the reabsorption of sodium and

water (Biller et al., 2000). The effects of angiotensin-II in the periphery restore extracellular

fluid volume by conserving water and sodium, allowing normal function of vital organs.

Accessibility of blood-borne angiotensin-II to the brain however, is blocked by the blood-brain

barrier, except for in regions that have a specialized blood-brain barrier architecture (McKinley

et al., 2002). These regions, termed circumventricular organs, allow the brain to sample the

immediate systemic environment through specialized neurons (Duvernoy and Risold, 2007). In

particular, the subfornical organ (SFO) and organum vasculosum of the lamina terminalis

(OVLT) are circumventricular organs that express receptors for angiotensin-II and aldosterone

(De Luca et al., 2013). Activation of AT1 receptors in the core of the SFO and lateral margins of

Page 23: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

23

the OVLT by angiotensin-II (Sunn et al., 2003) stimulates a neuronal network with nodes in

several brain regions.

The net result of angiotensin-II induced activation of this network is the enigmatic

behavioral and physiological phenomenon of thirst (Stricker and Sved, 2000). Thirst, which is

largely dependent on circumventricular organ afferent projections to the median preoptic nucleus

(Hollis et al., 2008), as well as complex regulation of neuropeptide release (McKinley et al.,

2004) and sympathetic activation (Stocker et al., 2008) designed to complement systemic

function. Moreover, it is now well-established that the brain is capable of producing all the

components of the RAAS (Phillips and Sumners, 1998) (Gomez-Sanchez et al., 2005), and

systemic angiotensin-II and aldosterone elevations may influence the activity of the local brain

RAAS to modulate limbic and neuroendocrine systems.

In regards to aldosterone, its central actions are not restricted by the blood-brain barrier,

but instead, are limited by competition with glucocorticoids for binding sites (Geerling and

Loewy, 2009). Because aldosterone circulates at concentrations ~1000 fold less than

glucocorticoids (Bledsoe et al., 2005), most mineralocorticoid receptors in the brain are not

activated by aldosterone. Exceptions include the PVN (Chen et al., 2014); amygdala and locus

coeruleus (Robson et al., 1998); and nucleus of the solitary tract (Geerling and Loewy, 2006),

where specialized subsets of cells express both mineralocorticoid receptors and 11β-

hydroxysteroid dehydrogenase (HSD), suggesting that these areas are susceptible to

physiological states that increase circulating angiotensin-II and aldosterone. However,

controversy exists over the ligand for the mineralocorticoid receptors in the PVN. 11β-HSD is

expressed in two isoforms: 11β-HSD1 reduces CORT in humans to cortisol which allows

activation of glucocorticoid receptors; while 11β-HSD2 oxidizes cortisol to prevent activation of

Page 24: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

24

mineralocorticoid receptors. Chen et al. found that 11β-HSD1, but not 11β-HSD2 was expressed

in the caudal PVN, and that mineralocorticoid receptors were present and likely activated with

sodium deprivation to increase sympathetic output (Chen et al., 2014). Haque et al. found that

11β-HSD2 was present with mineralocorticoid receptors in magnocellular neurons of the PVN

and SON, but did not provide any functional data (Haque et al., 2015). Knowledge of the ligand

for mineralocorticoid receptors in the PVN and SON is important to understand the physiological

relevance of steroid signaling to the individual types of neurons in the PVN, especially during an

osmotic perturbation.

The acute elevation of plasma tonicity that occurs following hypertonic saline

administration creates an osmotic gradient which moves water from the intracellular fluid

compartment to the extracellular fluid, which in turn, increases blood volume and pressure

(Caeiro and Vivas, 2008). The loss of water from the intracellular fluid compartment causes cell

shrinkage that is thought to activate osmoreceptors located on the membrane of specialized

osmosensitive cells. Research investigating osmoreceptor function supports the notion that

changes in membrane tension elicit conformational changes in the osmoreceptor that lead to

increased membrane ion permeability, thereby transducing the change in osmolality into a

signaling mechanism that can affect other cells (Voisin and Bourque, 2002; Bourque, 2008).

Osmosensitive cells are located in strategic areas including the throat (Kuramochi and

Kobayashi, 2000), intestines (Dooley and Valenzuela, 1984), hepatic portal vein (Baertschi and

Vallet, 1981) and circumventricular organs in the brain (McKinley et al., 1999). In regards to

circumventricular organs, the OVLT conveys information regarding a change in osmolality

directly to vasopressin and oxytocin producing cells in the SON and PVN (Johnson and Gross,

1993). As discussed previously, the release of vasopressin into the systemic circulation counters

Page 25: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

25

the increased plasma osmolality by promoting renal water retention. In rodents, the elevations in

systemic oxytocin that follows hypernatremia promotes the inhibition of renin release and

stimulates renal sodium excretion (Sjoquist et al., 1999). Thus, increased plasma osmolality due

to excess body sodium elicits activation of osmosensitive cells causing subsequent release of

neuropeptides into the systemic circulation that alter the renal handling of sodium and water to

alleviate hypernatremia.

In summary, neuropeptide-releasing PVN neurons and the RAAS are potently activated

by altered blood pressure, volume and tonicity. Through multiple sensory mechanisms, body

fluid homeostasis is maintained through cardiovascular and renal signals that converge on the

hypothalamus through circumventricular organ and hindbrain inputs, which in turn, are

reciprocally connected to limbic brain regions controlling mood and affect (Fig. 1-3). Thus,

brain regions controlling mood and affect promote cardiovascular and endocrine responses to

stress through hypothalamic and hindbrain nuclei. Consequently, overlapping neuronal networks

are positioned to influence cardiovascular and endocrine responses to hydromineral imbalance as

well as psychological stressors (Fig. 1-3).

Stress-Responsiveness During Osmotic Challenge

Stressors can be classified into the two broad categories of physiological or

psychological. Physiological stressors include those that threaten body fluid homeostasis, and

activation of the RAAS is an example of the body’s attempt to cope through endocrine,

autonomic, and behavioral adjustments. Stressors that are psychogenic in nature influence

anxiety level, sympathetic tone, and the HPA axis through anticipatory, memory, and fear-related

systems relayed to the hypothalamus via limbic brain regions (Jankord and Herman, 2008).

Although often unnecessary to the welfare of the individual, coping mechanisms that respond to

psychogenic stressors can include RAAS activation (Krause et al., 2011b). Both types of

Page 26: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

26

stressors have the potential to activate the HPA axis via stimulation of PVN parvocellular CRH

neurons that project to the median eminence. As discussed above, the release of CRH into the

primary capillary plexus of the hypothalamo-hypophyseal portal vasculature induces anterior

pituitary ACTH release. The binding of ACTH to the zona fasciculata layer of the adrenal gland

causes glucocorticoid synthesis and release into the systemic circulation that bind to

glucocorticoid receptors throughout the body to mobilize energy reserves and down-regulate

processes not essential to the stress response. Negative feedback of glucocorticoid occurs at

several levels including the pituitary and PVN to bring about termination of HPA axis activity.

Exposure to psychogenic stress activates the RAAS in the absence of threats to

hydromineral balance to increase central and systemic angiotensin-II (Yang et al., 1993; Krause

et al., 2011b) as well as the expression of AT1 receptors in the brain (Aguilera et al., 1995a).

Stress-induced up-regulation of AT1 receptor mRNA occurs in the SFO, anterior pituitary, zona

glomerulosa and medulla of the adrenal gland (Leong et al., 2002), and on CRH cells in the PVN

(Aguilera et al., 1995b). Similarly, administration of dexamethasone, a glucocorticoid receptor

agonist, significantly increases AT1 receptor binding in the PVN and SFO (Shelat et al., 1999).

Taken together, these results suggest that glucocorticoids derived from stress-induced HPA axis

activation may feed forward to increase AT1 receptor expression to alter sensitivity to centrally

released angiotensin-II. Consistent with this notion, pretreatment with dexamethasone augments

water drinking stimulated by central administration of angiotensin-II (Ganesan and Sumners,

1989). Elevated glucocorticoids accompanied by increased angiotensin-II release in the PVN

reduces 11β-HSD effectiveness resulting in increased preautonomic stimulation (Gabor and

Leenen, 2012) and potentiation of sympathetic activity (de Kloet et al., 2005). Thus, induction

Page 27: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

27

of the HPA axis may interact with an activated RAAS to sensitize subsequent behavioral and

autonomic responses to hydromineral imbalance.

The RAAS has gained considerable attention over the past decade as a potential

therapeutic target for treating stress-related disorders, anxiety, depression, alcoholism, cognitive

impairment as well as neuroinflammatory and neurodegenerative diseases (Saavedra and Pavel,

2005; Albrecht, 2010; Saavedra, 2012; Anil Kumar et al., 2014). Angiotensin receptor blockers

administered either peripherally or centrally, improve mood and cognition as well as attenuate

HPA axis and sympathetic responses to stressors having strong psychogenic components

(Saavedra and Pavel, 2005; Albrecht, 2010; Saavedra, 2012; Anil Kumar et al., 2014). While the

specific circuits that angiotensin-II receptor antagonists modulate to achieve these affects are not

fully understood, increased circulating angiotensin-II is known to influence hypothalamic

function through various pathways including those originating in circumventricular organs and

regions governing mood and affect.

The SFO and OVLT, together with the median preoptic nucleus represent key regulatory

areas that both share connectivity with one another and project to the hypothalamus. As

previously discussed, the SFO and OVLT are unique in their ability to sample the constituents of

the systemic circulation and promote activation of circuits to preserve homeostasis. Activation

has been shown to alter the balance between excitatory and inhibitory projections to the PVN

and SON which in turn strongly influences the activity of the HPA axis and behavior.

Recent investigation into the effects of hypertonic saline injection on humoral,

behavioral, and central measures of stress-responsiveness in male rats and mice revealed a

prominent role for activation of oxytocinergic pathways. Hypertonic (2.0 M) saline injected

subcutaneously followed by 60 min water deprivation resulted in decreased ACTH,

Page 28: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

28

glucocorticoid, and plasma renin activity during a 60 min restraint period, but significantly

increased plasma oxytocin levels (Krause et al., 2011a). Cardiovascular recordings determined

that rats rendered mildly hypernatremic and subsequently exposed to restraint stress had mean

arterial pressure and heart-rate variability return to prerestraint levels faster than normonatremic

controls (Krause et al., 2011a). Consistent with previous reports (Pirnik et al., 2004), acute

hypernatremia increased Fos induction in vasopressin and oxytocin expressing neurons in the

PVN and SON, an effect that was predictive of increased social interactions but decreased

anxiety-like behavior in the elevated plus maze (Krause et al., 2011a). Acute hypernatremia and

restraint stress interacted to increase Fos expression in the OVLT, oval capsule of the BNST,

ventral lateral septum, and central nucleus of the amygdala (Frazier et al., 2013). However, the

increased neuronal activation that occurred subsequent to acute hypernatremia and restraint was

specific to these brain regions as the medial prefrontal cortex, lateral ventral BNST, and

dorsomedial hypothalamus showed no significant change in Fos expression (Frazier et al., 2013).

Whole cell patch clamp recordings revealed that acute hypernatremia caused inhibition of CRH

neurons in the PVN that was dependent on activation of oxytocin receptors (Frazier et al., 2013).

Several lines of investigation (Huang et al., 2014) are actively pursuing the neuronal

mechanism underlying the dampened HPA axis activation that occurs with acute hypernatremia.

Seminal studies conducted by Ludwig et al. discovered that subsequent to the systemic release of

vasopressin and oxytocin that follows peripheral salt-loading, brain levels of these neuropeptides

exhibit a robust and sustained increase (Ludwig et al., 1994). The implication is that

magnocellular neurons in the SON and PVN release neuropeptides from dendrites and these

neuropeptides may act as autocrine or paracrine signals at their site of origin. In this regard,

dendritic release of vasopressin serves as a powerful autocrine signal that modulates firing

Page 29: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

29

patterns and activity of magnocellular vasopressin neurons in the SON and PVN (Ludwig and

Leng, 1997; Gouzenes et al., 1998). Recently, dendritically-released neuropeptide also was

found to signal in a paracrine fashion to influence the functional activity of neighboring neurons

of a differing neuronal phenotype. Specifically, the Stern laboratory discovered that dendritic

release of vasopressin stimulated neighboring presympathetic neurons in the PVN to drive

sympathoexcitation following a hyperosmotic stimulus (Son et al., 2013). Hyperosmotic stimuli

also trigger the dendritic release of oxytocin in the hypothalamus (Ludwig and Leng, 2006;

Mabrouk and Kennedy, 2012) and oxytocin neurons are in close proximity to CRH neurons that

express oxytocin receptor mRNA in the PVN (Dabrowska et al., 2011). Therefore, it is possible

that the inhibition of CRH neurons that accompanies acute hypernatremia maybe regulated by

dendritically released oxytocin that activates oxytocin receptors expressed on CRH neurons in

the PVN.

Identifying the oxytocinergic circuits underlying the anxiolytic effects of acute

hypernatremia may reveal neuronal mechanisms that quell excitation of neurons that facilitate

physiological and behavioral manifestations of fear and anxiety. The magnocellular neurons in

the PVN and SON are the major source of oxytocin but whether these neurons supplied oxytocin

to forebrain nuclei controlling mood and affect was unclear because their axons are known to

terminate in the posterior pituitary where they release oxytocin into the systemic circulation.

Consequently, endogenously generated oxytocin was hypothesized to activate oxytocin receptors

in forebrain nuclei by diffusion of dendritically released peptide (Ludwig and Leng, 2006) or via

axonal release from a subset of parvocellular oxytocinergic neurons residing in the PVN that do

not project to the pituitary (Swanson and Sawchenko, 1983). These hypotheses about the source

of centrally released oxytocin have undergone a paradigm shift as the result of recent research

Page 30: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

30

utilizing advanced techniques in neuroanatomical tract-tracing. These elegant studies conducted

by Grinevich and colleagues discovered that oxytocin neurons in the PVN and SON that project

to the posterior pituitary also send axon collaterals to forebrain nuclei mediating stress-

responsiveness and anxiety-like behavior (Knobloch et al 2012). In other words, oxytocinergic

magnocellular neurons in the PVN and SON simultaneously project to the posterior pituitary and

the forebrain, presenting the possibility that systemic and central release of oxytocin maybe

accomplished by the same neurons. In rodents, magnocellular oxytocinergic neurons in the PVN

and SON are robustly excited by elevations in the pNa+ (Krause et al., 2011a; Frazier et al.,

2013). Therefore, it is possible that acute hypernatremia is a systemic stimulus that excites

hypothalamic osmosensing neurons to promote the release of oxytocin into systemic circulation,

and subsequently, stimulates the release of oxytocin from axons terminating in forebrain nuclei

known to mediate physiological and behavioral manifestations of fear and anxiety. Amygdalar

CRH neurons are implicated in the development of anxiety disorders and further research is

warranted to determine whether this potential oxytocinergic pathway is a viable therapeutic

target.

In addition to the central nucleus of the amygdala, the extensive mapping of

oxytocinergic afferents throughout the rat brain isolated dense innervation of the BNST

(Knobloch et al., 2012). The BNST is a complex set of at least a dozen sub-nuclei that mediate

the information gathered by higher-order perceptual and memory-based brain regions through an

integrating process involving the hindbrain and hypothalamus (Dong et al., 2001; Bota et al.,

2012). Because of this pivotal role, the BNST is rich in neurochemical diversity, both locally

and in terms of the phenotype of afferent and efferent projections (Dong and Swanson, 2006;

Bota et al., 2012). Rainnie and colleagues used viral anterograde tracing of PVN neurons to

Page 31: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

31

reveal that oxytocin immunoreactive fibers terminating in the BNST originate from neurons in

the PVN (Dabrowska et al., 2011) and subsequent studies suggest that these cells are

osmosensing magnocellular neurons (Knobloch et al., 2012). Acute hypernatremia followed by

restraint-stress robustly activated oxytocin neurons in the PVN (Krause et al 2011a) as well as

neurons residing in the oval capsule portion of the BNST (Frazier et al., 2013). These studies

support the hypothesis that elevations in body sodium levels excite magnocellular oxytocin

neurons and subsequent psychogenic stress exposure may trigger the release of oxytocin from

axons terminating in the BNST, which in turn, activate neurons within this nucleus. The pattern

of neuronal activation that occurs after acute hypernatremia and restraint is predictive of

decreased anxiety-like behavior and increased social interactions (Krause et al., 2011a). In this

regard, Kim et al found that two BNST sub-regions have highly specific roles in mediating

separate features of anxiety such as risk-avoidance and respiratory rate (Kim et al., 2013). While

this study found that activation of the BNST increased anxiety, activation associated with acute

hypernatremia seemed to be related to a decrease in anxiety-like behavior. As mentioned, the

BNST is a highly diverse region in terms of cellular capabilities and connectivity that may be the

basis for understanding the apparent discrepancies in activation of this region. Ultimately, the

circuit level data that has produced valuable insights into the effects of activating cells in these

areas will have to be complemented by studies that take into consideration ligand-receptor

interactions.

An acute osmotic challenge induces anxiolytic and pro-social effects associated with an

increase in oxytocin-producing cell activation and a reduction in the HPA and behavioral

components of the stress response. As oxytocin cells in the PVN are both osmosensitive and

receive projections from osmosensitive areas of the brain, the increase in activation is likely due

Page 32: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

32

to the pivotal role of oxytocin in rectifying the fluid imbalance. Through dendritic release and

collateralized axonal projections that target the posterior pituitary and limbic brain regions

influencing mood and behavior, oxytocin signaling orchestrates the humoral and behavioral

responses necessary to maintain hydromineral homeostasis specifically in the context of an acute

hyperosmotic stimulus.

Chronic Salt-Loading

Maintenance on a high-salt diet involves the replacement of standard rodent chow usually

composed of 0.3-0.6% NaCl with chow containing 3-8% NaCl (Yu et al., 1998; Fang et al.,

2000; Sanders, 2009). In contrast to other forms of salt-loading, high-salt diets generally allow

the experimental subject ad libitum access to water. As a result, the physiological and

psychological stress may be considered less severe than other salt-loading paradigms and in

certain cases may be an ideal model of hypertension in salt-sensitive individuals (Sanders, 2009).

High-salt diets increase blood pressure via volume expansion (Cowley, 1997), by increasing

pNa+ levels which activate central osmosensors to reflexively increase blood pressure (Johnson

and Gross, 1993), or by both mechanisms (Fang et al., 2000). Perhaps the most striking long-

term effects of a high-salt diet are those published by Meneely and Ball who found that in a

study of 679 male rats, survival rates decreased dramatically when dietary salt was increased

(Meneely and Ball, 1958). In this study, 75% of rats eating chow composed of 0.15%-2.0%

NaCl were still alive after 16 months, however after 5 months only 5% of rats eating 21% NaCl

chow were still alive. Gradations of NaCl content between 0.15% and 21% predicted mortality

as consumption of increased concentrations were positively correlated with increased death rates.

Death was usually preceded by a “nephrosis-like syndrome” and abrupt onset of massive edema.

Thus, even with access to water, high dietary salt intake can still significantly impact mortality

through vascular lesions (Tobian and Hanlon, 1990), albuminuria and collagen deposition (Yu et

Page 33: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

33

al., 1998), cerebrovascular accidents (Tobian and Hanlon, 1990), and accelerated renal failure

(Sanders et al., 2001).

Chronic salt-loading can also be accomplished by replacing drinking water with sodium

chloride solutions generally consisting of 1.5%-3.0% NaCl. Perhaps the most severe

experimental paradigm of those described, chronic salt-loading both induces cellular dehydration

and with time depletes volume (Watts et al., 1995). Early reports (Duchen, 1962) of salt-loading

lasted up to 21 days after which point rats weighing 200-250g at the beginning of the study had

lost up to 60g. Additionally, rats lost all traces of what was then called “neurosecretory

material” in the posterior lobe of the pituitary which disappeared commensurate with vasopressin

secretion during the first week. This extensive remodeling of pituitary tissue morphology is

indicative of the intense involvement of pituitary hormones in the homeostatic response.

Comparison of the effects of rats eating 1% NaCl chow and rats drinking 1% NaCl saline found

that drinking saline resulted in a 50% reduction in life span compared with high-salt diet

consumption as well as a 66% rate in the development of hypertension (Koletsky, 1958; 1959).

Furthermore, it was determined that the vascular, renal, and cardiac lesions caused by drinking

saline were identical to those induced by pharmacologically activating mineralocorticoid

receptors via deoxycorticosterone treatment (Selye, 1943; Selye et al., 1943; Selye and Stone,

1943) which similarly overloads the body with sodium. The most common reason for death in

chronic salt loaded rats was pneumonia associated with renal dysfunction (Koletsky, 1961).

While the vascular injuries caused by salt-loading were very similar to those found in humans

with essential hypertension, they were not found to be identical (Koletsky, 1961). Essentially,

chronic salt-loading through forced saline imbibition is a potent stimulus with many

physiological correlates to the development of human hypertension especially in those that may

Page 34: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

34

be salt-sensitive; therefore, the alterations in the brain that occur during chronic salt-loading may

also be mechanistically related to the development of essential hypertension. In pursuit of these

mechanisms, we designed our experiments in Chapters 4 and 5 to investigate changes that occur

in the early (Day 5) stages of these central alterations when peripheral causes for hypertension

have not yet fully developed.

The chronic salt-loading paradigm typically results in systemic osmoregulatory responses

that are pushed close to or beyond homeostatic capacity. The phases described by Watts et al

involve an initial short (~24-48 hr) successful adaptation that allows for normal food intake and

defense against intravascular volume loss that is then followed by a progressively increasing

hematocrit and plasma osmolality (Watts et al., 1995). The initial phase involving successful

homeostasis is also characterized by elevated CORT, which then gradually becomes suppressed

during the course of the salt-loading term. Furthermore, the end products of the RAAS,

angiotensin-II and aldosterone, have the potential to become involved in an adverse relationship

as angiotensin-II has both the effects of maintaining blood pressure in response to volume loss

and increasing aldosterone levels (Laragh et al., 1960), the latter of which is counterproductive in

the context of elevated pNa+. Regarding this adverse relationship, Husain et al. found that while

aldosterone was potently suppressed during chronic salt-loading, plasma angiotensin-II levels

were not different from controls (Husain et al., 1987). This was attributed to a sequestration of

circulating angiotensin-II in the adrenal gland by receptor mediated endocytosis which would

allow for the action of angiotensin-II to constrict the vasculature in response to increasing

hypovolemia without stimulating aldosterone release. Rats drinking hypertonic saline instead of

water develop hypertension, vascular and renal necrosis, and have shortened life spans which are

complications that take longer to develop in, but that are similar to those observed in the

Page 35: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

35

deoxycorticosterone-salt model of cardiovascular disease (Koletsky, 1961). Thus, the highly

dysregulated humoral constituency is directed towards maintaining blood pressure and hydration

of each the body’s fluid compartments, but which also are involved with the development of

cardiovascular and renal damage.

During the course of a standard five- to seven-day experimental period of drinking

hypertonic saline, the continued threat of long-term intracellular dehydration and hypotension

results in behavioral adaptations characterized by drinking large volumes of fluid, but limiting

food intake (Watts et al., 1995). However, increased saline intake has not always been the

observed behavior, as rats were found to exhibit highly variable fluid intakes in response to salt-

loading (Watts et al., 1995). It is possible that in this case, the behavioral response variability

may reflect a conflict between goal-directed motor responses aimed towards increased water-

seeking behavior and altered salt-appetite which may vary with each individual.

As discussed above, the paradigm of chronic salt-loading results in several deviations

from corresponding homeostatic set points, requiring an appropriately altered stress response to

promote survival in the face of a maligned internal environment. Stressful psychological events

experienced in the context of the physiological stress of forced hypertonic saline administration

have shown that HPA axis sensitivity is augmented during hydromineral imbalance. Rats

maintained on 2% NaCl for seven days and then rehydrated presented with sensitization of HPA

axis measures upon rehydration (Amaya et al., 2001). These rats had decreased CORT during

dehydration, but exaggerated their basal and restraint-induced ACTH responses after one week

of rehydration. Parvocellular CRH decreased with salt-loading, but increased during rehydration

while vasopressin increased during both periods. Interestingly, plasma vasopressin levels

increased with restraint during dehydration, but CRH did not suggesting that vasopressin has an

Page 36: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

36

important role in stress-responsiveness specifically during the course of long-term osmotic

challenge when ACTH secretion is inhibited.

Chronic salt-loading is known to modulate the expression of CRH in the PVN through

circuits originating in the lamina terminalis. Kovacs and Sawchenko found that rats maintained

on 2% NaCl solution instead of water for seven days exhibited a shift in CRH mRNA expression

from a parvocellular region to a magnocellular distribution (Kovacs and Sawchenko, 1993).

This shift was abolished with knife cuts dissociating the connectivity between the OVLT and

SON, but not when pathways between the median preoptic nucleus and SFO or median preoptic

nucleus and SON were severed, promoting the idea that the OVLT is the primary osmosensitive

region influencing CRH plasticity in the PVN of rats (Kovacs and Sawchenko, 1993). This

study also provided evidence that the increased magnocellular expression of CRH was specific to

oxytocin cells.

In support of these findings, another study found increased CRH mRNA in oxytocin cells

of the PVN and SON in rats using the same protocol (Imaki et al., 2001), but also found

increased CRH type 1 receptors and increased urocortin (a CRH type-2 receptor ligand) in both

oxytocin and vasopressin magnocellular cells of the PVN and SON (Imaki et al., 2001). Salt-

loading also decreased CRH type 1 receptors in parvocellular neurons of the PVN, but increased

their expression in the intermediate pituitary (Imaki et al., 2001). The purpose of an increased

influence of CRH in the intermediate pituitary is unclear; however, it has been suggested that

corelease of CRH with magnocellular neuropeptides potentiates the release of the latter (Bondy

and Gainer, 1989). Overall, the relationship between where CRH synthesis goes up relative to

where receptor expression increases following osmotic stimulation is not exactly known.

However, CRH expression does increase with salt-loading in neurons that project to the PVN.

Page 37: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

37

These regions include the fusiform nucleus of the BNST, periventricular nucleus, and lateral

hypothalamic area (Watts et al., 1995). Furthermore, CRH expression is increased in the

amygdala with salt-loading, and the amygdala contains a dense population of neurons projecting

to the BNST (Sun et al., 1991). Interesting enough, it seems that concomitant increases in CRH

and its receptor is specific to PVN neurons (Rivest et al., 1995). Together, the changing CRH

system in the brain during salt-loading may reflect a concert of mechanisms working through the

PVN to alter the pituitary release of ACTH and posterior pituitary hormones.

The components of the HPA axis and osmoregulatory mechanisms are conserved across

mammalian species; however, it is possible that salt-related plastic changes in CRH and CRH

type 1 receptor gene expression in rats may not be generalized to the mouse model. In a

comparative study, restraint-induced CRH mRNA upregulation and Fos expression occurred in

the PVN of both rat and mouse (Imaki et al., 2003). However, restraint- and hypertonically-

induced increases in CRH type 1 receptors as well as hypertonically-stimulated CRH expression

in PVN magnocellular neurons were all found to be specific to rats with no significant changes

noted in mice (Imaki et al., 2003). In mice, CRH type 1 receptor expression was mainly

colocalized with CRH-positive cells and very few oxytocin or vasopressin cells suggesting an

autoregulatory or positive feedback mechanism as well as a potential species-specific

intermediary between salt-balance and stress-responsiveness (Imaki et al., 2003). The increase

in CRH expression found in rat SON with salt-loading was found to not occur in mice, a

difference confirmed by others (Hinks et al., 1995). Further investigation into the apparent

differences in mouse and rat neuroplastic responses to salt-loading are indeed warranted.

The alteration in components of stress-responsive neurons in the PVN is predictive of

altered autonomic stress-responsiveness. A year-long experiment was carried out to test the

Page 38: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

38

effects of conflict stress on baboons that were maintained on a normal or high-salt diet (Turkkan,

1994). In this endeavor, it was found that a five-month period elapsed before high-salt diet alone

significantly impacted a rise in mean arterial pressure of 6+/-5 mmHg (Turkkan, 1994) .

Following twice daily food/shock conflict sessions, stress and a high-salt diet combined had

almost three times the effect on blood pressure compared with high-salt diet alone with a rise in

mean arterial pressure of 17+/-3 mmHg. Individual representative cases revealed indices of

stress-sensitivity that positively correlated with the impact high-salt diet had on increased mean

arterial pressure. These include urinary free cortisol, degree of indecision during conflict stress,

as well as pressure diuresis and natriuresis. It is therefore evident that the physiological stress

brought about by high sodium intake can exacerbate the autonomic response to psychological

stress, thereby contributing to the development of hypertension and associated cardiovascular

disease.

Parvocellular neurons in the PVN likely modulate autonomic function through the release

of CRH in the rostroventrolateral medulla (RVLM). As with any long-term challenge involving

blood pressure regulation, the core network of barosensitive sympathetic efferents composed of

connectivity between the RVLM, nucleus of the solitary tract, spinal cord, and PVN is poised to

be altered to counter chronic hyperosmotic conditions. Regarding the RVLM, barosensitive

neurons residing in this region are hyperactive in spontaneously hypertensive, salt-sensitive, and

renal hypertensive rat strains (Guyenet, 2006) and PVN neurons projecting to the RVLM induce

this hyperactivity during water deprivation (Stocker et al., 2005; Stocker et al., 2006). Milner et

al. found that CRH-releasing parvocellular PVN neurons project to the RVLM where CRH

binding to the CRH receptor caused a sustained increase in mean arterial pressure (Milner et al.,

1993). Furthermore, in addition to local production of CRH by neurons in the RVLM, afferent

Page 39: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

39

input from CRH-positive PVN neurons forms excitatory asymmetric synapses on RVLM

neurons. Ultrastructural localization of these synapses were found most frequently innervating

distal dendrites suggesting that CRH release in the RVLM serves more of a modulatory role

rather than a direct influence on neuronal activity. Taken together, the PVN to RVLM circuit

composed of parvocellular CRH containing neurons is situated to coincidentally influence the

HPA axis and autonomic response to stress. Thus, CRH reconfiguration in the PVN during

chronic salt-loading may involve increased CRH production in neurons projecting to the RVLM.

Aims

Acute Salt-Loading

Previously published research found that acute salt-loading is anxiolytic and dampens

stress-responsiveness in laboratory rats (Krause et al., 2011a; Frazier et al., 2013). While these

effects were consistently replicated in rats, whether these effects generalize to mice, and whether

acute salt-loading decreases neuronal activation in PVN neurons genetically modified to

fluorescently report for CRH transcription has yet to be determined. Furthermore, methods to

identify CRH mRNA in the brain have been limited by poor spatial resolution suggesting that

utilizing an advanced method to detect CRH in the brain will be a significant contribution to this

area of research.

The aims of Chapter 3 were to determine whether acute salt-loading decreases anxiety-

like behavior and restraint-induced HPA axis activation in mice and to evaluate the associated

neuronal activation in the PVN. Mice rendered mildly hypernatremic via systemic

administration of 2.0 M NaCl were subjected to restraint or behavioral testing on the elevated

plus maze. Quantification of Fos-induction in both CRH- and oxytocin-positive neurons in the

PVN served to provide insight towards neural mechanisms contributing to the HPA axis-

dampening and anxiolytic effects of acute mild hypernatremia.

Page 40: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

40

Chronic Salt-Loading

The anxiolytic and HPA axis suppressing effects of acute salt-loading may have

reinforcing value leading to an over-consumption of salt. High sodium intake leads to

hypertension in salt-sensitive individuals, and elevated blood pressure is a risk factor for

developing a heart, vascular, and renal pathologies. Chronic salt-loading causes many of the

physiological manifestations associated with cardiovascular and renal disease, but also induces

relevant alterations in neuropeptide expression in the PVN. These alterations involve the same

PVN neuropeptides active in the response to acute hypernatremia, and include rearrangement of

CRH expression and increased systemic release of vasopressin and oxytocin. However, the

majority of chronic salt-loading experiments have been performed in rats and experiments in

mice have not always supported data using a rat model. Given the increased usage of mice in

research, substantiation of the central and behavioral effects of salt-loading in mice is a general

goal of these experiments. Collectively, the following aims serve to identify specific changes in

behavior and neuropeptide expression associated with a known progression of physiological

alterations that result in hypertension, cardiovascular and renal damage, and attenuated life-span.

The first and second aims are covered in Chapter 4 while the third and fourth aims are covered in

Chapter 5.

The first aim was to evaluate ingestive behavior and the effects of drinking 2% NaCl on

pNa+, hematocrit, and plasma proteins. One of the goals of Chapter 4 was to determine if

transgenic reporter mice drinking 2% NaCl for five days would alter ingestive behavior to

increase fluid intake and decrease food intake. The literature conflicts as to whether volume of

2% NaCl consumed is highly variable or only increases substantially. Additionally, feeding

behavior in the mouse is comparatively less well studied than the rat, meaning that these data

will serve to substantiate the murine response to salt-loading. Subsequent experiments

Page 41: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

41

investigated changes in neuronal activation and neuropeptide expression that were to be

interpreted with respect to the changes in ingestive behavior and plasma constituency found in

Aim 1 i.e. the changes in the brain viewed in the context of anorexia, polydipsia, hypovolemia,

etc.

The second aim included evaluation of neuronal activation and the HPA axis response to

restraint. While the literature indicates that in the rat PVN and SON neurons are activated and,

more specifically, vasopressin neurons are activated in these regions, similar experiments in mice

have yielded conflicting results. Moreover, the rostro-caudal organization of the PVN is such

that neuroendocrine neurons are largely parceled out from preautonomic neurons suggesting that

evaluation of total neuronal activation will indicate the degree to which these functions are

involved in the regulatory response. Therefore, the goals of the second aim were to determine if

1) vasopressin neurons were activated 2) restraint or salt-loading had an effect on total Fos in

neuroendocrine or preautonomic subregions of the PVN 3) restraint or salt-loading had an effect

on PVN neurons reporting for CRH transcription and 4) the CORT response to restraint was

suppressed. Based on the literature and our acute salt-loading experiments, we hypothesized that

vasopressin neurons would be activated to conserve water, but activation of CRH neurons would

be suppressed as ACTH and CORT in salt-loaded rats are attenuated. We hypothesized that the

surge in CORT in response to restraint would similarly be suppressed in mice. Furthermore,

activation of preautonomic neurons located mainly in the posterior region of the PVN has to our

knowledge never been addressed following chronic salt-loading. As substantial changes are

known to occur centrally to accommodate high salt-intake, we hypothesized that preautonomic

neurons in the PVN would increase their activity. The data obtained from these experiments

Page 42: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

42

serve to provide a foundational perspective of the cell types and therefore the functions that are

active to mediate the central response to salt-loading.

The third aim was to determine the extent of CRH colocalizations with vasopressin and

oxytocin neurons in the PVN and SON. This aim necessitated the detailed evaluation of CRH

mRNA expression using both a conventional method and genetic reporting for CRH

transcription. Salt-loading induced rearrangement of CRH expression in the rat has not always

been confirmed by similar experiments performed in mice. Furthermore, the limitations inherent

in using colchicine pretreatment to utilize antibodies directed against the CRH peptide make

colocalization experiments difficult. Moreover, the use of more than one method to isolate CRH

transcript represents an important methodological investigation into a novel technique to isolate

CRH positive neurons. The experiments in Chapter 5 included immunohistochemical

investigation with antibodies for vasopressin and oxytocin in brain tissue from CRH reporter

mice to compare increases in CRH induced by salt-loading within these particular cell types in

the PVN and SON. These results indicate the neuropeptides involved with the central response

to salt-loading.

The final aim was to determine the extent of salt-loading induced CRH and vasopressin

expression in PVN preautonomic neurons projecting to the RVLM. The PVN coordinates both

neuroendocrine and autonomic functions. Preautonomic neurons were identified using the

retrograde tract-tracer Fluorogold (FG) injected into the RVLM. The ultimate goal of Chapter 4

was to test the hypothesis that CRH and vasopressin increase expression in a defined circuit

composed of PVN projections to the RVLM, thereby coordinating neuroendocrine control of the

response to salt-loading with centrally-mediated autonomic regulation.

Page 43: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

43

Figure 1-1. Schematic of the major systems that interact with neurons in the paraventricular

nucleus of the hypothalamus to maintain hydromineral homeostasis. Image is of a

coronal section taken through the brain of a CRH reporter mouse (see Chapter 2 for

description). The image includes a unilateral portion of the PVN with fluorescent

reporting for CRH (red), immunohistochemical labeling of vasopressin neurons

(green), and neurons retrogradely-labeled with tract tracer injected into the

rostroventrolateral medulla (magenta).

Page 44: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

44

Figure 1-2. Schematic of CRH pathways projecting from the PVN. While the effects of CRH

released from neurons residing in the PVN is well characterized for pituitary-

projecting neurons, the functions of CRH released in other areas are less well

understood. However, the presence of CRH receptors and functional studies

microinjecting CRH in these areas suggests a role in autonomic regulation.

Page 45: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

45

Figure 1-3. Osmosensitive and stress-responsive circuits. Circuits adapted in part from Ulrich-

Lai and Herman, 2009 and Bourque, 2008 demonstrating key hypothalamic circuits

that mediate the behavioral, endocrine, and autonomic stress-response (shown in red

and green). Many of these circuits are also activated when the plasma osmolality is

elevated above a threshold set point (shown in blue and red). The circuits shown in

red represent stress-responsive pathways that may be altered during an osmotic

challenge such as hypernatremia. This figure is not exhaustive and does not represent

circuits that are necessarily monosynaptic.

Page 46: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

46

CHAPTER 2

MATERIALS AND METHODS

Generation of CRH Reporter Mice

Adult male CRH reporter mice were generated as previously described (Taniguchi et al.,

2011). Briefly, induction of tdTomato red fluorescent protein to indicate CRH transcription in

neurons was accomplished by the generation of B6(Cg)-Crhtm1(cre)Zjh/J knockin mice (Jackson

Laboratory Stock # 012704) expressing a Cre recombinase coding region immediately after the

STOP codon terminating CRH transcription. These mice were then crossed with

Gt(ROSA)26Sortm14(CAG-tdTomato)Hze congenic mice (Jackson Laboratory Stock # 007914)

expressing a loxP-flanked STOP cassette preventing tdTomato transcription. CRH reporter mice

were 8-12 weeks old at the beginning of the experiments and were maintained on a 12:12 h

light/dark cycle in clear plastic ventilated cages with ad libitum access to pelleted chow (Harlan

Teklad) and water (except where otherwise noted). All procedures were approved by the

Institutional Animal Care and Use Committee at the University of Florida.

Validation of CRH Reporter

RNAscope in situ hybridization (ISH) was performed on brain tissue collected from

CRH-reporter mice to determine the extent to which CRH mRNA colocalizes with tdTomato in

the PVN. Mice were overdosed with sodium pentobarbital, transcardially perfused with 0.9%

saline followed by 4% paraformaldehyde (PFA). Subsequently, brains were extracted, coronally

sectioned at 20 µm into 6 series and then immediately rinsed and mounted onto Superfrost Plus

Gold slides. Tissue collection, sectioning and mounting of sections were performed in RNase-

free conditions. Slides were allowed to air dry for 20-30 min and then were stored at -80°C until

processing for in situ hybridization. Three slides containing separate series of sections through

the PVN were allowed to reach room temperature for 30 min prior to performing the

Page 47: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

47

manufacturer’s protocol (Advanced Cell Diagnostics; Hayward, CA). RNAscope ISH was

performed using the following probes: (1) Negative Control, DapB, (2) Positive control, Ubc, (3)

CRH. All images were captured at 40x magnification and the exposure time was adjusted for

each image using the best-fit feature in Axiovision. Subsequently, the min-max feature was

utilized to minimize background fluorescence and provide optimal visualization of RNA signal.

All images were processed using the same automated parameters.

Figure 2-1 (A-C) depicts a unilateral PVN fluorescent photomicrograph from a coronal

section taken from the brain of a CRH-reporter mouse and processed for in situ hybridization.

Highly specific single-strand fluorescent labeling of CRH mRNA revealed consistent

colocalization of CRH probe and tdTomato fluorescence (Fig. 2-1C). Figure 2-1D shows a

qualitative image of the ventral portion of a coronal brain section taken from the hypothalamus

of a CRH-reporter mouse and processed for immunofluorescent labeling of oxytocin. Consistent

with hypothalamic-pituitary axonal projections, green oxytocin fibers and red tdTomato fibers

are seen in the internal and external zones of the median eminence, respectively.

Antibody Characterization

The details of the antibodies used in this study are given in Table 2-1. Both the PS-38

mouse anti oxytocin-neurophysin (oxytocin-NP) and PS-41 mouse anti vasopressin-neurophysin

(vasopressin-NP) antibodies were previously screened for specificity and lack of cross-reactivity

by Ben-Barak and colleagues (Ben-Barak et al., 1985). Following initial harvesting, solid phase

ELISA was used to isolate posterior pituitary-positive hybridoma supernatant from anterior

pituitary- or cerebellar-positive supernatant. An immunohistochemistry assay was then

performed to determine which antibodies selectively stained hypothalamic and pituitary cells,

and those with the predicted pattern were cloned. Liquid phase RIA results using iodinated rat

neurophysin from both normal and Brattleboro rats showed that PS-38 reacted potently to normal

Page 48: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

48

and Brattleboro tissue, and PS-41 failed to produce any signal in Brattleboro tissue.

Furthermore, iodinated vasopressin-NP was highly reactive with vasopressin-NP antibody, but

did not react with PS-38 and iodinated oxytocin-NP performed similarly (Ben-Barak et al.,

1985). In our experiments, labeling with PS-38 and PS-41 resulted in clear neuronal profiles

with distinctive patterns in the PVN and SON which were consistent with previously published

studies (Castel and Morris, 1988). Omission of the primary antibody during the

immunofluorescent labeling protocol outlined below resulted in a complete loss of detectable

fluorescence.

Fluorogold could readily be identified at the site of injection. Labeling was also apparent

unilaterally in PVN cells without the use of an antibody by imaging with a DAPI filter set.

However, labeling intensity was improved with primary incubation in AB153, which was

previously validated in mouse brain tissue (Dimitrov et al., 2013).

Two anti-c-Fos antibodies were used to identify activated neurons. Both antibodies

produced a nuclear staining pattern indicative of activated neurons in the PVN following

restraint, but produced little to no reactivity in the PVN of unrestrained mice. Furthermore, both

antibodies revealed bands in western blots in the range of 50-65 kDa.

Immunofluorescence

Immunohistochemical identification of oxytocin/vasopressin neurophysin, Fos and FG in

the PVN was performed using indirect fluorescence. Rinses and solutions were made using 50

mM potassium phosphate buffered saline and took place at room temperature on an orbital

shaker unless otherwise noted. Single labeling began by rinsing free-floating sections 5 X 5 min

to remove cryoprotectant. Blocking consisted of 2% normal donkey serum (Jackson

ImmunoResearch) with 0.2% Triton-X (Sigma) for 1 h followed by incubation with primary

antibody in blocking solution overnight at 4° C. The next day, sections were rinsed 5 X 5 min

Page 49: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

49

and incubated in blocking solution with either one or two secondary antibodies. After a final

series of rinses, sections were mounted on Superfrost Plus slides (Fisher) in KPB, allowed to air

dry, and then coverslipped using polyvinyl alcohol mounting medium with DABCO (Sigma).

Image Capture and Analysis

Images of the PVN were captured using an AxioImager M.2 fluorescent microscope

(Carl Zeiss, Thornwood, New York) equipped with an Apotome.2 and connected to a PC

running Axiovision 4.8. Using a Plan-Apochromat 20x/0.8 M27 objective, z-stacks were

captured from sections corresponding to various rostral-caudal levels through the PVN or SON.

Each image was assigned to a specific atlas figure by an individual experimenter that analyzed

the shape of the PVN proper and matched it with the shapes of adjacent anatomical landmarks

found in The Mouse Brain in Stereotaxic Coordinates 3rd Ed (Franklin and Paxinos, 2008).

Image size was approximately 1388 X 1040 pixels and each z-step was 1 µm with an average of

20 optical sections per PVN image. When capturing images from sections processed for

immunohistochemistry, exposure time was determined based on an overexposure reporting

function automatically set by the software. Exposure time during the imaging process varied and

are reported in each chapter. Images were adjusted using the brightness and contrast settings

available in Axiovision to increase clarity.

Quantification of colocalizations took place using a method similar to the “direct three-

dimensional counting” process (Williams and Rakic, 1988). Each z-stack was opened in

Axiovision 4.8 or Zen lite 2012 (Carl Zeiss) and then each image was inspected while alternately

turning the respective channels on and off to identify potential colocalizations. Candidate cells

were then further analyzed as necessary by removing the area of the image containing the cell

into a separate region of interest (ROI) image to generate a 3D reconstruction. If, as a 3D

reconstruction, the cell contained fluorescent labeling on each individual channel throughout the

Page 50: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

50

z-plane of the image, then the cell was counted. The number of bi-lateral colocalizations were

determined for each atlas-matched section containing the PVN. An identical procedure was

repeated for image capture and analysis of the SON. Statistical analyses were performed by

grouping according to atlas region and in a separate analysis by grouping all images obtained

from each subject.

Page 51: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

51

Table 2-1. Antibody information.

Antibody Antigen Species Dilution Manufacturer

Sc-52 c-Fos Rabbit Polyclonal 1:1000 Santa Cruz

MCA-2H2 c-Fos Mouse Monoclonal 1:2000 Encor

Biotechnology

PS-38 Oxytocin

Neurophysin

Mouse Monoclonal 1:400 Dr. Gainer,

NIH

PS-41 Arginine-

Vasopressin

Neurophysin

Mouse Monoclonal 1:400 Dr. Gainer,

NIH

AB153 Fluorogold Rabbit Polyclonal 1:1000 Millipore

Alex Fluor 488 Mouse IgG (H+L) Donkey 1:500 Jackson

Alexa Fluor 647 Rabbit IgG (H+L) Donkey 1:500 Jackson

Alkaline

Phosphatase, Fab

Fragments

Digoxigenin Sheep Polyclonal 1:1500 Roche

Page 52: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

52

Figure 2-1. Representative images validating CRH-Cre induced tdTomato expression. A-C: A

unilateral coronal section through the PVN depicting two images of the same set of

neurons labeled for (A) CRH mRNA probe amplification (green) and (B) tdTomato

(red soma). A merged image (C) illustrates a high degree of CRH mRNA and

tdTomato colocalization. Scale bars = 20 m. D: tdTomato expression adjacent to

oxytocin fibers in the median eminence. Scale bar = 100 m.

Page 53: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

53

CHAPTER 3

CENTRAL AND BEHAVIORAL EFFECTS OF ACUTE SALT-LOADING

Background

In mammals the pNa+ is regulated by neural, humoral and behavioral mechanisms that

maintain blood tonicity at levels that allow normal physiological function. Sodium deficiency

causes hyponatremia and increases circulating levels of angiotensin-II and aldosterone which

activate receptors in the kidney and brain to restore the pNa+ to homeostatic levels by promoting

the retention and consumption of sodium. Conversely, excess sodium causes hypernatremia and

suppresses angiotensin-II and aldosterone but causes the secretion of vasopressin and oxytocin

into the systemic circulation which alleviates the elevated pNa+ by promoting renal water

retention and sodium excretion. Thus, the pNa+ is tightly regulated by neurohumoral

compensatory responses that act in the brain and periphery to maintain blood tonicity at

homeostatic levels when challenged with sodium deficiency or excess.

The neuropeptides and hormones that maintain the pNa+ are also known to influence

mood, affect, and stress-responsiveness. For example, studies conducted in humans and animals

have found that elevated circulating levels of angiotensin-II and aldosterone are predictive of

affective disorder (Murck et al., 2003b; Grippo et al., 2005; Niebylski et al., 2012; Häfner et al.,

2013), and that oxytocin and vasopressin are mediators of stress-responsiveness and anxiety

(Neumann and Landgraf, 2012a; Benarroch, 2013a). As mentioned, the secretion of angiotensin-

II, aldosterone, vasopressin, and oxytocin are heavily influenced by the pNa+, and therefore, it is

possible that alterations in body sodium levels affect stress responding and anxiety-like behavior

through manipulation of these neuroendocrine signals. In this regard, previous results indicate

that sodium depletion is anxiogenic (Leshem, 2011), but acute salt-loading is anxiolytic and

dampens stress-responsiveness in laboratory rats (Krause et al., 2011a; Frazier et al., 2013).

Page 54: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

54

While the changes in the pNa+ are found to influence mood and stress responding in rats,

whether these effects generalize to mice, and therefore, allow the use of the genetic

manipulations that mouse models afford to investigate central mechanism(s) underlying the

stress limiting effects of acute hypernatremia has not been evaluated.

The goal of the present study was to determine whether acute modest increases in the

pNa+ affect anxiety-like behavior and HPA axis activation in laboratory mice. Mice were

rendered mildly hypernatremic via systemic administration of 2.0 M NaCl, and subsequently,

were subjected to psychogenic stress or tests of anxiety-like behavior. Administration of 2.0 M

NaCl produced a modest but significant increase in the pNa+ relative to control injection of 0.15

M NaCl. This modest rise in the pNa+ was associated with attenuated anxiety-like behavior and

decreased stress-induced HPA activation. Both quantitative and qualitative neuroanatomical

studies were conducted to provide insight towards neural mechanisms contributing to the

anxiolytic effects of acute mild hypernatremia. Collectively, the results demonstrate that the

stress limiting and anxiolytic effects of slight elevations in the pNa+ also occur in mice. The

implication is that acutely increasing the pNa+ may trigger interactions between neurons

expressing oxytocin and CRH to limit responding to psychological stress.

Materials and Methods

Animals

Studies examining the effects of acute hypernatremia on anxiety-like behavior and HPA

activation used adult male C57BL/6 mice obtained from Harlan. Neuroanatomical studies

utilized CRH reporter mice generated and housed as described in Chapter 2. Standard mouse

chow (Harlan) was suspended in a wire rack that also supported an accessory water bottle

allowing ad libitum access to both food and water except where otherwise noted. All procedures

were approved by the Institutional Animal Care and Use Committee of the University of Florida.

Page 55: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

55

Restraint Stress and Blood Sampling

Mice were injected subcutaneously with 0.1 mL of either 2.0 M (n=10) or 0.15 M NaCl

(n=10) and returned to their home cages where water was made unavailable. Saline injections

were preceded by 2% lidocaine (~0.01 mL) to minimize discomfort. Sixty-minutes after saline

injections, mice were placed in clear plastic ventilated tubes to initiate a stress response in the

context of normal or elevated pNa+. Tail blood samples (~20 µL) were collected in chilled

EDTA-coated plastic collection tubes immediately at the onset of restraint and again after 30 min

of immobilization in plastic restrainers. Mice were then released and allowed to recover in their

home cages where two more blood samples were taken at 60 min and 120 min relative to the

initiation of restraint. Blood samples were kept on ice until centrifuging at 4° C at 6500 rpm for

15 min. Microcapillary samples were measured for hematocrit, and plasma was extracted and

stored at -80° C until pNa+, plasma proteins, and CORT analyses took place. Plasma sodium

levels were determined for the blood sample taken at the onset of restraint using an auto flame

photometer as previously described (Frazier et al., 2013) (Instrumentation Laboratory,

Lexington, Massachusetts). Plasma CORT was determined for each time point a blood sample

was taken using an 125I RIA kit (MP Biomedicals, Santa Ana, California) as previously described

(Frazier et al., 2013). Plasma proteins and hematocrit were determined for the blood sample

taken at the onset of restraint using a handheld refractometer (VET 360, Reichert) and

microcapillary reader, respectively.

Immunohistochemistry

Two cohorts of CRH-reporter mice (n = 6 and 8) were each further divided into groups

given either 2.0 M NaCl or 0.15 M NaCl and then restrained 60 min later as described above.

Mice were sacrificed 120 min after the onset of restraint (180 min after injections) and

stimulated Fos induction, a marker of neuronal activity, is known to peak during this time

Page 56: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

56

(Hoffman et al., 1993). Mice were overdosed with sodium pentobarbital inducing a level of

anesthesia that rendered them unresponsive to toe-pinch before transcardial perfusion with 0.9%

saline. Following the clearing of blood, mice were perfused with 4% PFA and brains were

carefully extracted then postfixed for 2 hours in 4% PFA before cryoprotection in 30% sucrose.

Four series of coronal 30 µm brain sections were cut on a Leica CM3050 S cryostat (Leica,

Buffalo Grove, Illinois) and then stored in cryoprotective solution at -20° C.

Rinses and solutions were made using 50 mM potassium phosphate buffered saline and

took place at room temperature on an orbital shaker unless otherwise noted. Immunofluorescent

labeling of Fos in brain sections from CRH-reporter mice began by rinsing free-floating sections

5 X 5 min to remove cryoprotectant. Blocking consisted of 2% normal donkey serum (Jackson

ImmunoResearch, West Grove, Pennsylvania) with 0.2% Triton-X (Sigma) for 1 h followed by

primary antibody incubation with rabbit anti-Fos (sc-52 1:1000; Santa Cruz) in blocking solution

overnight at 4° C. The second day consisted of rinses 5 X 5 min and incubation for 2 h in

blocking solution with donkey anti-rabbit Alexa-Fluor 647 (1:500; Jackson ImmunoResearch,

West Grove, Pennsylvania). After a final series of rinses, sections were mounted on Superfrost

Plus slides (Fisher) in KPB, allowed to air dry, and then coverslipped using polyvinyl alcohol

with DABCO (Sigma).

CRH-reporter mice were used for double-immunofluorescent labeling of oxytocin and

Fos. The protocol used was identical to the one used above to label Fos except for the addition

of primary antibody (mouse anti-oxytocin neurophysin, PS-38 1:400; generously provided by Dr.

Gainer, NIH) (Ben-Barak et al., 1985) to the blocking solution containing Fos antibody on day

one. Additionally, the secondary antibody, donkey anti-mouse Alexa-Fluor 488 (1:500; Jackson

Page 57: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

57

ImmunoResearch, West Grove, Pennsylvania), was mixed with donkey anti-rabbit Alexa-Fluor

647 in blocking solution on day two.

Image Capture and Analysis

All images were captured using an AxioImager M.2 fluorescent microscope (Carl Zeiss,

Thornwood, New York) connected to a PC running Axiovision 4.8. Using a Plan-Apochromat

10x/0.45 M27 objective, z-stacks of tdTomato and Fos expression were captured through the

PVN (from bregma -0.46mm to -1.22mm) using anatomical landmarks found in The Mouse

Brain in Stereotaxic Coordinates 3rd Ed.(Franklin and Paxinos, 2008). Image size was 1388 X

1040 pixels and each z-step was 1 µm with an average of 20 optical sections per PVN image.

Excitation/emission spectra used to image tdTomato and Fos were 540/580 nm and 649/670 nm,

respectively. Exposure time was automatically set by the software and varied from 400-600 ms

for the channel capturing tdTomato images and 1-3 s for the channel capturing Fos images.

Quantification of Fos positive tdTomato neurons took place by first opening each z-stack in

Axiovision 4.8 and then marking colabeled neurons using an event marker function. To verify

that Fos-positive nuclei were within tdTomato neurons, z-stacks were scrolled through and

channels were turned on and off as needed. Fos counts were performed by personnel blind to

treatment conditions on matched sections from both sides of the PVN and then averaged for

animals injected with 0.15 M NaCl and 2.0 M NaCl. Image capture and analysis of Fos

positive oxytocin neurons was performed using the same methods. Visualization of oxytocin

used an excitation/emission of 470/509 and exposure of ~700 ms while parameters for Fos were

similar to those described above.

Behavioral Testing

Naïve C57BL/6 mice were injected with 2.0 M NaCl (n=13) or 0.15 M NaCl (n=12) and

water deprived for 60 min before assessment of anxiety-like behavior on an EPM. During the

Page 58: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

58

light phase, mice were brought one at a time into a procedure room with a black curtain which

separated the EPM from the experimenters. Testing began by placing a mouse in the center of an

orthogonally-oriented two plank maze (62 cm X 62 cm) facing an open arm. A 5 min period of

exploration was recorded by a ceiling-mounted video camera connected to a PC running

TopScan software (TopScan; CleverSys, Reston, Virginia). Simultaneous video tracking of the

mouse’s position allowed for the automated scoring of time spent in the open arms and total

distance traveled. The testing apparatus was cleaned with 30% ethanol between subjects.

Statistics

All data presented as mean +/- SEM. Plasma sodium, hematocrit, plasma proteins, EPM

data, Fos positive oxytocin cells and Fos positive CRH cells were assessed with a 2-tailed t test.

Plasma CORT was analyzed using a 2-factor ANOVA using GraphPad Prism version 5.04 for

Windows (GraphPad Software; San Diego, California).

Results

Subcutaneous Delivery of 2.0 M NaCl Modestly Increases pNa+ but Blunts Restraint-

Induced Elevations in Plasma CORT

Analysis of blood samples taken from mice given injections of 2.0 M NaCl found a slight

but significant increase in pNa+ relative to controls injected with 0.15 M NaCl (Fig. 3-1A).

Hematocrit and plasma protein measurements were similar between hypernatremic and control

mice.

Plasma CORT analysis revealed a significant time X condition interaction [Fig. 3-2,

F(3,39) = 3.48, (P<0.05)]. Restraint increased CORT at 30 min compared to baseline

concentrations in both groups; however, mice injected with 2.0 M NaCl had an attenuated rise at

30 and 60 min compared with mice injected with 0.15 M NaCl. This attenuation was specific to

these time points as both baseline and 120 min CORT concentrations were not significantly

Page 59: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

59

different between groups. With respect to interpolated time points, the area under the curve was

also decreased relative to controls [Fig. 3-2 inset, (P<0.05)].

Acute Hypernatremia Inversely Affects Restraint-Induced Activation of CRH and

Oxytocin Neurons in the PVN

Figure 3-3 shows photomicrographs of atlas matched sections through the PVN

containing tdTomato reporting of CRH positive neurons and immunofluorescent labeling of Fos.

Whereas control mice (Fig. 3-3A) expressed robust restraint-induced Fos expression in CRH

neurons, this same region displayed fewer activated CRH neurons following 2.0 M NaCl

injection (Fig. 3-3B). Compared with mice injected with 0.15 M NaCl, mice injected with 2.0 M

NaCl had significantly less (P<0.05) restraint-induced Fos positive CRH neurons (Fig. 3-3C).

Conversely, mice injected with 2.0 M NaCl and restrained (Fig. 3-4B) exhibited a significant

[Fig. 3-4C, (P<0.05)] increase in Fos induction in oxytocin neurons relative to controls (Fig. 3-

4A).

Slight Elevations in the pNa+ are Associated with Decreased Anxiety-Like Behavior in the

EPM

Mice injected with 2.0 M NaCl spent more time exploring the open arms of an EPM than

mice injected with 0.15 M NaCl [Fig. 3-5A, (P<0.05)]. This increased exploration was not due

to an overall increase in locomotion as the total distances traveled were similar between groups

(Fig. 3-5B).

Discussion

The goal of the current study was to determine whether the anxiolytic and stress

dampening effects of acute mild hypernatremia occur in laboratory mice. To this end, mice

systemically delivered 2.0 M NaCl had a modest but significant increase in the pNa+ relative to

control mice treated with 0.15 M NaCl. The slight rise in the pNa+ that was observed in mice

administered 2.0 M NaCl was associated with an attenuated restraint-induced activation of the

Page 60: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

60

HPA axis and increased time spent in the open arms of an EPM. We conducted neuroanatomical

studies to evaluate how acute hypernatremia may interact with restraint-stress to affect Fos

induction, a marker of neuronal activation, in PVN neurons expressing oxytocin or CRH. Acute

hypernatremia elicited robust Fos induction within oxytocin neurons but significantly decreased

Fos within CRH neurons. Our results, in conjunction with previous studies, suggest that acute

hypernatremia dampens stress-responsiveness, in part, by promoting oxytocin-mediated

inhibition of CRH neurons. The implication is that increased salt intake may be used as a coping

strategy to alleviate the impact of psychological stressors.

Mice rendered hypernatremic via systemic administration of 2.0 M NaCl had a slight

(≈2%) but significant increase in the pNa+ relative to control mice delivered 0.15 M NaCl.

While this increase stimulates oxytocin and vasopressin release as well as osmotically driven

water intake (Stricker and Verbalis, 1986; Krause et al., 2008), the elevation in the pNa+

observed in our study is modest relative to the order of magnitude (10% increase) that is required

to increase brain osmolytes to defend the CNS against the damaging effects of severe

hypernatremia (Heilig et al., 1989). Our experimental paradigm, does however, produce an

increase in the pNa+ that is strikingly similar to that which occurs in humans and animals given

excess dietary sodium to cause increases in blood pressure or the development of salt-sensitive

hypertension, respectively (He et al., 2005) (de Wardener et al., 2004) (Fang et al., 2000;

O'Donaughy and Brooks, 2006; O'Donaughy et al., 2006). Thus, our experimental model of

hypernatremia produces a modest increase in the pNa+ that likely affects cardiovascular function

as well as endocrine and behavioral responses to psychogenic stress.

Activation of the HPA axis is an endocrine measure of stress-responsiveness that is

initiated by excitation of CRH neurons in the PVN which stimulate ACTH secretion from the

Page 61: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

61

anterior pituitary, which in turn, promotes CORT release from the adrenal cortex. Relative to

isonatremic controls, mice rendered mildly hypernatremic had decreased plasma CORT

subsequent to the onset of restraint. These results are in agreement with our previous studies

demonstrating that rats subjected to a similar degree of hypernatremia have decreased plasma

CORT subsequent to a restraint challenge (Krause et al., 2011a; Frazier et al., 2013). Our

previous study also determined that acute hypernatremia decreases the secretion of ACTH into

the systemic circulation (Krause et al., 2011a). Because the release of ACTH is controlled by

excitation of CRH neurons in the PVN, we hypothesized that acute hypernatremia may attenuate

Fos induction in PVN neurons producing tdTomato, a marker for CRH expression. Importantly,

mRNA for CRH was present in ≈ 95% of all tdTomato producing neurons in the PVN (Fig. 2-1),

thereby demonstrating the validity of CRH-reporter mice. Consistent with our hypothesis,

restraint-evoked Fos induction of tdTomato producing neurons was significantly decreased in

mice subjected to acute hypernatremia relative to controls. Taken together, these results suggest

that acute hypernatremia dampens stress-induced activation of the HPA axis by inhibiting CRH

neurons in the PVN.

Hypernatremic mice had strong Fos induction in oxytocin neurons in the PVN relative to

isonatremic controls and these results are consistent with those from our previous research using

rats to demonstrate that acute hypernatremia combined with restraint more than tripled the

number of Fos positive oxytocin neurons in the PVN compared to restraint alone (Krause et al.,

2011a). Work conducted by Ludwig and colleagues established that systemic hypernatremia

triggers the dendritic release of oxytocin, which produces robust and sustained elevations in

central levels of this peptide (Ludwig and Leng, 2006). Given that exogenous administration of

oxytocin decreases stress-induced activation of the HPA axis (Windle et al., 1997; Windle et al.,

Page 62: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

62

2004), it is possible that acute hypernatremia triggers endogenous release of oxytocin within the

CNS, which inhibits activation of CRH neurons, and consequently, blunts HPA activity. In

support of this, electrophysiological studies conducted by Frazier and colleagues (Frazier et al.,

2013) revealed that acute hypernatremia created an inhibitory tone on putative CRH neurons that

was dependent on activation of oxytocin receptors. Therefore, elevating the pNa+ may cause

activation of osmosensitive oxytocin neurons in the hypothalamus, which influence excitation of

neurons responsive to psychogenic stress by releasing oxytocin into the CNS.

Hypernatremia and restraint likely utilize different neural circuits to activate the PVN and

both are known to elicit robust Fos induction in this nucleus (Sharp et al., 1991; Herman and

Cullinan, 1997). Our previous study found that hypernatremia followed by restraint significantly

increased Fos in oxytocin and vasopressin containing neurons; however, the total number of cells

expressing Fos within the PVN was similar to that of isonatremic animals subjected to restraint

(Krause et al., 2011a). These results suggested that the effects of hypernatremia and restraint

were not additive. Given that hypernatremia blunted restraint-induced activation of the HPA

axis (Krause et al., 2011a), we hypothesized that hypernatremia activated oxytocin and

vasopressin containing neurons, but inhibited CRH containing neurons in the PVN. Consistent

with this hypothesis, the present study determined that hypernatremia followed by restraint

significantly decreased Fos induction in tdTomato expressing neurons in the PVN. Taken

together, the result suggest that hypernatremia selectively activates oxytocin and vasopressin

containing neurons in the PVN but inhibits those that express CRH, which accounts for the

similarities in total Fos induction observed between hypernatremic and isonatremic animals

subjected to restraint.

Page 63: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

63

In addition to attenuating stress-induced activation of the HPA axis, the present study

found that acute hypernatremia decreased anxiety-like behavior in the EPM relative to that of

isonatremic controls. Once again, these results are consistent with those from our previous

studies using rats that found that slight elevations in the pNa+ are anxiolytic, especially in social

situations (Krause et al., 2011a; Frazier et al., 2013). Central administration of oxytocin

decreases anxiety-like behavior in the EPM (Mak et al., 2012) and oxytocin efferents are found

in limbic brain regions (Knobloch et al., 2012) like the amygdala, that are heavily implicated in

the expression of fear and anxiety-like behavior (Davis et al., 1994). Of relevance, some of the

oxytocin efferents in the amygdala arise from magnocellular oxytocinergic neurons in the PVN

and supraoptic nucleus (Knobloch et al., 2012). Magnocellular oxytocinergic neurons are

osmosensitive and become excited by elevations in the pNa+ (Hattori et al., 1990). Therefore, it

is possible that the anxiolytic effects of acute hypernatremia result from excitation of

osmosensitive oxytocin neurons with axonal projections to limbic brain nuclei controlling the

expression of fear and anxiety-like behavior.

Collectively our results demonstrate that the anxiolytic and stress dampening effects of

acute hypernatremia extend to mice. Consequently, these results allow the use of mice and the

genetic manipulations that this animal model affords to further investigate the neural and

humoral mechanism(s) underlying the stress limiting effects of acute hypernatremia. From a

broader perspective, our results may provide insight as to why some patients with salt-sensitive

hypertension and high levels of life-stress (Calhoun, 1992; Calhoun and Oparil, 1995; Barnes et

al., 1997) also have difficulties complying with restricted dietary sodium intake (Epstein et al.,

2012). These more far-reaching conclusions are discussed further in Chapter 6.

Page 64: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

64

Figure 3-1. Plasma measurements from mice 60 min after 2.0 M or 0.15 M NaCl injections. (A)

Injections of 2.0 M NaCl significantly increased pNa+ relative to control injections of

0.15 M NaCl, but had no effect on (B) hematocrit or (C) plasma proteins. *p<0.05

Error bars indicate SEM.

Page 65: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

65

Figure 3-2. 2.0 M NaCl administration reduces the CORT response to restraint stress relative to

0.15 M NaCl. The integrated CORT response was also significantly reduced.

*p<0.05 0.15 M NaCl vs 2.0 M NaCl. AUC=Area Under the Curve. Error bars

indicate SEM.

Page 66: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

66

Figure 3-3. 2.0 M NaCl injection attenuates restraint-induced activation of CRH neurons. (A)

Representative photomicrograph of a unilateral coronal section through the PVN

depicting Fos induction (cyan nuclei) in tdTomato (red soma) containing neurons

following 0.15 M NaCl and restraint. (B) Representative photomicrograph of a

unilateral coronal section through the PVN depicting Fos induction in tdTomato

containing neurons following 2.0M NaCl and restraint. (C) The group mean for Fos

induction in tdTomato containing neurons was significantly less for mice subjected to

2.0 M NaCl injection and restraint relative to control. *p<0.05 Scale bars = 50 m.

Page 67: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

67

Figure 3-4. 2.0 M NaCl administration and restraint increases activation of oxytocin-producing

cells in the PVN. (A) Representative photomicrograph of a unilateral coronal section

depicting Fos induction (red nuclei) in oxytocin (green cell bodies) containing

neurons following 0.15 M NaCl and restraint. (B) Representative photomicrograph

of a unilateral coronal section depicting Fos induction in oxytocin neurons following

2.0 M NaCl and restraint. (C) The group mean for Fos induction in oxytocin-

producing cells was significantly more for mice subjected to 2.0 M NaCl injection

and restraint relative to control. *p<0.05 Scale bars = 50 m.

Page 68: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

68

Figure 3-5. Acute hypernatremia attenuates anxiety-like behavior. (A) Mice treated with 2.0 M

NaCl spent a greater proportion of time exploring the open arms of an EPM than mice

treated with 0.15 M NaCl. (B) Overall locomotion was unaffected by 2.0 M NaCl as

total distance traveled was similar for each group. * p < 0.05. Error bars indicate

SEM.

Page 69: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

69

CHAPTER 4

CENTRAL AND BEHAVIORAL EFFECTS OF CHRONIC SALT-LOADING

Background

Chronic salt-loading through the replacement of drinking water with concentrated sodium

chloride solution has been used experimentally in rats and mice to investigate factors that

accompany persistently hyperactive osmoregulatory mechanisms. Specifically, an initial period

of intracellular dehydration during the first 24 hrs stimulates a variety of physiological

alterations to maintain normal fluid intake and plasma volume with increased urinary output

(Watts et al., 1995; Polito et al., 2006). These alterations along with increased central and

systemic vasopressin (Amaya et al., 1999) secretion promote water retention and the excretion of

excessive amounts of sodium. However, following three days of drinking 2% NaCl, rats fail to

maintain adequate hydration of the extracellular fluid compartment which is apparent by an

increasing hematocrit (Watts et al., 1995). Without sufficient hydration rats become increasingly

aphagic and lose weight although fluid intakes range from hyperdipsic to adipsic as homeostasis

continues to be unattainable (Watts et al., 1995). Furthermore, after an initial period of

heightened CORT (Watts, 1992a), ACTH and CORT are suppressed to lower than control levels

and restraint-induced CORT is suppressed as well (Amaya et al., 2001). Neuronal activation

increases in diverse brain areas and neuronal phenotypes as a result of chronic salt-loading

(Watts, 1992a; b; Watts et al., 1995; Watts, 1996). Overall Fos activation increases in the

OVLT, median preoptic nucleus, SFO, SON, and PVN as these regions sense and relay

hyperosmotic plasma conditions to brain areas responsible for hydromineral homeostasis

(Kovacs and Sawchenko, 1993). However, there have been inconsistencies reported across

species as Fos expression in the SON of the rat (Kovacs and Sawchenko, 1993) was not found to

be expressed in similarly salt-loaded mice (St-Louis et al., 2012). Additionally, accounting for

Page 70: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

70

Fos expression throughout the entire rostral-caudal extent of the PVN has not been done, nor has

Fos expression in the PVN been assessed with the combined stressors of chronic salt-loading and

restraint. Finally, in contrast to several studies in rats, restraint-induced CORT has yet to be

determined in chronically salt-loaded mice. Thus, during the final period of chronic salt-loading

a variety of neural, humoral, and autonomic processes caused by prolonged hyperosmotic

conditions and which are putatively involved in the pathogenesis of hypertension and

cardiovascular disease can be characterized.

Methods

Animals

Studies examining the effects of chronic hypernatremia on HPA activation used adult

male C57BL/6 mice obtained from Harlan. Neuroanatomical studies utilized adult male CRH

reporter mice which were generated as detailed in Chapter 2. All mice either continued ad

libitum access to water (control group; Water) or had water replaced with a 2% NaCl solution for

five days (experimental group; 2% NaCl). All procedures were approved by the Institutional

Animal Care and Use Committee at the University of Florida.

Procedure

Assessment of food and fluid intake was performed at the same time each day for the

duration of the 5 d study. Weights of both chow and water- or saline-filled bottles were recorded

by an individual experimenter and are presented as mean daily intakes corrected for body weight.

Body weights were assessed during the same monitoring session and are presented as daily

averages. Tail blood samples (~20 µL) were collected from mice in both groups using chilled

EDTA-coated plastic collection tubes immediately before water was replaced with saline.

Furthermore, microcapillary tubes were used to sample blood which was immediately

centrifuged to measure hematocrit. After the fifth consecutive day of salt-loading, a final tail

Page 71: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

71

blood sample was taken from mice in both groups upon initiation of restraint and hematocrit was

again assessed. Restraint was accomplished by placing mice in clear plastic ventilated restraint

devices constructed of 50 mL conical tubes for 30 min and allowed a 60 min recovery period.

Mice were then overdosed with sodium pentobarbital and transcardially perfused with 0.9%

saline (30-50 mL) followed by ice-cold 4% PFA (30-50 mL). Brains were carefully extracted

and postfixed for 4 h in 4% PFA before cryoprotection in 30% sucrose. Four series of coronal

30 µm sections were cut on a Leica CM3050 S cryostat (Leica, Buffalo Grove, Illinois) and then

stored in cryoprotective solution at -20° C. Blood samples were kept on ice until centrifuging at

4° C at 6500 rpm for 15 min. Plasma was extracted and stored at -80° C until pNa+ levels were

determined using an auto flame photometer (Instrumentation Laboratory, Lexington,

Massachusetts) and plasma protein could be determined using a hand-held refractometer

(Reichert, Depew, New York).

Restraint Stress and CORT

Mice were placed in clear plastic ventilated tubes to initiate a stress response in the

context of normal or elevated pNa+. Tail blood samples (~20 µL) were collected in chilled

EDTA-coated plastic collection tubes immediately at the onset of restraint and again after 30 min

of immobilization in plastic restrainers. Mice were then released and allowed to recover in their

home cages where two more blood samples were taken at 60 min and 120 min relative to the

initiation of restraint. Blood samples were kept on ice until centrifuging at 4° C at 6500 rpm for

15 min and stored at -80° C CORT analyses took place. Plasma CORT was determined for each

time point a blood sample was taken using an 125I RIA kit (MP Biomedicals, Santa Ana,

California) as previously described in Chapter 3.

Page 72: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

72

Image Capture and Analysis

Images of the PVN were captured using an AxioImager M.2 fluorescent microscope

(Carl Zeiss, Thornwood, New York) equipped with an Apotome.2 and connected to a PC

running Axiovision 4.8. Using a Plan-Apochromat 20x/0.8 M27 objective, z-stacks of tdTomato

and fluorescently-labeled vasopressin-NP or Fos expression were captured from sections

corresponding to various rostral-caudal levels through the PVN (from Bregma -0.46 mm to -1.22

mm). Each image was assigned to a specific atlas figure by an individual experimenter that

analyzed the shape of the PVN proper and matched it with the shapes of adjacent anatomical

landmarks found in The Mouse Brain in Stereotaxic Coordinates 3rd Ed (Franklin and Paxinos,

2008). Image size was approximately 1388 X 1040 pixels and each z-step was 1 µm with an

average of 20 optical sections per PVN image. When capturing images from sections processed

for immunohistochemistry, exposure time was determined based on an overexposure reporting

function automatically set by the software. Exposure time during the imaging process varied

from 30-40 ms for the channel capturing tdTomato images, 125-175 ms for vasopressin-NP

images, and 400-800 ms for Fos images. Images were adjusted using the brightness and contrast

settings available in Axiovision to increase clarity.

Quantification of colocalized tdTomato or vasopressin-NP with Fos positive nuclei took

place using a method similar to the “direct three-dimensional counting” process (Williams and

Rakic, 1988). Each z-stack was opened in Axiovision 4.8 or Zen lite 2012 (Carl Zeiss) and then

each image was inspected while alternately turning the respective channels on and off to identify

potential colocalizations. Candidate cells were then further analyzed as necessary by removing

the area of the image containing the cell into a separate region of interest (ROI) image to

generate a 3D reconstruction. If, as a 3D reconstruction, the cell contained fluorescent labeling

on each individual channel throughout the z-plane of the image, then the cell was counted. The

Page 73: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

73

number of bi-lateral colocalizations were determined for each atlas-matched section containing

the PVN. Statistical analyses were performed by grouping according to atlas region and in a

separate analysis by grouping all images obtained from each subject.

Statistics

All grouped data presented as mean +/- SEM. Fluid intake, food intake, body weight,

total Fos, percentage of Fos positive tdTomato neurons, and plasma CORT were assessed with a

two-factor analysis of variance. Main effects or interactions (P<0.05) were assessed post hoc

with the Bonferroni method. Within group (Pre vs post salt-loading) and between group pNa+,

hematocrit, and plasma protein percentage were each assessed with a paired and unpaired 1-

tailed t-test, respectively. Significance for all analyses was set at P<0.05. Statistical analyses

were performed and graphs were created using GraphPad Prism 5.

Results

Salt-Loaded Mice Exhibit Increased Fluid Intakes and pNa+ Concentrations, but Maintain

Normal Indices of Blood Volume

Daily monitoring of fluid intake revealed that mice in the control condition drank an

average of 5.27+/-0.10 mL of water each day and intake did not significantly change over the

course of the study. In contrast, mice with 2% NaCl as a sole source of fluid drank significantly

more after 48 h of salt exposure (time x treatment interaction; [F(4, 115) = 3.14, (P<0.05)]) and

intake remained significantly (P<0.0001 on days 2, 3, and 4; P<0.001 on day 5) higher than

controls for each day thereafter (Fig. 4-1, A). Consumption of 2% NaCl for five days

significantly elevated the pNa+ relative to Day 1 and compared with water drinking controls

(P<0.05) (Fig. 4-1, B). Volume of fluid consumed after the first day was similar between

groups, however, mice drinking 2% NaCl exhibited highly variable intakes after the second day

(Fig. 4-2, A). Despite the challenge to hydromineral homeostasis, hematocrit and plasma protein

Page 74: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

74

concentrations were similar between groups both at the beginning and the end of the experiment

(Fig. 4-1, C & D).

Salt-Loaded Mice Maintain Food Intake and Body Weight

Analysis of chow consumption revealed a main effect of 2% NaCl [F(1, 115) = 9.53,

(P<0.05)]; however, post hoc analysis did not reveal significant differences during any particular

24 period (Fig. 4-4, A). During the first 24 hs, salt-loaded mice showed no correlation

(R2=0.0065, P>0.05) between food and fluid intake (Fig. 4-3, A). In contrast, there was a

positive correlation (R2=0.5393; P <0.05) observed in water-drinking controls after the first 24

hrs. Following the second 24 hrs of salt-loading this trend changed (Fig. 4-3, B) with a

significant positive correlation emerging in salt-loaded mice (R2=0.3283; P <0.05). Water-

drinking controls failed to show any correlation between food and fluid intake for the remainder

of the experiment. Figure 4-4B shows both groups maintained stable daily average body weights

throughout the five-day period.

Salt-Loading Does Not Increase Neuronal Activation in PVN Vasopressin Neurons or the

SON

Figure 4-5 depicts the Fos immunoreactivity and vasopressin-NP expression in the PVN

of Water (Fig. 4-5, A) and 2% NaCl (Fig. 4-5, B) mice subjected to restraint. Fos expression

was not evident in vasopressin-NP neurons in the PVN of control mice and only 7.35 +/- 4.52%

of vasopressin-NP neurons were Fos positive in the PVN of mice drinking 2% NaCl (P=0.12).

Furthermore, within the SON neither group was found to express Fos (Fig. 4-5; C, D).

Salt-Loading Increases Fos Expression in the Posterior PVN

Total Fos counts for PVN “neurosecretory” regions (Bregma -0.70mm and -0.82mm) and

“preautonomic” regions (Bregma -0.94mm and -1.06mm) were averaged and compared for

groups of mice drinking water, 2% NaCl, water + restraint, and 2% NaCl + restraint (Fig. 4-5,

Page 75: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

75

E). In the “neurosecretory” regions, restraint increased the number of Fos-positive nuclei

independent of saline treatment [F(1,24) = 58.13, P <0.05; Fig. 4-5E]. In “preautonomic” regions,

there was an interaction between salt-loading and restraint [F(1,24) = 44.09, P <0.05; Fig. 4-5E].

In addition to restraint-induced Fos increases, mice subjected to salt-loading alone exhibited

increased Fos immunoreactivity [Main effect of 2% NaCl: F(1,24) = 4.53, P <0.05; Fig. 4-5E]

compared to water-drinking controls.

Salt-Loading does not Affect Restraint-Induced Fos Induction in tdTomato Neurons

Figure 4-5F shows the average percentage of Fos-positive tdTomato neurons in the PVN.

The average percentage of Fos-positive tdTomato neurons in the PVN was similar in

unrestrained mice drinking water or 2% NaCl. The percentage of Fos-positive tdTomato neurons

in both PVN regions was significantly increased by restraint [F(1,44) = 65.89, P<0.05] in both

euhydrated and salt-loaded mice. There was no main effect of salt-loading or interaction

between salt-loading and restraint.

Salt-Loading Blunts Restraint-Induced Elevations in Plasma CORT During Recovery

Plasma CORT analysis revealed a significant time X treatment interaction [Fig. 4-6,

F(3,54) = 3.32, (P<0.05)]. Restraint increased CORT at 30 min compared to baseline

concentrations in both groups; however, salt-loaded mice recovered faster as CORT at 120 min

was significantly reduced (P<0.05). Attenuation was specific to this time point as both baseline

and 60 min CORT concentrations were not significantly different between groups. With respect

to interpolated time points, the area under the curve was not different relative to controls (Fig. 4-

6 inset).

Discussion

We evaluated how salt-loading caused by consumption of 2% NaCl changes ingestive

behavior, neuronal activation, and restraint-induced CORT. The behavioral adaptations that

Page 76: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

76

occurred to counter the osmotic stress included significant increases in daily saline intake that

correlated with the maintenance of chow intake. Relative to water drinking controls, salt-loading

significantly increased the pNa+ but had no effect on hematocrit or plasma protein

concentrations. Fos induction within vasopressin-NP and tdTomato expressing neurons was

similar between salt-loaded and control mice and no Fos induction was evident in the SON;

however, total Fos increased in the posterior preautonomic region of the PVN with salt-loading

alone. These measures of neuronal activation did not predict similar HPA axis functionality as

salt-loaded mice had CORT levels significantly lower at 120 min after the onset of restraint.

In the current study, drinking 2% NaCl instead of water robustly increased fluid intake

but did not substantially affect food consumption. The behavioral adaptations of salt-loaded

mice were accompanied by a significant rise in pNa+, but did not elicit changes in body weight,

hematocrit or plasma proteins. In contrast to our results here using a mouse model, rats drinking

saline instead of water exhibit highly variable fluid intake and decreased food consumption

resulting in significantly reduced blood volume and body weight (Watts, 1992a; Watts et al.,

1995; Yue et al., 2008; Boyle et al., 2012). The renal handling of sodium and water seems to be

more efficient in mice relative to rats (Zhai et al., 2006; Christensen et al., 2014). Whereas rats

respond to forced saline consumption by limiting food intake, mice quickly adapt to maintain

intake levels by drinking and excreting large volumes of fluid. Consistent with this, the higher

ratio of short to long loop nephrons in the mouse kidney (81:18, (Zhai et al., 2006)) compared

with rats (70:30, (Kriz, 1967)) is predictive of the lower rise in plasma osmolality observed in

mice (3.3%, (Polito et al., 2006)) compared with rats (6.5%, (Watts et al., 1995)) following five

days of drinking 2% NaCl. Interestingly, the ratio of short to long loop nephrons in humans

(81:15, (Peter et al., 1927)) is closer to the mouse kidney than that of the rat, suggesting more

Page 77: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

77

similarities between the human and murine renal handling of sodium than with the rat. Thus,

relative to rats, it seems that mice are less susceptible to the hypernatremia that occurs with salt-

loading which may explain why dehydration anorexia was limited in the present study.

During the first 24 hrs, mice drinking 2% NaCl instead of water did not have fluid intake

that correlated with food consumption, but subsequently, a positive correlation emerged

suggesting the engagement of compensatory mechanisms that allow normal food intake by

alleviating intracellular dehydration. Our experiments evaluating Fos induction may provide

insight towards such mechanisms. We found that Fos induction within vasopressin neurons was

not affected by five days of salt-loading and these results are consistent with those of (St-Louis et

al., 2012) which determined that magnocellular vasopressin neurons are activated after two days

of salt-loading but this activation subsided six days later. Taken together, these results suggest

that in mice the activity of vasopressin neurons may return to basal levels earlier than previously

thought and activation of neurosecretory vasopressin neurons may be transient while vasopressin

mRNA and hnRNA continue to rise with enduring hypernatremia (Ozaki et al., 2004; St-Louis et

al., 2012).

The differences in Fos expression observed in the PVN of chronic salt-loaded mice may

be indicative of neuroendocrine and autonomic adaptations aimed at coping with chronic osmotic

stress. The results from Chapter 3 indicate that acute hypernatremia blunts HPA axis activation

elicited by psychogenic stress, an effect that was observed here 90 min after chronically salt-

loaded mice were released from restraint tubes. Basal CORT levels were not affected by acute

or chronic salt-loading. Our results in Chapter 3 also indicate that acute salt-loading is

associated with decreased Fos-induction within CRH neurons in the PVN; however, our results

in the chronic salt-loading paradigm indicate that this decrease does not occur in CRH neurons of

Page 78: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

78

the PVN. Specifically, salt-loaded and control mice exhibited similar restraint-induced Fos

expression within tdTomato neurons in the PVN. Interestingly, salt-loading alone increased total

Fos in the posterior preautonomic region of the PVN; however, this effect also cannot be

attributed to increased activation of tdTomato neurons within the posterior preautonomic region

because colocalization of Fos and tdTomato was similar in salt-loaded and water-drinking mice.

These results suggest that the increased sympathetic drive that occurs with chronic salt-loading

may not derive from an increase in the number of active preautonomic CRH neurons as

detectable by Fos, but rather, may be the result of altered hypothalamic expression of

neuropeptides, like CRH and vasopressin, which allows modulation of neuronal activity in

hindbrain nuclei that control sympathetic outflow.

Collectively these experiments evaluated in a broad way the murine response to salt-

loading. In agreement with previous studies, mice respond to the presentation of only saline as a

source of water by drinking large volumes after an initial period of somewhat normal intakes.

However, in stark contrast to similar studies in rats, mice maintained food intake and body

weights with consumptions of fluid and food that were positively correlated with one another

after the first day. Furthermore, while the acute salt-loading experiment in Chapter 3 indicated

that restraint-induced CORT is suppressed at 30 and 60 min following the onset of restraint, the

chronic salt-loading experiment here indicated no effects at these time points and suppression

with salt-loading at 120 min. Analysis of the neuronal activation in the PVN and SON that may

have contributed to the adaptive ingestive behaviors included lack of evidence for magnocellular

vasopressin activity at the time of tissue collection, but rather an indication that the activity of

preautonomic neurons may be involved in the response to salt-loading. Further investigation into

Page 79: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

79

the potential for neuropeptide plasticity and preautonomic involvement are explored in Chapter

5.

Page 80: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

80

Figure 4-1. Fluid consumption and the effects on indices of hydration. Mice with access to 2%

NaCl (shaded circles) instead of water (open squares) drank significantly more (time

x treatment interaction; [F(4, 115) = 3.14, (P<0.05)]) each day after the first 48 h (A).

Mice maintained on 2% NaCl had higher pNa+ on Day 5 compared with Day 1 (B;

P<0.05). Hematocrit (C) and plasma proteins (D) were similar between groups.

Water n=12, 2% NaCl n=13.

Page 81: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

81

Figure 4-2. Individual fluid and food intake values. Circles represent the individual water (blue)

and 2% NaCl (red) daily fluid intakes values normalized to body weight for each 24

hr period (A). Bars represent the daily average of all mice. Individual daily chow

intake values normalized to body weight for mice with access to water (blue circles)

or 2% NaCl (red circles) (B). Bars represent the daily average of all mice.

Page 82: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

82

Figure 4-3. Correlation graphs for food and fluid intake. Individual values plotted for daily food

and fluid intake for mice drinking water (black squares) or 2% NaCl (red circles).

Significant correlations were determined for Water Day 1 (p=0.01; R2=0.5393) and

2% NaCl Day 2-5 [Day 2 (p=0.04; R2=0.3283); Day 3 (p=0.04; R2=0.3356); Day 4

(p=0.02; R2=0.4477); Day 5 (p=0.01; R2=0. 4876)].

Page 83: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

83

Figure 4-4. Daily food intake and body mass measurements. The consumption of chow was

similar between salt-loaded mice and water-drinking controls as there were no

significant differences in daily intakes (A), however, there was a main effect of 2%

NaCl [F(1, 115) = 9.53, (P<0.05)]. Body weights did not differ significantly (B). Water

n=12, 2% NaCl n=13.

Page 84: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

84

Figure 4-5. Stress induced neuronal activation. (top left) Coronal sections through the PVN

depicting neurons labeled with vasopressin-NP (green cell bodies) and Fos (red

nuclei). Fos expression was similar for restrained water drinking (A) and saline-

drinking mice (B). (bottom left) Coronal sections through the SON depicting Fos

labeling. Fos expression was similar for restrained water drinking (C) and saline-

drinking mice (D) in the SON. (top right) Overall Fos activation in the PVN was

increased with restraint in both the “neurosecretory” region (Bregma -0.70 to -0.82)

[F(1,24) = 58.13, P <0.05; Fig. 3-5E] and “preautonomic” region (Bregma -0.94 to -

1.06) [F(1,24) = 44.09, P <0.05; Fig. 3-5E]. Unrestrained salt-loaded mice had

increased Fos compared with unrestrained controls in the “preautonomic” region of

the PVN [Main effect of 2% NaCl: F(1,24) = 4.53, P <0.05; Fig. 3-5E]. (bottom right)

Fos activation in tdTomato neurons was increased in the PVN with restraint at both

PVN regions (P<0.05). Salt-loading did not affect Fos expression in tdTomato

neurons (F). Water n=12, 2% NaCl n=12. Error bars = SEM. Scale bars = 100 µm.

3v, 3rd ventricle. ot, optic tract.

Page 85: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

85

Figure 4-6. Chronic salt-loading reduces the CORT response to restraint during the recovery

period. The integrated CORT response was not significantly reduced. *p<0.05.

AUC=Area Under the Curve. Error bars indicate SEM.

Page 86: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

86

CHAPTER 5

CHRONIC SALT-LOADING-ASSOCIATED NEUROPEPTIDE PLASTICITY

Background

The PVN contains peptidergic neurons that respond to dehydration through endocrine

and neural compensatory mechanisms that maintain and restore hydromineral balance.

Specifically, the intracellular dehydration that occurs with acute elevations in the pNa+

influences the activation of PVN neurons to prevent diuresis but promote natriuresis and the

maintenance of blood pressure by controlling the systemic and central release of vasopressin,

oxytocin and CRH. Elevations in the pNa+ or hypernatremia activates magnocellular

vasopressin and oxytocin neurons in the PVN to elicit neurohypophyseal secretion of these

neuropeptides which act on the kidney to promote water retention and the excretion of sodium in

urine (Verbalis et al., 1991; Ludwig et al., 1994; Pirnik et al., 2004). In contrast, acute

hypernatremia inhibits neurosecretory CRH neurons and plasma renin activity resulting in

blunted stress-induced HPA axis activation and lowered levels of corticosterone (Krause et al.,

2011a), thereby attenuating renal sodium retention (Husain et al., 1987; Papanek et al., 1997).

The PVN contains preautonomic neurons that also produce vasopressin, oxytocin and CRH that

project to the hindbrain and spinal cord to control sympathoexcitation and cardiovascular

function (Lee et al., 2013). In contrast to their neurosecretory counter-parts, peptidergic

preautonomic neurons are thought to influence sympathetic nervous system activity and

cardiovascular responses to hydromineral imbalance by affecting neurotransmission through the

release of vasopressin, oxytocin and CRH from axons terminating on postsynaptic neurons.

Thus, acute elevations in body sodium levels trigger coordinated excitation and inhibition of

peptidergic neurons in the PVN to initiate endocrine and neural signals that buffer against further

dehydration while maintaining blood pressure.

Page 87: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

87

Chronic salt-loading, as a consequence of drinking hypertonic saline instead of water,

also is associated with attenuated HPA axis activity and increased sympathetic outflow and

blood pressure (Sapirstein et al., 1950; Amaya et al., 2001), and these changes occur

concomitantly with altered expression of neuropeptides in the PVN and amygdala (Watts et al.,

1995). For example, studies conducted in rats found that magnocellular neurons of the PVN and

SON adapt to chronic salt-loading by upregulating CRH (Lightman and Young, 1987; Kovacs

and Sawchenko, 1993), which may facilitate the systemic release of oxytocin and natriuresis

(Verbalis et al., 1991); however, parvocellular neurons in the PVN down-regulate CRH

expression (Amaya et al., 2001) while neurons in the amygdala increase CRH (Watts et al.,

1995). As mentioned, the neuropeptide plasticity that follows chronic salt-loading is associated

with increased osmoregulatory capacity, decreased responsiveness of the HPA axis and

sympathetically-mediated hypertension (Lightman and Young, 1987; Watts, 1992a; 1996;

Amaya et al., 2001; Hartner et al., 2003). Thus, chronic salt-loading stimulates neuropeptide

plasticity in the PVN that may serve as a compensatory mechanism that permits the altered

endocrine axis activity and enhanced sympathetic outflow that is required to cope with chronic

osmotic stress. Whether chronic salt-loading elicits neuropeptide plasticity in preautonomic

neurons of the PVN or amygdala has not been investigated.

The primary goal of this study was to characterize the effect of chronic salt-loading on

CRH plasticity within vasopressin, oxytocin, and preautonomic neurons in the PVN. For

comparative purposes, preautonomic neurons in the amygdala were analyzed. Here, we utilized

a CRH-reporter mouse to specifically focus on the neural subtypes that respond to dehydration

by upregulating CRH. Additionally, relatively few chronic salt-loading experiments have been

performed in mice and the organization of neurons in the mouse PVN has only recently been

Page 88: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

88

thoroughly characterized (Biag et al., 2012). Consequently, we also sought to analyze CRH

expression in the PVN and SON for comparative purposes with similar studies using rats.

Collectively, our results serve to solidify the structure and functional neuroanatomy of the mouse

PVN and provide evidence that a prolonged hyperosmotic stimulus can influence CRH

expression in neuroendocrine and preautonomic neurons.

Methods

Animals

CRH reporter mice were generated as detailed in Chapter 2. A group of mice (n=23)

were used to quantify the effect of chronic salt-loading on CRH colocalization with oxytocin or

vasopressin and another group of CRH reporter mice (n=33) were used to determine the extent to

which these colocalizations occurred in PVN neurons projecting to the hindbrain. Mice either

continued ad libitum access to water (control group; Water) or had water replaced with a 2%

NaCl solution for five days (experimental group; 2% NaCl). All procedures were approved by

the Institutional Animal Care and Use Committee at the University of Florida.

In situ Hybridization

Localization of CRH mRNA through fluorescent detection was performed as described in

Chapter 2. Localization of CRH mRNA without fluorescence was accomplished with an alkaline

phosphatase reaction. Following the overnight hybridization, coverslips were removed and

slides were washed 3 X 5 min in maleic acid buffer containing Tween 20 at room temperature

and 2 X 30 min in 50% formamide, 1x SSC, 0.1% Tween-20. Sections were then blocked for 1

h at room temperature with 1x Blocking Reagent (Roche) in maleic acid buffer containing

Tween 20 with 10% heat-inactivated sheep serum before coverslipping and incubation with

primary antibody (Sheep anti-digoxigen alkaline phosphatase Fab Fragment, 1:1500; Roche)

overnight at 4° C. The following day, coverslips were removed and slides were rinsed in maleic

Page 89: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

89

acid buffer containing Tween 20 3 X 10 min and predeveloping buffer (100 mM Tris pH 9.8,

100 mM NaCl, 50 mM magnesium chloride) for 2 X 2 min at room temperature. Slides were

then transferred to developing solution (100 mM Tris pH 9.8, 100 mM NaCl, 50 mM magnesium

chloride, 5% polyvinyl alcohol, 0.11 mM nitroblue tetrazolium salt in 70% dimethylformamide,

5-bromo-4-chloro-3-indolyl-phosphate in 100% dimethylformamide) in a foil-wrapped coplin jar

at 37° C for 36 h. Slides were then briefly rinsed in tap water to stop the color reaction. Sections

were then dehydrated in 2 min rinses of increasing concentrations of ethanol, defatted 2 X 2 min

in xylene, and then coverslipped with DPX mounting media (VWR).

Image Capture and Analysis

Brightfield images of CRH digoxigen in situ hybridization labeled sections were taken

with a Plan-Apochromat 10x/0.45 M27 objective using light filters to obtain maximum contrast.

Densitometry of CRH mRNA labeling was performed using Image J (NIH) by first outlining a

region of interest based on boundaries established in The Mouse Brain in Stereotaxic

Coordinates 3rd Ed (Franklin and Paxinos, 2008) for separate rostral-caudal levels of the PVN,

and then using the “measure” function to determine the raw integrated density of the confined

area. Bi-lateral densities for each PVN atlas-matched level were averaged for brain sections

taken from control and salt-loaded mice.

Brightfield images of thionin-stained were taken with a Plan-Apochromat 20x/0.8 M27

objective using light filters to obtain maximum contrast. PVN boundaries were established using

figures in The Mouse Brain in Stereotaxic Coordinates 3rd Ed (Franklin and Paxinos, 2008) for

each rostral-caudal levels of the PVN.

Fluorescent images were captured using an AxioImager M.2 fluorescent microscope

(Carl Zeiss, Thornwood, New York) equipped with an Apotome.2 and connected to a PC

running Axiovision 4.8. Using a Plan-Apochromat 20x/0.8 M27 objective, z-stacks of

Page 90: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

90

tdTomato, fluorescently-labeled CRH mRNA, and vasopressin/oxytocin-NP were captured from

sections corresponding to various rostral-caudal levels through the PVN (from Bregma -0.46 mm

to -1.22 mm). The same sections that included the three most caudal regions of the PVN were

also used to image the amygdala. Each image was assigned to a specific atlas figure by an

individual experimenter that analyzed the shape of the PVN and amygdala proper and matched it

with the shapes of adjacent anatomical landmarks found in The Mouse Brain in Stereotaxic

Coordinates 3rd Ed (Franklin and Paxinos, 2008). Image size was approximately 1388 X 1040

pixels and each z-step was 1 µm with an average of 20 optical sections per PVN image. When

capturing images from sections processed for immunohistochemistry, exposure time was

determined based on an overexposure reporting function automatically set by the software.

Exposure time during the imaging process varied from 30-40 ms for the channel capturing

tdTomato images and 125-175 ms for vasopressin/oxytocin neurophysin images. When

capturing images from sections processed for in situ hybridization, the longest exposure time that

still resulted in a lack of signal when imaging sections hybridized with the negative control probe

was used. Images were adjusted using the brightness and contrast settings available in

Axiovision to increase clarity.

Quantification of colocalized tdTomato with either vasopressin or oxytocin neurophysin

neurons took place using a method similar to the “direct three-dimensional counting” process

(Williams and Rakic, 1988). Each z-stack was opened in Axiovision 4.8 or Zen lite 2012 (Carl

Zeiss) and then each image was inspected while alternately turning the respective channels on

and off to identify potential colocalizations. Candidate cells were then further analyzed as

necessary by removing the area of the image containing the cell into a separate region of interest

(ROI) image to generate a 3D reconstruction. If, as a 3D reconstruction, the cell contained

Page 91: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

91

fluorescent labeling on each individual channel throughout the z-plane of the image, then the cell

was counted. The number of bi-lateral colocalizations were determined for each atlas-matched

section containing the PVN. An identical procedure was repeated for image capture and analysis

of the SON. Statistical analyses were performed by grouping according to atlas region and in a

separate analysis by grouping all images obtained from each subject.

Sections from CRH-reporter mice were processed for double immunofluorescent labeling

of vasopressin-NP and FG and imaged for quantification of preautonomic neurons expressing

vasopressin-NP or tdTomato. FG injection site analysis was accomplished by imaging untreated

hindbrain sections with a DAPI filter set (E/A, 329/461 nm) to observe FG fluorescence and

determine injection sites that were comparable for PVN analysis. Specifically, each image was

opened in PowerPoint (Microsoft) and manipulated to make the image transparent. The images

corresponding to a particular atlas region were then superimposed over a tracing of the relevant

atlas figure. After correcting for the boundaries and matching anatomical landmarks, a circle

was drawn over the FG-labeled injection site and included in a composite figure comprised of all

injections for either control or salt-loaded mice for each atlas figure containing the RVLM.

The total number of FG positive cells in the PVN were determined from mice with

injection sites that were observed to be similar in location and area. Tissue from mice with

injection sites that did not show FG coverage in the area of the RVLM were excluded from the

study. Analysis of the forebrain from successful injections included three atlas-matched PVN

images per mouse (from Bregma -0.94, -1.06, and -1.22mm) that also contained the amygdala.

These areas were determined to be rich in hindbrain-projecting neurons based on the distribution

of FG positive cells. Images were acquired as described above and z-stacks were scrolled

through to identify FG positive cells based on a clear cellular profile and uniform labeling. Cells

Page 92: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

92

that met these criteria were marked with an event marker function and totals were recorded. This

process was repeated for tdTomato and vasopressin-NP. Finally, single-channel images and

superimpositions were used to determine the total number of FG positive cells, FG+vasopressin-

NP positive cells as a percentage of all FG cells, and FG+tdTomato positive cells as a percentage

of all FG cells.

Statistics

All grouped data presented as mean +/- SEM. Atlas-matched colocalization data were

assessed with a two-factor analysis of variance. Main effects or interactions (P<0.05) were

assessed post hoc with the Bonferroni method. Total neuropeptide/tdTomato colocalization data

obtained from grouping all PVN atlas regions from each subject were assessed with an unpaired

2-tailed t-test. Significance for all analyses was set at P<0.05. Statistical analyses were

performed and graphs were created using GraphPad Prism 5.

Results

Salt-Loading Reduces CRH mRNA Expression in the PVN

Figure 5-1 shows representative images of sequential (A and C, B and D) coronal brain

sections taken from a series through the PVN and processed for in situ hybridization and

digoxigen immunohistochemistry. At Bregma -0.70 mm, the CRH mRNA labeling represents a

dense triangular region (Fig. 5-1, A) that is markedly reduced in the ventral portion of the PVN

with salt-loading (Fig. 5-1, B). At Bregma -0.82 mm, CRH expression following salt-loading

appears localized to a cluster in the dorsal PVN (Fig.5-1, C) that is lessened in area, but not in

density with salt-loading (Fig. 5-1, D). Quantification of the average density for each PVN

region revealed a significant (P<0.05) decrease in CRH mRNA density with salt-loading at

Bregma -0.70 mm, but no overall difference at Bregma -0.82 mm (Fig. 5-1, E).

Page 93: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

93

Fluorescent Tracking of CRH mRNA by tdTomato

Assessment of the specificity of Cre/loxP-mediated reporting for CRH transcription was

performed using in situ hybridization. CRH transcript was labeled in PVN sections of mice

drinking water, 2% NaCl, water + restraint, or 2% NaCl + restraint. Adjacent sections were

processed for thionin staining for determination of PVN subregions at Bregma -0.82 (Fig. 5-2,

A). Subregions were outlined and used to compare DAPI-labeled nuclei, CRH mRNA+DAPI

nuclei, and CRH mRNA+tdTomato expressing cells. Semi-quantitative analysis revealed that

the area of each PVN subregion contained comparable numbers of neurons in each group as the

number of DAPI-positive nuclei were similar (Fig.5-2; D, E). Furthermore, while restraint

increased the number of DAPI+CRH mRNA counts and DAPI+tdTomato counts, salt-loading

alone decreased the number of DAPI+CRH mRNA counts in each subregion, suggesting a trend

toward inhibition of CRH mRNA with salt-loading; however, this inhibitory trend was not

reflected in the number of DAPI+tdTomato counts.

Fluorescently-labeled CRH mRNA and tdTomato-filled neurons of coronal sections

through the PVN were quantified in three atlas–matched brain sections representing the majority

of the rostral-caudal extent of the PVN from an euhydrated unrestrained mouse (Fig 5-2, F1).

CRH mRNA was colocalized with tdTomato 94.36 +/- 4.63% of the time while CRH mRNA that

was not colocalized with tdTomato (arrows) represented 9.63 +/-5.74% of total CRH mRNA

positive cells (Fig 5-2, F2).

Total tdTomato Unaffected by Salt-Loading or Restraint

The total number of tdTomato positive neurons was similar between groups (No main

effects or interaction; Fig. 5-3) although the mean count tended to be higher for salt-loaded

restrained mice vs water-drinking restrained mice at Bregma -0.82mm (Water+Restraint:

176.60+/-16.56, 2% NaCl+Restraint: 196.00+/-4.97) and Bregma -0.94mm (Water+Restraint:

Page 94: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

94

72.00+/-9.85, 2% NaCl+Restraint: 109.33+/-18.35). In contrast, the mean counts for

unrestrained mice were nearly the same: Bregma -0.82mm (Water No Restraint: 183.00+/-8.22,

2% NaCl No Restraint: 183.500+/-5.41) and Bregma -0.94mm (Water No Restraint: 87.60+/-

15.19, 2% NaCl No Restraint: 84.00+/-12.29).

Salt-Loading is not Associated with Alterations in the Extent of Oxytocin-NP/tdTomato

Colocalizations in the PVN

Figures 5-4 and 5-5 show representative images of coronal sections through the PVN of

mice drinking water or 2% NaCl. Comparable atlas-matched sections from both groups were

immunofluorescently labeled for oxytocin-NP. As indicated by the large number of tdTomato

positive neurons in Fig. 5-4, this region contains a large number of putative neurosecretory

neurons, while Fig. 5-5 contains primarily preautonomic neurons (confirmed in Fig. 5-11).

Sections corresponding to Bregma -0.70 mm or Bregma -0.82 mm had only scant colocalizations

in both the Water (Fig. 5-4 A, C) and 2% NaCl (Fig. 5-4 B, D) groups. The distribution of

tdTomato and oxytocin-NP in these two particular regions was clearly distinct from other regions

of the PVN. At Bregma -0.70 mm, tdTomato cells were organized as a dense cluster of neurons

surrounded by a comparatively much smaller number of oxytocin-NP cells. The tdTomato and

oxytocin-NP expression at Bregma -0.70 mm was very similar to that seen at Bregma -0.82 mm

except for a reduction in the amount of tdTomato expression in the ventral portion of the PVN.

At Bregma -0.94 mm and Bregma -1.06 mm, colocalizations were observed more frequently in

both control and salt-loaded mice compared with the other regions analyzed. tdTomato

expression in these two caudal regions decreased compared with the more rostral areas and

furthermore oxytocin-NP expression was more laterally distributed (Fig. 5-5 A-D).

Two-way ANOVA revealed a main effect of atlas region [F(4, 40) =17.30 (P<0.05)], but

not treatment and no interactions were found. The average number of tdTomato and oxytocin-

Page 95: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

95

NP colocalizations increased in the two most caudal sections through the PVN but this increase

was similar for both groups (summarized by atlas region in Fig. 5-4 E and Fig. 5-5 E, and

averaged for all subjects in Fig. 5-8 A).

Salt-Loading is Associated with an Increase in Vasopressin-NP/tdTomato Colocalizations

in the PVN

Figures 5-6 and 5-7 shows representative unilateral coronal sections

immunofluorescently labeled for vasopressin-NP through the PVN of CRH-reporter mice

drinking either water or 2% NaCl. The overall distribution of vasopressin-NP neurons appeared

much closer to and interspersed with tdTomato cells compared with the distribution of oxytocin-

NP cells. This was especially true in the ventral portion of the PVN (Fig. 5-6 A-D).

Furthermore, vasopressin-NP-positive soma were found more frequently in the neurosecretory

region, with the majority of vasopressin-NP immunoreactivity occurring at Bregma -0.70 mm

and Bregma -0.82 mm. The high density of tdTomato and vasopressin-NP cells observed in

these two atlas regions were also associated with a significant increase in colocalizations.

Specifically, an atlas-region x treatment interaction [F(4,81)=4.59 (P<0.05)] occurred with main

effects of both atlas region [F(4,81)=25.01 (P<0.05)] and 2% NaCl treatment [F(1,81)=21.68

(P<0.05)]. Post hoc analysis revealed sections related to Bregma -0.70 mm and Bregma -0.82

mm having significantly more (P<0.05) tdTomato/vasopressin-NP colocalizations with salt-

loading (Fig. 5-6 E). This effect contributed to an overall increase in the average number of

tdTomato/vasopressin-NP colocalizations throughout the PVN which more than doubled in

response to salt-loading (P<0.05) (Fig. 5-8 B).

Salt-Loading does not Affect CRH Expression in Vasopressin-NP or Oxytocin-NP

Expressing Neurons of the SON

Figure 5-9 depicts representative images of sections through the SON from Water and

2% NaCl mice. Analysis of tdTomato/vasopressin-NP colocalizations (Fig. 5-9 C) and

Page 96: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

96

tdTomato/oxytocin-NP colocalizations (Fig. 5-9 D) indicates no apparent effects of salt-loading.

Furthermore, both groups exhibited only sparse expression of tdTomato in the SON that

amounted to only one or two cells (data not shown).

Fluorogold Injections were Localized to the Nucleus Ambiguous, Pre-Bötzinger Complex,

and Rostroventrolateral Medulla

The distribution patterns of FG injection sites were consistent between Water and 2%

NaCl mice, indicating similar delivery of retrograde tracer to the hindbrain of both groups (Fig.

5-10). Of 33 injected mice, 7 were included from the Water group and 8 from the 2% NaCl

group. Overall, the injections delivered similar amounts of tracer to regions including and also

adjacent to the RVLM, specifically the nucleus ambiguous and pre-Bötzinger complex.

Only injection sites determined to have tracer injected into the RVLM were included for

subsequent FG quantification in the PVN. The most rostral atlas region (Bregma -6.48 mm) that

shows the RVLM (Fig. 5-10, A1 and B1) is also characterized by the lack of a nucleus

ambiguous, and at this level one section from the Water group was counted as a hit (Fig. 5-10

A1). While five sections from the 2% NaCl exhibited FG at this same level, only one of these

overlapped with the RVLM area (Fig. 5-10 B1, excluded injection sites shown in red). The next

region moving caudally (Bregma -6.64 mm; Fig. 5-10, A2 and B2) shows four RVLM hits in the

Water group (Fig. 5-10 A2) and four out of five hits in the 2% NaCl group (Fig. 5-10 B2). The

region at Bregma -6.72 mm included five hits in the Water (Fig. 5-10 A3), and four out of five

from 2% NaCl groups, respectively (Fig. 5-10 B3). At Bregma -6.84 mm, five out of six hits

were determined in the Water group (Fig. 5-10 A4) and five out of seven in the 2% NaCl group

(Fig. 5-10 B4).

Page 97: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

97

Retrogradely-Labeled Neurons Increase Expression of tdTomato and Vasopressin-NP in

PVN

Injection site analysis resulted in the inclusion of 7 mice from the Water group and 8

mice from the 2% NaCl group for quantification of soma in the PVN. Sections labeled for FG

and vasopressin-NP were analyzed first for the number of FG-labeled cells in each section.

Representative images are of the three atlas-matched regions containing the majority of

preautonomic cells (Fig. 5-11 A-F). Quantification of FG-labeled cells did not reveal a

significant interaction or a main effect of 2% NaCl but there was a main effect of atlas-region

[F(2, 36)=3.90 (P<0.05)] (Fig. 5-11 G). Average FG-expressing cells counted for all sections were

not significantly different (Fig. 5-11 H).

Vasopressin-NP cells in the caudal PVN were not labeled with the same intensity as more

anterior regions (Fig. 5-12 A). Labeling was often faint and granular (Fig. 5-12 B). There were

significant effects of both atlas-region [F(2,36)=4.91 (P<0.05)] and 2% NaCl [F(1,36)=4.40

(P<0.05)]. Post hoc analysis did not reveal significant differences in any particular atlas region

(Fig. 5-12 E); however, the average percentage of vasopressin-NP/FG colocalizations were

increased in response to salt-loading (P<0.05) (Fig. 5-12 F).

The tdTomato expression in all three atlas regions comprising the caudal PVN was much

less abundant and scattered than in more anterior PVN regions. Figure 5-13 shows

representative images of coronal atlas-matched sections through the PVN. Compared with the

controls (Fig. 5-13 A), salt-loading increased tdTomato colocalizations with FG (Fig. 5-13 B).

Two-way ANOVA detected a main effect of 2% NaCl [F(1,36)=17.02 (P<0.05)]. Post hoc

analysis revealed a significant increase in tdTomato/FG colocalizations at Bregma -0.94 mm

(P<0.05) (Fig. 5-13 C). The average percentage of tdTomato/FG colocalizations within the PVN

was increased in response to salt-loading (P<0.05) (Fig. 5-13 D).

Page 98: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

98

Retrogradely-Labeled Neurons in the Amygdala are Distinct from tdTomato-Expressing

Neurons

Analysis of brain sections including the amygdala was performed from three salt-loaded

mice and three water-drinking controls for the same atlas-matched regions described above for

the PVN. Figure 5-14 shows the distribution of tdTomato-expressing neurons and neurons

labeled with FG at Bregma -1.06. The densest expression of tdTomato was found in the lateral

portion of the amygdala which was almost entirely populated with neurons reporting for CRH.

FG expression was found almost exclusively in neurons of the medial amygdala, and there were

no colocalizations with very sparsely distributed tdTomato-expressing neurons. Other regions

with scant expression of tdTomato include the basolateral amygdala, the extended amygdala, the

anterior part of the basomedial amygdaloid nucleus, and the central nucleus of the amygdala;

regions which also had no retrogradely-labeled neurons.

Discussion

We used fluorescent indication of CRH transcription in conjunction with

immunohistochemistry and retrograde neuronal tract tracing to evaluate how salt-loading caused

by consumption of 2% NaCl changes neuropeptide expression within distinct populations of

PVN neurons. Salt-loading significantly decreased and redistributed CRH mRNA within

subregions of the PVN. Subsequent experiments utilized tdTomato reporting of CRH

transcription to decipher which PVN neurons exhibited changes in neuropeptide expression as a

consequence of salt-loading. Consumption of 2% NaCl did not affect tdTomato expression in

the SON or tdTomato colocalization with oxytocin-NP in the PVN. However, salt-loading did

significantly increase the colocalization of vasopressin and tdTomato in putative magnocellular

PVN neurons. To evaluate changes in neuropeptide expression within preautonomic neurons,

CRH-reporter mice were injected with FG targeting the RVLM. Interestingly, salt-loading

Page 99: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

99

significantly increased the colocalization of vasopressin and FG as well as tdTomato and FG

within preautonomic neurons in the PVN. This effect seemed to be specific to the PVN, as no

tdTomato colocalizations with FG were observed in the amygdala. Our results demonstrate that

chronic osmotic stress increases expression of CRH in neuronal circuits that maintain plasma

tonicity and blood pressure. The implication is that such neuropeptide plasticity alleviates the

deleterious consequences of chronic dehydration but may progressively become a

pathophysiological mechanism that contributes to the etiology of stress-related disease.

The parvocellular and magnocellular subdivisions of the mouse PVN are not well defined

which impedes the localization of CRH mRNA to these subdivisions (Biag et al., 2012). To

circumvent this impediment, we used the Cre-loxP system to report the transcription of CRH

mRNA with the red fluorescent protein, tdTomato, which allowed evaluation of CRH production

within specific neuronal phenotypes. Using this approach, we determined that salt-loading did

not affect the colocalization of tdTomato with oxytocin-NP, suggesting that salt-loading does not

increase CRH and oxytocin colocalization within the PVN of mice. In addition, we observed

few tdTomato expressing cells in the SON under basal conditions and did not observe an effect

of salt-loading on tdTomato expression within this nucleus. Studies conducted using rats have

found that CRH mRNA is expressed in magnocellular neurons of the PVN and SON that

produce oxytocin, and that oxytocin mRNA and oxytocin/CRH colocalizations increase in

response to chronic salt-loading (Majzoub et al., 1983; Lightman and Young, 1987; Kovacs and

Sawchenko, 1993; Imaki et al., 2003). This plasticity may influence pituitary physiology

(reviewed in (Bondy et al., 1989)) because CRH derived from magnocellular neurons acts in the

intermediate lobe of the pituitary to augment systemic neuropeptide release through a mechanism

dependent, in part, on melanocyte-stimulating hormone signaling. Our results, consistent with

Page 100: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

100

those of others (Imaki et al., 2003), suggest that the osmotically-induced CRH plasticity that

occurs in magnocellular oxytocin neurons within the PVN and SON of rats does not occur in

mice. The implication is that the effects of salt-loading on CRH expression and systemic

neuropeptide release may also differ between rats and mice.

In contrast to what was observed with tdTomato and oxytocin-NP, salt-loading increased

the colocalization of tdTomato with vasopressin-NP in the PVN. These colocalizations occurred

most frequently in a region of the PVN known to contain a dense cluster of CRH-positive

parvocellular neurons. In this area, vasopressin-positive neurons closely surround and are

interspersed with CRH neurons that project to the median eminence (Biag et al., 2012).

Colocalization of CRH and vasopressin in neurosecretory parvocellular neurons allows their

corelease, which facilitates the secretion of ACTH into the systemic circulation (Gillies et al.,

1982; Vale et al., 1983). It is unlikely that the increase in tdTomato/vasopressin-NP

colocalizations that we observed occurred in neurosecretory parvocellular neurons that initiate

HPA-axis activation because salt-loading attenuates CRH and does not increase vasopressin

mRNA in the parvocellular subdivision of the PVN (Scott Young, 1986; Kovacs and

Sawchenko, 1993; Aguilera and Rabadan-Diehl, 2000; Grinevich et al., 2001). It is possible that

chronic osmotic stress allows the corelease of CRH and vasopressin to act at intermediate

pituitary CRH receptors to facilitate vasopressin secretion, similar to the corelease of oxytocin

and CRH that occurs in magnocellular neurons of rats (Bondy and Gainer, 1989). This

interpretation is consistent with previous reports demonstrating that chronic salt-loading elevates

vasopressin (Morita et al., 2001). Furthermore, osmotically-induced release of vasopressin into

the systemic circulation is inhibited by corticosterone (Papanek et al., 1997) and suppressing

HPA axis activity could permit heightened release of vasopressin to alleviate the hypertonicity

Page 101: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

101

that occurs with chronic salt-loading. Here, we propose that, within the PVN, salt-loading

increases CRH expression in magnocellular neurons to elicit water reabsorption through

vasopressin secretion.

The expression of CRH in the PVN is not limited to neurons that project to the median

eminence, but is also found in neurons making extrahypothalamic synapses where CRH acts as a

neurotransmitter or modulator (Dunn and Berridge, 1990). In particular, neurons within the PVN

that project to hindbrain nuclei, deemed preautonomic neurons, are known to express CRH and

are implicated in the control of sympathetic outflow and arterial blood pressure (Sawchenko,

1987; Milner et al., 1993; Yang and Coote, 1998). In the present study, we identified

preautonomic neurons as those labeled with retrograde tracer delivered into the RVLM and its

adjacent nuclei. Importantly, we established similar tracer distribution and number of

retrogradely labeled preautonomic neurons in the PVN of control and salt-loaded mice, thereby

allowing quantitative comparisons between these groups. A portion of retrogradely-labeled PVN

neurons were also positive for tdTomato and the distribution of these colocalizations was similar

to that of RVLM-projecting CRH neurons previously reported in rats (Milner et al., 1993). In

contrast, none of the FG-positive neurons in the amygdala were positive for tdTomato as neurons

labeled for FG were confined to the medial portion of the amygdala while tdTomato-positive

neurons were found densely expressed in the lateral portion. Interestingly, mice subjected to

salt-loading had a greater percentage of retrogradely labeled neurons in the PVN that expressed

tdTomato, suggesting chronic osmotic stress initiates the transcription of CRH in PVN neurons

with efferent projections to the RVLM. The RVLM is a hub for many converging and diverging

signals relating to sympathetic tone (reviewed (Guyenet, 2006)) and injection of CRH into the

RVLM increases arterial pressure in rats (Milner et al., 1993). Thus, the increased expression of

Page 102: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

102

CRH that occurs in PVN preautonomic neurons subsequent to salt-loading may allow these

neurons to utilize CRH receptors in the RVLM to promote sympathetic outflow and

cardiovascular adaptations to chronic osmotic stress.

The increased expression of CRH in preautonomic neurons that occurs with salt-loading

may work in concert with vasopressin to influence autonomic outflow subsequent to chronic

osmotic stress. Consistent with the results of Pretel & Piekut (Pretel and Piekut, 1989), we found

that PVN neurons retrogradely identified as projecting to the RVLM and its adjacent nuclei were

immunoreactive for vasopressin-NP. These results are consistent with prior reports that PVN

vasopressinergic neurons project to the RVLM in both rats and mice (Kc et al., 2002; Kc et al.,

2010; Rood et al., 2013) but conflict with other studies indicating that very few if any

vasopressin neurons project to the RVLM (Gomez et al., 1993; Stocker et al., 2006). This

discrepancy can be explained by the induction of vasopressin in preautonomic neurons by salt-

loading or could also be due to inadvertent administration of retrograde tracer to the ambiguous

nucleus, which receives vasopressinergic projections from the PVN (Wang et al., 2002; Rood et

al., 2013). In the latter case, there is evidence that vasopressin promotes parasympathetic

withdrawal in the ambiguous nucleus (Wang et al., 2002) that may complement the action of

vasopressin by increasing sympathetic output via the RVLM (Kc et al., 2010).

The increase in colocalization of tdTomato with FG and AVP does not agree well with

our data that did not find any significant changes in tdTomato expression. The average number

of tdTomato positive neurons in the PVN of salt-loaded restrained mice was higher than the

average number determined for restrained water-drinking controls (Fig. 5-3), suggesting a trend

towards increased reporting for CRH. The statistical methods employed to determine any

differences between groups rely on comparing mean counts relative to the variability observed to

Page 103: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

103

make the claim that the increase was significant. It is highly likely that the variability

contributing to total tdTomato came from additional sources beyond those influencing the

variability observed in the determination of colocalization of tdTomato with AVP and FG. If

this is the case, then the true effect size of salt-loading on total tdTomato may be small or

medium, while the true effect size of salt-loading on tdTomato in discrete populations of PVN

neurons may be large. In support of this, the experiments performed by Imaki et al. suggest that

in mice, but not rats, total CRH expression in the PVN is only subtly altered during salt-loading,

promoting the idea that significant increases in CRH expression are likely only observed in a

narrowly defined population (Imaki et al., 2003).

Taken together, our results illustrate potential mechanisms by which CRH and

vasopressin may be differentially expressed in the PVN to counter chronic saline consumption.

Both of these neuropeptides become dysregulated in widespread diseases and disorders including

those with psychiatric and cardiovascular components (discussed in more detail in Chapter 6).

The increased colocalization of CRH and vasopressin may work to increase systemic

vasopressin, but the consequences of this plasticity on the rest of the network of which the PVN

is a hub has yet to be determined. Similarly, the increase in preautonomic neuropeptide

expression suggests either an adaptation to maintain normal cardiovascular function during an

extreme overload of salt or is a marker for a maladaptive consequence of some other mechanism.

Given our data supporting a lack of dehydration-induced anorexia or hypovolemia, it is likely the

former case; however, it may be that the duration of salt-loading in the present experiments did

not allow for sufficient development of other mechanisms to surface. These mechanisms might

include the involvement of SON neurons or oxytocin neurons and future experiments would

indeed shed additional light on the overall plasticity of PVN neurons during chronic salt-loading.

Page 104: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

104

Figure 5-1. Representative images of PVN sections processed for in situ hybridization of

digoxigenin-conjugated probe and CRH mRNA. Comparable sections from mice in

the Water (A, C) and 2% NaCl (B, D) groups. Salt-loading was associated with a

decrease in density at Bregma -0.70 mm, but did not differ at Bregma -0.82 (E,

P<0.05). Water n=4-6, 2% NaCl n=4. Scale bars = 100 µm. 3v, 3rd ventricle.

Page 105: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

105

Figure 5-2. Analysis of tdTomato reporting for CRH mRNA expression in the PVN. Thionin

staining for determination of PVN subregions at Bregma -0.82 (A). Coronal section

through the PVN depicting DAPI (blue nuclei) and CRH mRNA in situ hybridization

(magenta dots) (B). Superimposed image of CRH mRNA in situ hybridization

(magenta dots) and tdTomato-filled cells (green) (C). Total counts of DAPI,

DAPI+CRH mRNA, and DAPI+tdTomato in the lateral magnocellular (PaLM) (D)

and dorsal cap (PaDC) (E). Coronal section through the PVN depicting CRH mRNA

and tdTomato positive neurons as well as neurons with CRH mRNA expression

without tdTomato expression (F1). Graph depicting reporter fidelity (F2). Inset:

closeup of CRH mRNA expression colocalized with DAPI and tdTomato, or DAPI

only. Arrows indicate false negatives. Lateral Magnocellular, PaLM. Dorsal Cap,

PaDC. Error bars = SEM. Scale bars = 50 µm (A-C, F), 10 µm (F1, inset). 3v, 3rd

ventricle.

Page 106: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

106

Figure 5-3. Analysis of total tdTomato in the PVN. Neither restraint or salt-loading significantly

increased the number of tdTomato positive neurons in the PVN.

Page 107: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

107

Figure 5-4. Representative unilateral images and quantification of oxytocin-neurophysin (OT-

NP)/tdTomato colocalizations in the “neurosecretory” region of the PVN. Total

bilateral colocalizations were determined for mice maintained on water (left side of

each outline; A and C) and mice maintained on 2% NaCl (right side of each outline;

B and D). Maintenance on 2% NaCl did not affect the number of OT-NP/tdTomato

colocalizations although there was a main effect of atlas-region [F(1, 16)=29.18

(P<0.05)] with more colocalizations occurring at Bregma -0.82 (E). Water n=5, 2%

NaCl n=5. Error bars = SEM. Scale bars = 100 µm. 3v, 3rd ventricle.

Page 108: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

108

Figure 5-5. Representative unilateral images and quantification of oxytocin-neurophysin (OT-

NP)/tdTomato colocalizations in the “preautonomic” region of the PVN. Total

bilateral colocalizations were determined for mice maintained on water (left side of

each outline; A and C) and mice maintained on 2% NaCl (right side of each outline;

B and D). Maintenance on 2% NaCl did not affect the number of OT-NP/tdTomato

colocalizations (P>0.05E). Water n=5, 2% NaCl n=5. Error bars = SEM. Scale bars

= 100 µm. 3v, 3rd ventricle.

Page 109: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

109

Figure 5-6. Representative unilateral images and quantification of vasopressin-neurophysin

(AVP-NP)/tdTomato colocalizations in the “neurosecretory” region of the PVN.

Total bilateral colocalizations were determined for mice maintained on water (left

side of each outline; A and C) and mice maintained on 2% NaCl (right side of each

outline; B and D). Maintenance on 2% NaCl revealed a significant effect of 2% NaCl

[F(1,34)=19.64 (P<0.05)] with sections related to Bregma -0.70 mm and Bregma -0.82

mm having significantly more (P<0.05) AVP-NP/tdTomato colocalizations with salt-

loading (E). Water n=8, 2% NaCl n=9-10. Error bars = SEM. Scale bars = 100 µm.

3v, 3rd ventricle.

Page 110: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

110

Figure 5-7. Representative unilateral images and quantification of vasopressin-neurophysin

(AVP-NP)/tdTomato colocalizations in the “preautonomic” region of the PVN. Total

bilateral colocalizations were determined for mice maintained on water (left side of

each outline; A and C) and mice maintained on 2% NaCl (right side of each outline;

B and D). Maintenance on 2% NaCl had no effect on AVP-NP/tdTomato

colocalizations (P>0.05, E). Water n=8, 2% NaCl n=9-10. Error bars = SEM. Scale

bars = 100 µm. 3v, 3rd ventricle.

Page 111: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

111

Figure 5-8. Average total colocalizations in the PVN. The average total number of oxytocin-

neurophysin (OT-NP)/tdTomato colocalizations for all PVN regions from each

subject was not significantly different between groups (P>0.05, A). Salt-loading was

associated with significantly increased (P<0.05, B) total vasopressin-neurophysin

(AVP-NP)/tdTomato colocalizations. Water (open bars) n=5, 2% NaCl (grey bars)

n=5. Error bars = SEM.

Page 112: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

112

Figure 5-9. Representative unilateral images used for quantification of colocalizations in the

supraoptic nucleus. There was no obvious qualitative change in tdTomato expression

due to salt-loading (A, B). The tdTomato expression that did occur was not

significantly affected in terms of colocalization with either vasopressin-NP (C) or

oxytocin-NP (D). Water (open bars) n=3, 2% NaCl (grey bars) n=3. Error bars =

SEM. Scale bars = 50 µm. ot, optic tract.

Page 113: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

113

Figure 5-10. A summary of FG injections targeting the RVLM. Injections from the control

group are shown in the left column (A1-4) and injections from salt-loaded mice are

shown to the right (B1-4). The atlas figures are tracings from individual plates

published in The Mouse Brain in Stereotaxic Coordinates 3rd Ed containing the four

regions (from top to bottom: Bregma -6.48 mm, -6.64 mm, -6.72 mm, -6.84 mm)

outlining the boundaries of the RVLM. The images to the left of each atlas drawing

(a, b, c, etc) are the actual images from which the location of FG-labeled area was

determined. Each image is labeled with a lower case letter designating its

corresponding injection site on the figure tracing to the right. Any image with FG

labeling was included for analysis regardless of whether or not it was determined to

be close enough to the RVLM to be acceptable for PVN analysis, thus hits are

designated with a green outline while misses are designated with a red outline.

Images are ordered with misses first (if any) followed by hits. Scale bars = 500µm

Page 114: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

114

Page 115: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

115

Figure 5-11. Representative unilateral images of z-stack 3D reconstructions used for

quantification of PVN neurons labeled with Fluorogold (FG). Images shown are

sequential from a series of sections taken from a single control subject (A, C, E) and a

single subject from the salt-loaded group (B, D, F). Total bilateral counts of FG-

positive cells were determined. Quantification revealed a similar number of FG-

positive cells for each PVN atlas region in each group (G). The average total number

of FG cells for all PVN regions from each subject was also not significantly different

(H). Water n=6-7, 2% NaCl n=7-8. Error bars = SEM. Scale bars = 50 µm. 3v, 3rd

ventricle.

Page 116: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

116

Figure 5-12. Representative images and quantification of Fluorogold (FG)/vasopressin-NP

(AVP-NP) colocalization. In contrast to more anterior sections, AVP-NP labeling

appeared granular (B). Quantification revealed a non-significant increase in the

number of FG/AVP-NP colocalizations (D) with salt-loading for each PVN atlas

region (E). The average total number of FG/AVP-NP colocalizations for all PVN

regions from each subject was higher (F) in salt-loaded mice compared with controls

(P<0.05). Error bars = SEM. Scale bars: A = 50 µm, B-D = 10 µm. 3v, 3rd ventricle.

Page 117: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

117

Figure 5-13. Representative images and quantification of Fluorogold (FG)/tdTomato

colocalization (A, B). Quantification revealed a significant increase in the number of

FG/tdTomato colocalizations (C) with salt-loading at Bregma -0.82 mm. The average

total number of FG/tdTomato colocalizations for all PVN regions from each subject

(D) was higher in salt-loaded mice compared with controls (P<0.05). Error bars =

SEM. Scale bars = 50 µm; 3v, 3rd ventricle.

Page 118: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

118

Figure 5-14. Representative image of Fluorogold (FG) and tdTomato in the amygdala. Typical

FG expression was located in the medial portion of the amygdala (CeM) while the

densest expression of tdTomato was in the lateral division (CeL). Only scant

expression of tdTomato occurred in the CeM, and in no case did tdTomato colocalize

with FG. Central nucleus of the amygdala (CeC), basolateral amygdala (BLA),

internal capsule (ic). Scale bar = 200 µm.

Page 119: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

119

CHAPTER 6

GENERAL DISCUSSION AND FUTURE DIRECTIONS

The data presented here serves two important general functions that allow for more

thorough future investigation into neuroplastic changes in the PVN that accompany salt-loading.

First, the extension of reported findings in the acute salt-loading paradigm from the rat model to

the mouse model was a necessary step towards gaining valuable tools to investigate the effects of

salt-loading on a genetic level. Similarly, the comparison of the general effects of chronic salt-

loading on ingestive behavior and measures of plasma constituency to results obtained using rats

is valuable to the researcher considering utilizing mice in place of or in conjunction with

experiments using rats. Second, the use of a transgenic reporter mouse to visualize CRH-

positive neurons represents a significant step forward in overcoming the technical limitations

inherent in visualizing CRH-positive neurons with high spatial precision. Granted this

improvement, like any method, has its limitations that must be acknowledged upfront, and

associated data must be interpreted in light of these limitations. These two general points will be

discussed together with the more specific points that might be better understood in the context of

species differences and methodological concerns.

The results of Chapter 3 confirmed that the effects of acute salt-loading involving

suppression of the HPA axis using a rat model could be replicated in a mouse model and further

implicated a centrally-mediated inhibitory mechanism. The results obtained by both Krause et

al. and Frazier et al. in rats included slight increases in pNa+ levels, without associated increases

in hematocrit as well as a blunted restraint-induced CORT (Krause et al., 2011a; Frazier et al.,

2013). All of these effects were found and reported here in Chapter 3 suggesting that the murine

response to restraint experienced in the context of acute hypernatremia involves suppression of

the HPA axis. However, Krause et al. also reported that restraint-induced ACTH was attenuated

Page 120: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

120

during restraint which was a measure not investigated with mice (Krause et al., 2011a). The

significance of this finding in rats was that the decrease in ACTH with CORT suggested that the

inhibitory effect on the HPA axis associated with acute hypernatremia was centrally mediated.

However, because only CORT and not ACTH were measured in our experiments in mice, we did

not have sufficient evidence to implicate a centrally-mediated effect based solely on endocrine

measures. To more directly test this hypothesis, and to do so in a direct way that is not possible

in rats, we utilized a CRH-reporter mouse to visualize Fos activation in PVN CRH neurons from

mice rendered acutely hypernatremic. In these mice, Fos-positive nuclei were significantly

reduced in transgenically-identified CRH neurons compared with controls providing evidence for

a source of inhibition to these cells. Regarding this inhibition, the study in rats reported by

Frazier et al. similarly lacked data showing a decrease in restraint-induced ACTH, but provided

electrophysiological data supporting an inhibitory tone on putative CRH neurons in the PVN that

was dependent on oxytocin-receptors (Frazier et al., 2013). The results from the current study

using mice support the activity of oxytocin as an inhibitory factor in attenuation of CRH neurons

and the HPA axis as mice rendered acutely hypernatremic had more frequent Fos-positive nuclei

in PVN neurons expressing oxytocin. However, additional experimentation is necessary to

confirm an action of oxytocin on CRH neurons or on neurons that innervate CRH neurons in the

PVN. These additional experiments benefit from the translation of findings from rat to mouse as

well as the initial investigation into the validity of the CRH reporter mouse used here. The

validation of the CRH reporter mouse included not only that fluorescent reporting for CRH

transcription colocalized well with fluorescent labeling of CRH mRNA (Figs. 2-1 and 5-2), but

for the first time showed that an osmotic stimulus via 2.0 M NaCl injection had an effect on

activation of genetically modified neurons altered to report for CRH in the PVN. These steps

Page 121: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

121

provide for future experiments to be performed that require the in vitro or post hoc identification

of PVN neurons that are known to be CRH-producing and which are also sensitive to inputs

which convey information regarding elevation of pNa+ concentration.

Chapter 3 included experiments to assess whether the anxiolytic effects of acute

hypernatremia produced in rats extended to the mouse model. Krause et al. found that slightly

elevating the pNa+ concentration resulted in pro-social effects as hypernatremic rats spent more

time interacting a novel conspecific (Krause et al., 2011a). Furthermore, Frazier et al. found that

the same treatment reduced anxiety-like behavior as assessed on the elevated plus maze (Frazier

et al., 2013). The increased activation of oxytocin neurons in the PVN noted in both of these

studies as well as our neuroanatomical data here imply that oxytocin is being released in the

brain, as oxytocin acting in the brain has been shown to have pro-social and anxiolytic effects on

behavior (Landgraf and Neumann, 2004; Neumann and Landgraf, 2012b; Peters et al., 2014).

Anecdotally speaking, rats generally appear to be less anxious and struggle less when confronted

with human handling which indicates that basal anxiety or reactionary behavior is pronounced in

mice in comparison to rats, thus it was unclear whether this behavior could be overcome with the

physiological alterations associated with hypertonic saline treatment. Our results here in mice

found that exploration of the open arms of the elevated plus maze was increased following

hypertonic saline injections indicating that the mechanism influencing anxiety in rats likely is

occurring in mice as well, and that innate behavioral differences between these two species do

not seem to influence the behavioral alterations caused by the hypernatremic state. This could be

because of the well-conserved mechanisms that balance hydromineral homeostasis in both

species which may be very similar despite other neuropeptide or circuit-level differences.

Page 122: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

122

The well-conserved mechanisms responsible for differences in innate behavior, but

similar anxiolytic responses during hypernatremia are likely mediated through the actions of

oxytocin and vasopressin, which seem to have similar centrally- and systemically-mediated

effects in both rats and mice (Insel et al., 1998), but which also display marked gender

differences (Francis et al., 2002). It has been postulated that gender differences exist in

neuropeptide expression for two possible reasons, 1) to promote behavioral differences and 2) to

prevent them (De Vries, 2004) i.e. in the first case, males and females must have very different

social and sexual behavior mediated by oxytocin and vasopressin, but in the latter case both

genders must be able to maintain adequate hydration, blood pressure, body temperature, and

other critical functions which are in part mediated by the same neuropeptides. It may be that this

explanation for differences on a sex basis also may be applicable on a species level as there are

many differences between rats and mice as far as oxytocin and vasopressin expression are

concerned. As previously discussed, perhaps the most obvious is the lack of segregated groups

of magnocellular neurons in the PVN of the mouse (Biag et al., 2012). This anatomical

difference may have functional consequences as magnocellular neurons that are more

interspersed with other neurons in the PVN may have more opportunity for cross-talk with other

neuronal phenotypes. The potential for cross-talk is an interesting and ongoing area of research

(Potapenko et al., 2013; Stern, 2015; Biancardi et al., 2016) involving the study of molecular and

glial-mediated signaling between neurons of the PVN to locally coordinate the output of specific

types of neurons. It is possible that the difference in the “scattered” murine expression of

magnocellular oxytocin and vasopressin in the PVN represents an anatomical basis for

differential control of basal anxiety levels through differential cross-talk with other PVN neurons

or differential projection patterns to limbic brain regions mediating fear and anxiety-like

Page 123: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

123

behavior. However, when confronted with an osmotic challenge, the differences in oxytocin and

vasopressin expression may either be inhibited or become secondary to similarly expressed

oxytocin and vasopressin neurons leading to a reduction in anxiety in both species in a similar

way. If true, the reason for a species difference in basal anxiety and anatomical PVN

magnocellular distribution, but a species similarity in the anxiolytic response to acute

hypernatremia might be found in consideration for the natural habitats of rats and mice. Both

rats and mice evolved in a commensalistic relationship with humans (Atkinson, 1985), but mice

are generally considered to be much more dependent on this association than rats are, thus the

common name house mouse (the house mouse or Mus musculus is the most widely-used

laboratory mouse species). Because of this close association, it may be that in order to survive,

mice developed a heightened sense of awareness due to the greater likelihood of encountering

natural predators that also associate with humans such as dogs, cats, foxes, rats, etc. which would

make them seem more anxious when handled. Rats on the other hand, being a predator of mice

and often exhibiting non-commensal behavior (Aplin et al., 2011; Varudkar and Ramakrishnan,

2015), may have developed a comparatively more “calm” demeanor. However, regardless of the

environmental obligations, each species would need to face exposure in seeking out water which

is analogous to the physiological state hypertonic saline injections induce before behavioral

testing in an apparatus such as the elevated plus maze. Thus it may be that while basal anxiety

levels are different between rats and mice, the activation of anxiolytic pathways related to

hyperosmotic conditions similarly induce changes in behavior.

There are several effects of acute salt-loading in rats that have yet to be confirmed in

mice. Krause et al. found that plasma vasopressin levels were not elevated at the time of

behavioral testing, restraint, and brain sample collection (Krause et al., 2011a). Furthermore,

Page 124: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

124

this study found that plasma renin activity was suppressed during restraint, but that plasma

oxytocin levels were increased, which are also effects not determined by the study performed

here in mice. These measures are important controls for several reasons. While hematocrit and

plasma protein levels are important indicators of plasma volume and therefore activation of the

RAAS, they do not provide a comprehensive view of the humoral consistency. The

determination of plasma vasopressin concentration and renin activity are important controls to

evaluate the endocrine response to hypertonic saline injections. These measures would indicate

whether the dehydration induced by the injections caused only intracellular dehydration or in

addition, caused extracellular dehydration and volume loss through increased urinary output.

High pNa+ generally suppresses the RAAS. Furthermore, vasopressin release is known to

quickly and transiently be released before returning to normal when the behavioral tests and

samples were collected in the acute salt-loading paradigm employed (Stricker and Verbalis,

1986). These points suggest that vasopressin and plasma renin activity would similarly be

affected in our experiments as previous reports using rats. However, in both cases water bottles

were removed following saline injections and the water deprivation period may have impacted

mice differently than rats. Regarding motivated behavior, the additional measures are important

because of the distinct mechanisms involved with hypovolemic thirst compared with hypertonic

thirst. Each would have specific consequences in terms of the circuits activated and the degree

to which various neurotransmitter and neuropeptides were involved in behavior and stress-

responsiveness (Stricker and Verbalis, 1986; Stricker and Sved, 2000). However, given that the

behavioral and CORT responses we observed in our experiments here with mice were consistent

with previous reports in rats it is likely that plasma vasopressin and renin activity were similarly

suppressed. Moreover, we found that oxytocin neurons were significantly activated following

Page 125: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

125

acute hypernatremia which implies that plasma oxytocin levels would be higher than controls

(Krause et al., 2011a), but this cannot be concluded without the proper experiments being

performed. Evidence exists that oxytocin is released systemically in response to elevated pNa+,

but that this is specific to rodents and does not occur in humans (Goodwin et al., 1970; Williams

et al., 1985; Seckl et al., 1989; Rasmussen et al., 2004). It would therefore be beneficial to know

if the increased activation of oxytocin neurons we observed was specific to neurons dedicated to

central projections or included pituitary projections.

As the primary purpose of these experiments was to better understand the capacity for

neuroendocrine and behavioral alterations associated with mediators of hydromineral

homeostasis in humans, it is important to relate these data to possible mechanisms of human

physiology. Clinical correlates to the presented experimental findings using animal models are

abundant, and the following discussion highlights these studies. They show that increased

activity of the RAAS is predictive of depression and anxiety, and when medication that blocks or

inhibits a specific component of the RAAS is prescribed for hypertension, there has also been

improvement in mood and affect. Large scale studies have provided evidence that aldosterone

levels may be a possible solution to the long-sought after biomarker for affective disorders.

Furthermore, direct evidence for the anxiolytic and stress-dampening effects of oxytocin is

available although there have been conflicting reports.

Interest in the relationship between anxiety or depression and RAAS activity found

significant correlations. In a two-year long study involving 1805 participants, researchers found

that depressed hypertensive individuals had elevated aldosterone levels compared with

individuals that were not depressed or those that were not hypertensive (Hafner et al., 2013).

Participants that were either only depressed or, surprisingly, only hypertensive did not present

Page 126: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

126

with heighted renin or aldosterone. In a smaller, but much more specific study carried out while

participants were asleep, Murck et al. monitored HPA and RAAS hormones of unmedicated

depressed and age-matched control subjects finding that aldosterone was significantly higher in

depressed individuals (Murck et al., 2003a; Murck et al., 2012). Furthermore, more than half of

patients with primary aldosteronism surveyed also suffered from an anxiety disorder and these

individuals had higher levels of self-reported stress, psychological distress, and a lower level of

well-being (Sonino et al., 2006; Sonino et al., 2011). As previously discussed, aldosterone

affects specific areas of the brain in rodent models, including the PVN, nucleus of the solitary

tract, locus coeruleus, and amygdala. The clinical data showing a correlation between elevated

aldosterone with anxiety and depression suggests that these areas are perhaps viable targets for

intervention if the primary cause of hyperaldosteronism cannot be directly addressed.

MacDonald and Feifel recently reviewed progress in developing oxytocin-based

therapeutics, citing over 30 clinical studies using intranasal administration of oxytocin in normal

individuals as well as single and multi-dose trials in those suffering from a brain-based illness

(Macdonald and Feifel, 2014). Results in terms of efficacy in reducing anxiety were conflicting;

however, some promising anatomical data was identified. For instance, females with a common

oxytocin receptor gene variant present with a psychological phenotype that includes excessive

worrying, cautiousness, pessimism, shyness, and fearfulness (Wang et al., 2013a). This variant is

associated with reduced resting state PFC to amygdala connectivity as assessed by fMRI and this

circuit’s functional connectivity is improved with oxytocin administration (Dodhia et al., 2014).

It may be that more specific routes of altering the central oxytocin system will be necessary in

order to tailor oxytocin-based therapies to individual disorders.

Page 127: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

127

A better understanding of the oxytocinergic circuitry may lead to novel therapeutic

strategies for treating core symptoms of autism spectrum disorders (ASD). Anecdotal reports and

clinical case studies have found that illness accompanied by fever markedly alleviates symptoms

of ASD (Sullivan et al., 1980; Cotterill, 1985; Curran et al., 2007). A hallmark of fever is

insensible water loss that causes acute hypernatremia or increases in the pNa+ concentration.

Thus, it is plausible that hypernatremia driven increases in central levels of endogenous oxytocin

might alleviate symptoms of ASD. This might be an especially attractive mechanism to exploit

if systemic levels of oxytocin are not affected, which as previously discussed, seems to be the

case in humans. Accumulating preclinical and clinical evidence suggests that brain

oxytocinergic circuits are valid candidate therapeutic targets for relief of ASD symptoms (Andari

et al., 2010; Guastella et al., 2010; Domes et al., 2013; Taurines et al., 2014).

The ability of oxytocin to suppress the activity of PVN CRH neurons seems to be an

effect seen in both humans and animal models. As discussed previously, an acute hypernatremic

state is associated with increased central release of oxytocin that potentiates an inhibitory

oxytocin-dependent tone on CRH neurons of the PVN. In support of this, human males were

given intranasal oxytocin in a randomized, placebo-controlled, double-blind experiment to test

the effect on exercise-induced salivary cortisol levels finding 24 IU, but not 48 IU oxytocin

significantly attenuated cortisol levels compared with placebo (Cardoso et al., 2013). Further

research of this caliber is necessary to determine if similar results are obtained with the induction

of a psychogenic stressor.

These clinical findings and experimental results using human subjects illustrate a

relationship between factors that mediate salt and water handling in the body, and neural circuits

involving the PVN which are often intimately involved in a variety of mental health disorders.

Page 128: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

128

The possibility exists that the reduced anxiety, HPA axis responsiveness, and increased

activation of oxytocin-producing neurons in the PVN presented here in association with acute

hypernatremia might be in some way involved with these disorders. It is possible that repeated

instances of recruiting these mechanisms may lead to conscious or unconscious learning of the

association between the hypernatremic state and perhaps a state of well-being - if it is true that a

human can induce a similar acute hypernatremic state. The injection of hypertonic saline

subcutaneously is an introduction of sodium that is for practical purposes never a mode of

increasing pNa+ levels in humans; rather, this method is meant to efficiently raise pNa+ in a

controlled experimental setting. Any translational potential of these experiments must therefore

ask the question of what a similar stimulus might be in humans in their normal environment or

simply, is it possible for one to raise his/her own pNa+ level to an extent that causes anxiolytic

and HPA axis-dampening effects. These are important questions to answer in order to make the

case that raising one’s pNa+ has reinforcing effects due to a “calming” or positive effect on a

subjective feeling of well-being. One of the major differences between our experimental design

and human behavior in terms of raising the pNa+ is that in our experiments there is no sensation

of taste while obviously humans mediate salt intake largely through gustatory processing. In this

sense, it may be that taste-related central processing might augment the effects of acute salt-

loading in some way and additional experiments that stimulate these pathways might test this

hypothesis. It is also possible that the effects of acute salt-loading might reinforce raising the

pNa+ through unconscious mechanisms as well. Another possibility is that an individual might

not increase the pNa+ solely by increasing salt-intake, but by altering fluid intake and urination

frequency to experience brief periods of hypertonic plasma without an associated reduction in

plasma volume. Finally, there may be no direct translation from rodent model to human

Page 129: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

129

physiology and behavior, but only an indirect summation of findings that, taken together with the

collective findings of others, adds to the general understanding of how mechanisms that control

hydromineral homeostasis also influences behavior, stress-responsiveness, and the activity of

PVN neurons.

To begin to understand the similarities and differences between acute and chronic salt-

loading and to add the element of ingestive behavior to our experimental paradigm, we

performed chronic salt-loading experiments to assess changes in the PVN and behavior.

Importantly, these experiments relied heavily on the identification of neurons in the PVN as

CRH-positive using a CRH-reporter mouse. Direct quantification of CRH protein or transcript

allows for the determination of CRH expression levels at the time brain tissue is sampled. In

contrast, the CRH reporter mouse utilized in the present study allows visualization of cells that

are stimulated to produce CRH during development or in response to an experimental treatment,

but is not indicative of the quantity of CRH. Regarding the possibility of ectopic expression, the

reliable colocalization of tdTomato with CRH mRNA has been previously validated using

various methods (Wamsteeker Cusulin et al., 2013; Chen et al., 2015; Pleil et al., 2015),

suggesting that increases in the number of cells that transcribe CRH in response to a stimulus can

be observed, quantified, and compared to controls. Reporting for CRH transcription following

Cre/lox-mediated recombination is accomplished through ROSA-26 promoter-driven tdTomato

expression, and consequently, the fluorescent indicator is understood to be persistent, meaning

that downregulation of CRH expression will not be accompanied by a downregulation of

tdTomato. While downregulation of CRH mRNA in response to salt-loading is a known and

interesting phenomenon (Watts, 1992a; Watts et al., 1995; Amaya et al., 2001), we designed our

experiments to focus on potential upregulation of CRH in vasopressin, oxytocin, and

Page 130: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

130

preautonomic neurons. Thus, while further characterizing the rearrangement of CRH in the PVN

during salt-loading was the impetus for the current study, the primary goals were to use the

improved spatial resolution of the CRH reporter mouse to isolate specific PVN cell-types that

increased CRH expression. It is true that in terms of CRH expression, spatial precision was

prioritized over temporal precision in that a tdTomato-positive neuron may have expressed CRH

at any point and to any extent; however, comparison of the total number of colocalizations

relative to water-drinking controls is indicative of CRH upregulation in specific cellular

phenotypes.

Another technical consideration that is important to the interpretation of the

colocalization data presented in Chapter 5 is the potential for changing numbers of vasopressin-

and oxytocin-positive neurons which might contribute to the colocalization counts presented in

Figures 5-3 to 5-7. The overall purpose of counting colocalizations in the PVN and SON was to

test the hypothesis that chronic salt-loading causes CRH expression to increase in magnocellular

oxytocin neurons, a hypothesis that was later extended to include magnocellular vasopressin

neurons after it was determined that the data did not support our hypothesis. Qualitative

assessment of the images and quantification of colocalized neurons that were labeled for

oxytocin and tdTomato indicate that salt-loading did not increase the number of

tdTomato/oxytocin colocalizations. However, as discussed elsewhere, the mouse PVN is

characterized by less segregation between magnocellular and parvocellular neurons and the

fluorescent indication for CRH is not known to extinguish when CRH transcription rates go

down or stop. Thus, interpretation of any colocalization becomes more complicated. If there

were changes, however small, in the total number of oxytocin neurons or the total number of

tdTomato neurons in salt-loaded mice compared with controls, then it may be that a change in

Page 131: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

131

the number of colocalizations would reflect this difference, but this need not be the case. No

differences in colocalization of oxytocin and tdTomato was detected through repeated random

sampling of several specific areas of the PVN in either group during a statistical test of

significance. We similarly assessed vasopressin- and tdTomato-expressing neurons and found

significant increases in colocalization which, in contrast to the investigation regarding oxytocin,

brought up an important question of whether it was CRH neurons making vasopressin or

vasopressin neurons making CRH. However, from a functional perspective the real question is

whether or not salt-loading is increasing CRH expression in neurons projecting to the posterior

pituitary or increasing vasopressin expression in neurons projecting to the median eminence

which might be answered through future experiments utilizing tract tracing of those projections.

As discussed in Chapter 5, it is likely that increased CRH in magnocellular projections to the

pituitary may potentiate vasopressin or oxytocin release (Bondy and Gainer, 1989) while

increased vasopressin in median eminence projecting neurons may potentiate ACTH release

(Abousamra et al., 1987). Because exposure to stress has likely already caused the reporting for

CRH in most if not all of the parvocellular median eminence-projecting neurons which cannot

subsequently decrease, tdTomato is likely going up or staying the same in the population of

magnocellular vasopressin neurons. Regarding the possibility of vasopressin increasing in

parvocellular median eminence-projecting neurons, total vasopressin counts might provide

additional information as to the nature of the colocalized neurons, but would not be a good

substitute for isolating neurons based on their projection targets, as salt-loading increased

vasopressin in another population of hindbrain-projecting neurons in this study. Essentially, the

colocalization data is best interpreted as an indication of possible functional reconfigurations

Page 132: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

132

occurring in the PVN, and these indications require additional experiments for conclusive

evidence concerning functionality.

The fluorescent indication of CRH-positive neurons was first used in the chronic salt-

loading study to identify PVN neurons that were also activated by salt-loading or restraint. In

contrast to our acute salt-loading experiment, we found that chronic salt-loading was not

associated with a decrease in restraint-induced Fos-positive nuclei specifically within neurons

that reported for CRH transcription. The results of Krause et al. and Frazier et al. found that in

acute salt-loaded rats, restraint-induced CORT was attenuated (Krause et al., 2011a; Frazier et

al., 2013), which was also a finding here in acutely salt-loaded mice. As discussed above, there

were several indications that this effect was centrally-mediated, including reduced ACTH, an

inhibitory tone on putative CRH neurons, and reduced Fos activation in CRH neurons.

However, similar restraint-induced Fos between chronic salt-loaded mice and controls is not

sufficient evidence to conclude that restraint-induced HPA axis activation is not inhibited during

chronic salt-loading. On the contrary, the results of Chapter 5 suggested that the recovery to

baseline levels of CORT after restraint stress occurs faster in chronic salt-loaded mice.

Additional mechanisms for HPA axis inhibition exist both upstream and downstream of CRH

neurons that are likely influencing levels of ACTH and CORT during the course of the extended

osmotic challenge. Previous reports in rats have shown that the rise and fall of CORT which

normally occurs in a circadian rhythm becomes dysregulated during chronic salt-loading by

eliminating the daily nadir (Watts, 1992a). Circadian rhythms are controlled by neurons in the

suprachiasmatic nucleus which changes transcriptional activity in response to salt-loading (Watts

et al., 1995), suggesting that information regarding external and internal cues that are sent from

the suprachiasmatic nucleus to the PVN (Watts et al., 1987; Vujovic et al., 2015) conveying the

Page 133: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

133

appropriate timing for low activity of CRH neurons may also be dysregulated. It is possible that

this dysregulation in a circadian regulatory circuit generates heightened activity of CRH neurons

at all times, and this effect supersedes an inhibitory input generated by the osmosensitive circuits

converging on the PVN CRH neurons during chronic salt-loading. Apart from central

mechanisms, it is also possible that the HPA axis may be attenuated while restraint-induced Fos

was unaffected through negative feedback involving actions in the pituitary. The incongruence

of the activity of CRH neurons with our results showing an attenuated HPA axis with chronic

salt-loading warrants further investigation.

Chronic salt-loading without restraint was associated with a significantly increased

number of Fos-positive nuclei in the caudal portion of the PVN. This region contains a large

population of neurons projecting to hindbrain and spinal cord autonomic centers (Biag et al.,

2012). The method of phenotyping PVN neurons based on position in the rostral-caudal extent is

a very general tool useful for a broad indication of the types of neurons involved with the

response to a particular stimulus. However, immunohistochemical and tract tracing molecules

directly provide evidence related to the function and destination of neuronal projections. Our

hypothesis that restraint-induced Fos in CRH neurons would be inhibited with chronic salt-

loading was not supported by the data we collected. However, the pattern of Fos activation

without respect to phenotypical marker provided some insight as to which populations of

neurons might be activated. These neurons were then specifically isolated in the experiments of

Chapter 5, which found that there was an increase in markers for vasopressin and CRH in

neurons projecting to the RVLM. Future experiments are necessary to determine the

consequences of this increase in expression.

Page 134: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

134

One of the most striking differences between what we observed in the PVN and SON of

salt-loaded mice and previously published studies using rats was the lack of a robust alteration in

CRH expression. Our findings did include significant changes in digoxigen-labeled hybridized

probe as well as increases in colocalization of tdTomato reporting for CRH with vasopressin and

FG. However, based on the majority of the findings in the literature, we hypothesized that the

increase in CRH expression in magnocellular PVN and SON would be dramatic in a visually-

noticeable way. It may be that the other differences between what we found in mice and

previously published findings in rats (Watts et al., 1995) may provide insight into this disparity.

Most chronic salt-loading experiments included results that indicated dehydration-induced

anorexia, that is, coexistent markers for hypovolemia such as increased hematocrit along with

reduced body weight. Chronic salt-loading is associated with increased CRH mRNA expression

in the lateral hypothalamus, a region of the brain that is part of a feeding behavior network

(Watts, 1999), receives afferent projections from osmosensitive neurons located in the SFO

(Miselis, 1981), and sends efferent projections to the PVN (Watts, 1999). It is therefore possible

that without an increase in angiotensin-II caused by hypovolemia, and associated stimulation of

SFO neurons projecting to the lateral hypothalamus, the stimulation and alteration of anorexic

signals in the lateral hypothalamus may not occur. In the same way, a lack of stimulation in the

lateral hypothalamus could result in a lack of input to PVN neurons which may be involved with

a dramatic increase in CRH expression with salt-loading. Furthermore, knife cuts dissociating

the OVLT from the PVN prevented chronic salt-loading induced rearrangement of CRH

suggesting that signals in the plasma are necessary for causing the change (Kovacs and

Sawchenko, 1993). Together it seems that without adequate stimulation by hypovolemia-related

plasma-born substances acting in the circumventricular organs, neuropeptide rearrangement may

Page 135: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

135

not occur or at least not occur in a dramatic way. Alternatively, there may be a species-related

difference that would prevent any such rearrangement from occurring under any circumstances,

and future studies would be necessary to confirm this.

Our results show that in mice CRH mRNA expression following five days of drinking

2% NaCl differs from the expression previously reported in rats and illustrates that CRH is

involved in both the neuroendocrine and autonomic reconfiguration caused by chronic salt-

loading. Our results together with previous reports (Amaya et al., 2001) suggest that salt-loading

reduces parvocellular CRH mRNA expression and increases expression of CRH and vasopressin

in neurons that project to the RVLM. The major implication is that this shift in expression is

coupled to increased sympathetic drive that may either maintain or elevate blood pressure during

the salt-loading period, and may contribute to long-lasting cardiovascular pathology. Sustained

exposure to either a stressful environment or chronic elevation of the pNa+ results in significant

risk for the development of psychiatric disorders or cardiovascular disease, respectively

(Chrousos and Gold, 1992; Strazzullo et al., 2009). Thus, the mechanisms that support the

changes in expression detailed here along with their respective components may be viable

therapeutic targets for intervention in certain disease states. The mechanisms that contribute to

the plasticity likely include changes on the transcriptional level. Increased transcription for CRH

may be induced via steroid signaling, as chronic salt-loading increases ACTH and CORT after

the first day before becoming suppressed (Elias et al., 2002). Activation of glucocorticoid

receptors in PVN parvocellular CRH neurons generally represents negative feedback of the HPA

axis (Herman et al., 2012); however, in other PVN neurons and areas of the brain, glucocorticoid

receptor activation increases CRH expression (Swanson and Simmons, 1989) in a similar way to

the increase caused by water deprivation (Watts and Sanchezwatts, 1995). It may be that the

Page 136: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

136

increased number of RVLM-projecting neurons reporting for CRH in our study was also

glucocorticoid-mediated early in the salt-loading period. As discussed previously, additional

influences from osmosensitive regions of the brain also contribute to altered CRH expression and

the relationship between these influences and steroid signaling should be included in future

endeavors.

There are interesting similarities between the neuropeptide plasticity found in the current

study and that observed in postmortem tissue obtained from patients with cardiovascular or

affective disorders. For example, the increased colocalization of CRH and vasopressin that

occurred with salt-loading is also observed in patients that have died from complications arising

from chronic heart failure. Specifically, patients that died from chronic heart failure have twice

as many neurons colocalizing CRH and vasopressin in the PVN when compared to that of

controls (Sivukhina et al., 2009). Relative to normotensive individuals, patients that have died

from complications due to essential hypertension exhibit a profound increase in CRH mRNA

within discrete regions of the PVN (Goncharuk et al., 2007); alterations that are evocative of the

CRH plasticity that we found within PVN preautonomic neurons as a consequence of salt-

loading. Cardiovascular disease and affective disorders are highly comorbid (Vogelzangs et al.,

2010), and intriguingly, affective disorders are also accompanied by neuroendocrine and

autonomic adaptations predictive of neuropeptide plasticity in the PVN. That is, affective

disorders are associated with increased CRH and vasopressin expression in the PVN and SON as

well as sympatho-vagal imbalance (Raadsheer et al., 1994; Meynen et al., 2006; Kemp et al.,

2012). Here, we propose that chronic challenges to body fluid homeostasis lead to neuropeptide

plasticity in hypothalamic nuclei that promote endocrine and autonomic alterations that mitigate

Page 137: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

137

the challenge, but progressively contribute to the etiology of cardiovascular and affective

disorders.

Our results show that in mice CRH mRNA expression following five days of drinking

2% NaCl differs from the expression previously reported in rats and further illustrates that CRH

is involved in both the neuroendocrine and autonomic reconfiguration caused by chronic salt-

loading. However, more detailed cellular investigation revealed that salt-loading increased CRH

expression in putative magnocellular vasopressinergic neurons of the PVN and increased

expression of CRH and vasopressin-NP in neurons that project to the RVLM. The major

implication is that this shift in expression is coupled to increased vasopressin release and

sympathetic drive that not only maintains blood tonicity and pressure during the salt-loading, but

also may contribute to the etiology of cardiovascular and affective disorders.

Closing Remarks

Relatively free, widespread access to potable water and an excess of salt are

contemporary additions to most economically advanced societies, representing only a small

fraction of human evolution. This suggests that the hydration-related instincts that for millennia

guided behavior, may be as limited in their usefulness to modern man as many misappropriated

aspects of the stress response. However, for many non-human species the centrally-mediated

mechanisms connecting hydromineral homeostasis with anxiety levels may have developed to

improve the likelihood of survival. Activation of the RAAS is associated with physiological

states such as sodium and volume depletion, situations which would also have the potential to

produce an anxious state. For example, species that do not consume meat have a dietary

imperative to locate and consume salt. Freely accessible salt may represent isolated areas that

may also be known to predators as a successful hunting area. Thus, the low-sodium state and

locating of a salt source may be hazardous and necessitate the increased alertness of heightened

Page 138: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

138

anxiety. Furthermore, a mild hypernatremic state is relieved primarily by approaching a water

source that may also be frequented by a variety of species implying that the increased sociability

and decreased anxiety that accompanies central oxytocin release may be of survival value as

well. It may be that these evolutionarily conserved systems as well as the circuits that interact

with them may become dysregulated by a variety of genetic and environmental influences.

The understanding of a pathological process obviously begins with a clear understanding

of normal physiological function. Indeed, the development of our understanding of the RAAS

occupies a large segment of modern medical history which has revolutionized pharmacological

treatment for hypertension. Pushing the boundaries of this knowledge has uncovered local

tissue-specific systems with RAAS components as well as a counter-regulatory limb that will

likely result in novel therapeutic targets. Furthermore, it is clear that increased activation of the

RAAS promotes anxiety and potentiates the response to a psychogenic stressor. While existing

drugs have been shown to counteract these effects, it is still unclear exactly how these off-target

effects are accomplished. Similarly, oxytocin and vasopressin have well-characterized endocrine

effects in the body, but are just beginning to be thought of as centrally acting mediators of

behavior. Together, the mechanisms of salt and water homeostasis represent a toolbox for

uncovering what may be fundamental central processes with unrealized potential.

The primary advantages of studying the central effects of both the RAAS and the

neuropeptides oxytocin and vasopressin is that their systemic roles in normal physiological

function have been extensively studied and the routes by which they activate brain circuits are

well-understood. With these rigorously reviewed research findings at hand, it is possible to

manipulate the normal homeostatic mechanisms governing hydromineral balance for the purpose

of studying the effect on the brain circuits involved. Often this can be accomplished without the

Page 139: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

139

potential confounds of introducing a foreign substance into the brain or facing the many

drawbacks of injecting a receptor blocker or agonist. Moreover, hydromineral balance is

primarily regulated by a complex series of behavioral adjustments involving motivation, reward,

social function, problem solving, and risk assessment. Given the diversity and far-reaching

implications of these areas of research, there exists a translational potential for basic science

understanding targeted at further revealing the mechanistic relationships between homeostatic

and stress-responsive systems.

The unpleasant anxious state associated with water deprivation is an obvious example of

how mood can be governed by hydromineral imbalance; however, the idea that the neurological

basis of the cravings for water, or salt for that matter, is somehow also the basis for the

pathogenesis of a clinical psychiatric disorder is beyond the scope of the present work.

Furthermore, to contend that the treatment of anxiety disorders, or psychiatric disorders in

general, can be trivialized to simply a monitoring and responding to the hydration state of tissues

is also not being suggested. In the extreme, such a connection does exist however, in the case of

the very serious condition of polydipsic schizophrenia in which patients are susceptible to water

intoxication (Goldman, Gnerlich, & Hussain, 2007). Less severe and more common are the

anecdotal cases of poor eating and drinking habits that result in transient states of subclinical

malaise due to mild dehydration or dysnatremias, as well as the relief that comes from satiating

thirst or satisfying a craving for salty food. Again, it would be easy to make the connection that

individuals use water or salt intake to self-medicate, consciously or unconsciously, in an effort to

alleviate the subjective effects of stress. While these are interesting and plausible ideas, none of

them are directly addressed here; rather, experimental evidence for mechanisms and circuits that

are reliably influenced by treatments involving water or salt are analyzed for the purpose of

Page 140: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

140

understanding the parameters in which normal and dysfunctional stress-responsiveness occurs

and associated measures of anxiety are altered.

Future Directions

A major finding of these experiments was that the robust alteration in chronic salt-

loading induced CRH expression reported previously using a rat model did not occur in a

dramatic way using a mouse model. Therefore, repeating the experiments with increased

concentrations of saline to determine if the osmoregulatory capacity of mice prevented the

dynamic changes in CRH expression reported in rats is an important experiment.

We found that there was an increase in tdTomato/vasopressin colocalizations with salt-

loading. Future experiments will determine conclusively whether it is vasopressin increasing in

CRH parvocellular neurons or CRH increasing in vasopressin magnocellular neurons. One such

set of experiments would systematically measure soma diameter to classify neurons as

magnocellular or parvocellular. These experiments might include a method or methods that

could detect decreases in CRH such as standard IHC and in situ hybridization.

The time course of Cre expression is related to the expression of tdTomato reporting for

CRH which is increased with restraint. It may be possible to begin to determine this time course

experimentally in vitro by stimulating PVN neurons to induce tdTomato expression. For

example, this could be accomplished with time-lapse microscopy in an in vitro preparation of a

hypothalamic section bathed in artificial cerebrospinal fluid. While it has been determined that

CRH mRNA increases in response to restraint, it has not been determined experimentally how

long after a restraint-induced increase in CRH mRNA it takes for a rise in Cre and tdTomato. It

would be best to perform these experiments in young mice that had been protected from

exposure to stress as much as possible to limit the number of tdTomato neurons in the PVN.

While not a specific contribution to furthering the understanding of PVN plasticity that this

Page 141: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

141

dissertation embodied, the technical advancement would include a better understanding of the

capabilities of a relatively new transgenic mouse model. Future experiments utilizing reporter

mice would benefit from knowing how long after stimulation it takes to induce Cre-mediated

excision of the STOP codon and subsequent expression of tdTomato.

Further experimentation will be aimed at exploring the mechanistic relationship between

increased glutamatergic innervation to the PVN with salt-loading (Fig. 6-1) and altered

neuropeptide expression. Qualitative images of astrocytic coverage suggest that retraction of

astrocytic processes during dehydration may influence synaptic physiology related to the

induction of plasticity in the PVN (Fig. 6-2).

Page 142: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

142

Figure 6-1. Increased vesicular glutamate transporter 2 (vGlut-2) appositions on tdTomato

neurons in the PVN with salt-loading. The dorsal-medial parvocellular region of the

PVN (A) labeled for vGlut-2 (magenta) in CRH reporter mouse tissue (B). High

magnification images in the hydrated state (C1) and salt-loaded state (C2) were

analyzed and quantified. A significant increase in vGlut-2 appositions were found on

tdTomato neurons in the PVN (D).

A B

C1

C2

D

Page 143: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

143

Figure 6-2. Decreased labeling for the astrocytic marker glial fibrillary acidic protein (GFAP)

around tdTomato reporting for CRH in the PVN with salt-loading.

B

A

Page 144: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

144

LIST OF REFERENCES

Abousamra AB, Harwood JP, Catt KJ, Aguilera G. 1987. Mechanisms of Action of Crf and

Other Regulators of Acth Release in Pituitary Corticotrophs. Annals of the New York

Academy of Sciences 512:67-84.

Adachi T, Hisano S, Daikoku S. 1985. Intragranular Colocalization of Immunoreactive

Methionine Enkephalin and Oxytocin within the Nerve-Terminals in the Posterior

Pituitary. Journal of Histochemistry & Cytochemistry 33(9):891-899.

Aguilera G, Kiss A, Luo X. 1995a. Increased Expression of Type 1 Angiotensin II Receptors in

the Hypothalamic Paraventricular Nucleus following Stress and Glucocorticoid

Administration. Journal of Neuroendocrinology 7(10):775-783.

Aguilera G, Rabadan-Diehl C. 2000. Vasopressinergic regulation of the hypothalamic-pituitary-

adrenal axis: implications for stress adaptation. Regul Pept 96(1-2):23-29.

Aguilera G, Young WS, Kiss A, Bathia A. 1995b. Direct regulation of hypothalamic

corticotropin-releasing-hormone neurons by angiotensin II. Neuroendocrinology

61(4):437-444.

Akmayev IG. 1971. Morphological aspects of the hypothalamic-hypophyseal system. II.

Functional morphology of pituitary microcirculation. Z Zellforsch Mikrosk Anat

116(2):178-194.

Akmayev IG, Popov AP. 1977. Morphological Aspects of Hypothalamo-Hypophyseal System .7.

Tanycytes - Their Relation to Hypophyseal Adrenocorticotropic Function -

Ultrastructural-Study. Cell and Tissue Research 180(2):263-282.

Albrecht D. 2010. Physiological and pathophysiological functions of different angiotensins in the

brain. British Journal of Pharmacology 159(7):1392-1401.

Amaya F, Tanaka M, Hayashi S, Tanaka Y, Ibata Y. 2001. Hypothalamo-pituitary-adrenal axis

sensitization after chronic salt loading. Neuroendocrinology 73(3):185-193.

Amaya F, Tanaka M, Tamada Y, Tanaka Y, Nilaver G, Ibata Y. 1999. The influence of salt

loading on vasopressin gene expression in magno- and parvocellular hypothalamic

neurons: an immunocytochemical and in situ hybridization analysis. Neuroscience

89(2):515-523.

Andari E, Duhamel JR, Zalla T, Herbrecht E, Leboyer M, Sirigu A. 2010. Promoting social

behavior with oxytocin in high-functioning autism spectrum disorders. Proceedings of the

National Academy of Sciences of the United States of America 107(9):4389-4394.

Anil Kumar K, Nagwar S, Thyloor R, Satyanarayana S. 2014. Anti-stress and nootropic activity

of drugs affecting the renin-angiotensin system in rats based on indirect biochemical

evidence. Journal of the renin-angiotensin-aldosterone system : JRAAS.

Page 145: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

145

Antunes-Rodrigues J, Ruginsk SG, Mecawi AS, Margatho LO, Cruz JC, Vilhena-Franco T, Reis

WL, Ventura RR, Reis LC, Vivas LM, Elias LL. 2013. Mapping and signaling of neural

pathways involved in the regulation of hydromineral homeostasis. Brazilian journal of

medical and biological research. 46(4):327-338.

Aoyagi T, Koshimizu TA, Tanoue A. 2009. Vasopressin regulation of blood pressure and

volume: findings from V1a receptor-deficient mice. Kidney Int 76(10):1035-1039.

Aplin KP, Suzuki H, Chinen AA, Chesser RT, Ten Have J, Donnellan SC, Austin J, Frost A,

Gonzalez JP, Herbreteau V, Catzeflis F, Soubrier J, Fang YP, Robins J, Matisoo-Smith E,

Bastos AD, Maryanto I, Sinaga MH, Denys C, Van Den Bussche RA, Conroy C, Rowe

K, Cooper A. 2011. Multiple geographic origins of commensalism and complex dispersal

history of Black Rats. Plos One 6(11):e26357.

Arborelius L, Owens MJ, Plotsky PM, Nemeroff CB. 1998. The role of corticotropin-releasing

factor in depression and anxiety disorders. The Journal of endocrinology 160(1):1-12.

Armario A. 2006. The Hypothalamic-Pituitary-Adrenal Axis: What can it Tell us About

Stressors? CNS & Neurological Disorders - Drug Targets 5(5):485-501.

Atkinson IAE. 1985. The spread of commensal species of Rattus to oceanic islands and their

effects on island avifaunas. International Council for Bird Preservation Technical

Publication:35-81.

Badaue-Passos D, Jr., Godino A, Johnson AK, Vivas L, Antunes-Rodrigues J. 2007. Dorsal

raphe nuclei integrate allostatic information evoked by depletion-induced sodium

ingestion. Exp Neurol 206(1):86-94.

Baertschi AJ, Vallet PG. 1981. Osmosensitivity of the hepatic portal vein area and vasopressin

release in rats. The Journal of physiology 315:217-230.

Bains JS, Ferguson AV. 1993. Paraventricular nucleus neurons projecting to the spinal cord in

the rat are influenced by subfornical organ stimulation. Society for Neuroscience

Abstracts 19(1-3):956-956.

Bancila M, Giuliano F, Rampin O, Mailly P, Brisorgueil MJ, Calas A, Verge D. 2002. Evidence

for a direct projection from the paraventricular nucleus of the hypothalamus to putative

serotoninergic neurons of the nucleus paragigantocellularis involved in the control of

erection in rats. The European journal of neuroscience 16(7):1240-1248.

Bardgett ME, Chen QH, Guo Q, Calderon AS, Andrade MA, Toney GM. 2014. Coping with

dehydration: sympathetic activation and regulation of glutamatergic transmission in the

hypothalamic PVN. Am J Physiol Regul Integr Comp Physiol 306(11):R804-813.

Page 146: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

146

Barnes V, Schneider R, Alexander C, Staggers F. 1997. Stress, stress reduction, and hypertension

in African Americans: an updated review. Journal of the National Medical Association

89(7):464-476.

Beinfeld MC, Meyer DK, Brownstein MJ. 1980. Cholecystokinin Octapeptide in the Rat

Hypothalamo-Neurohypophyseal System. Nature 288(5789):376-378.

Ben-Barak Y, Russell JT, Whitnall MH, Ozato K, Gainer H. 1985. Neurophysin in the

hypothalamo-neurohypophysial system. I. Production and characterization of monoclonal

antibodies. J Neurosci 5(1):81-97.

Benarroch EE. 2005. Paraventricular nucleus, stress response, and cardiovascular disease.

Clinical autonomic research : official journal of the Clinical Autonomic Research Society

15(4):254-263.

Benarroch EE. 2013a. Oxytocin and vasopressin Social neuropeptides with complex

neuromodulatory functions. Neurology 80(16):1521-1528.

Bernard C. 1949. An introduction to the study of experimental medicine. New York : Henry

Schuman, Inc.

Biag J, Huang Y, Gou L, Hintiryan H, Askarinam A, Hahn JD, Toga AW, Dong HW. 2012.

Cyto- and chemoarchitecture of the hypothalamic paraventricular nucleus in the

C57BL/6J male mouse: a study of immunostaining and multiple fluorescent tract tracing.

J Comp Neurol 520(1):6-33.

Biancardi VC, Stranahan AM, Krause EG, de Kloet AD, Stern JE. 2016. Cross talk between AT1

receptors and Toll-like receptor 4 in microglia contributes to angiotensin II-derived ROS

production in the hypothalamic paraventricular nucleus. Am J Physiol Heart Circ Physiol

310(3):H404-415.

Biller KJ, Unwin RJ, Shirley DG. 2000. Distal tubular electrolyte transport during inhibition of

renal 11beta-hydroxysteroid dehydrogenase. American journal of physiology Renal

physiology 280(1):9.

Bledsoe RK, Madauss KP, Holt JA, Apolito CJ, Lambert MH, Pearce KH, Stanley TB, Stewart

EL, Trump RP, Willson TM, Williams SP. 2005. A Ligand-mediated Hydrogen Bond

Network Required for the Activation of the Mineralocorticoid Receptor. Journal of

Biological Chemistry 280(35):31283-31293.

Boler J, Enzmann F, Folkers K, Bowers CY, Schally AV. 1969. The identity of chemical and

hormonal properties of the thyrotropin releasing hormone and pyroglutamyl-histidyl-

proline amide. Biochem Biophys Res Commun 37(4):705-710.

Page 147: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

147

Bondy CA, Gainer H. 1989. Corticotrophin-Releasing Hormone Stimulates Neurohypophyslal

Hormone Release through an Interaction with the Intermediate Lobe of the Pituitary. J

Neuroendocrinol 1(1):5-8.

Bondy CA, Whitnall MH, Brady LS, Gainer H. 1989. Coexisting Peptides in Hypothalamic

Neuroendocrine Systems - Some Functional Implications. Cell Mol Neurobiol 9(4):427-

446.

Bosch OJ, Dabrowska J, Modi ME, Johnson ZV, Keebaugh AC, Barrett CE, Ahern TH, Guo J,

Grinevich V, Rainnie DG, Neumann ID, Young LJ. 2016. Oxytocin in the nucleus

accumbens shell reverses CRFR2-evoked passive stress-coping after partner loss in

monogamous male prairie voles. Psychoneuroendocrinology 64:66-78.

Bota M, Sporns O, Swanson LW. 2012. Neuroinformatics analysis of molecular expression

patterns and neuron populations in gray matter regions: the rat BST as a rich exemplar.

Brain Res 1450:174-193.

Bourque CW. 2008. Central mechanisms of osmosensation and systemic osmoregulation. Nat

Rev Neurosci 9(7):519-531.

Boyle CN, Lorenzen SM, Compton D, Watts AG. 2012. Dehydration-anorexia derives from a

reduction in meal size, but not meal number. Physiol Behav 105(2):305-314.

Brazeau P, Ling N, Bohlen P, Esch F, Ying SY, Guillemin R. 1982. Growth-Hormone

Releasing-Factor, Somatocrinin, Releases Pituitary Growth-Hormone Invitro.

Proceedings of the National Academy of Sciences of the United States of America-

Biological Sciences 79(24):7909-7913.

Burgus R, Dunn TF, Desiderio D, Ward DN, Vale W, Guillemin R. 1970. Characterization of

ovine hypothalamic hypophysiotropic TSH-releasing factor. Nature 226(5243):321-325.

Burgus R, Ling N, Butcher M, Guillemi.R. 1973. Primary Structure of Somatostatin -

Hypothalamic Peptide That Inhibits Secretion of Pituitary Growth-Hormone. Proceedings

of the National Academy of Sciences of the United States of America 70(3):684-688.

Busnardo C, Alves FHF, Crestani CC, Scopinho AA, Resstel LBM, Correa FMA. 2013.

Paraventricular nucleus of the hypothalamus glutamate neurotransmission modulates

autonomic, neuroendocrine and behavioral responses to acute restraint stress in rats.

European Neuropsychopharmacology 23(11):1611-1622.

Byrum CE, Guyenet PG. 1987. Afferent and efferent connections of the A5 noradrenergic cell

group in the rat. J Comp Neurol 261(4):529-542.

Caeiro X, Vivas L. 2008. beta-Endorphin in the median preoptic nucleus modulates the pressor

response induced by subcutaneous hypertonic sodium chloride. Exp Neurol 210(1):59-66.

Page 148: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

148

Calhoun DA. 1992. Hypertension in blacks: socioeconomic stress and sympathetic nervous

system activity. Am J Med Sci 304(5):306-311.

Calhoun DA, Oparil S. 1995. Racial differences in the pathogenesis of hypertension. Am J Med

Sci 310 Suppl 1:S86-90.

Cardoso C, Ellenbogen MA, Orlando MA, Bacon SL, Joober R. 2013. Intranasal oxytocin

attenuates the cortisol response to physical stress: a dose-response study.

Psychoneuroendocrinology 38(3):399-407.

Castel M, Morris JF. 1988. The neurophysin-containing innervation of the forebrain of the

mouse. Neuroscience 24(3):937-966.

Chan-Palay V, Zaborszky L, Kohler C, Goldstein M, Palay SL. 1984. Distribution of tyrosine-

hydroxylase-immunoreactive neurons in the hypothalamus of rats. J Comp Neurol

227(4):467-496.

Chen J, Gomez-Sanchez CE, Penman A, May PJ, Gomez-Sanchez E. 2014. Expression of

mineralocorticoid and glucocorticoid receptors in preautonomic neurons of the rat

paraventricular nucleus. Am J Physiol Regul Integr Comp Physiol 306(5):R328-340.

Chen Y, Molet J, Gunn BG, Ressler K, Baram TZ. 2015. Diversity of Reporter Expression

Patterns in Transgenic Mouse Lines Targeting Corticotropin-Releasing Hormone-

Expressing Neurons. Endocrinology 156(12):4769-4780.

Christensen EI, Grann B, Kristoffersen IB, Skriver E, Thomsen JS, Andreasen A. 2014a. Three-

dimensional reconstruction of the rat nephron. Am J Physiol Renal Physiol 306(6):F664-

671.

Chrousos GP, Gold PW. 1992. The concepts of stress and stress system disorders. Overview of

physical and behavioral homeostasis. Jama 267(9):1244-1252.

Clarke IJ, Cummins JT. 1982. The Temporal Relationship between Gonadotropin-Releasing

Hormone (Gnrh) and Luteinizing-Hormone (Lh) Secretion in Ovariectomized Ewes.

Endocrinology 111(5):1737-1739.

Coote JH. 1995. Cardiovascular function of the paraventricular nucleus of the hypothalamus.

Biological signals 4(3):142-149.

Coote JH. 2005. A role for the paraventricular nucleus of the hypothalamus in the autonomic

control of heart and kidney. Exp Physiol 90(2):169-173.

Cotterill RM. 1985. Fever in autistics. Nature 313(6002):426.

Page 149: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

149

Cowley AW. 1997. Role of the renal medulla in volume and arterial pressure regulation.

American Journal of Physiology-Regulatory Integrative and Comparative Physiology

273(1):R1-R15.

Curran LK, Newschaffer CJ, Lee LC, Crawford SO, Johnston MV, Zimmerman AW. 2007.

Behaviors associated with fever in children with autism spectrum disorders. Pediatrics

120(6):e1386-1392.

Dabrowska J, Hazra R, Ahern TH, Guo JD, McDonald AJ, Mascagni F, Muller JF, Young LJ,

Rainnie DG. 2011. Neuroanatomical evidence for reciprocal regulation of the

corticotrophin-releasing factor and oxytocin systems in the hypothalamus and the bed

nucleus of the stria terminalis of the rat: Implications for balancing stress and affect.

Psychoneuroendocrinology 36(9):1312-1326.

Daniels D, Mietlicki EG, Nowak EL, Fluharty SJ. 2009. Angiotensin II stimulates water and

NaCl intake through separate cell signalling pathways in rats. Exp Physiol 94(1):130-137.

Davis M, Rainnie D, Cassell M. 1994. Neurotransmission in the rat amygdala related to fear and

anxiety. Trends Neurosci 17(5):208-214.

de Kloet ER, Joels M, Holsboer F. 2005. Stress and the brain: from adaptation to disease. Nat

Rev Neurosci 6(6):463-475.

De Luca LA, Menani JV, Johnson AK. 2013. Neurobiology of Body Fluid Homeostasis:

Transduction and Integration: Taylor & Francis Group.

De Vries GJ. 2004. Minireview: Sex differences in adult and developing brains: compensation,

compensation, compensation. Endocrinology 145(3):1063-1068.

de Wardener HE, He FJ, MacGregor GA. 2004. Plasma sodium and hypertension. Kidney Int

66(6):2454-2466.

Dimitrov EL, Yanagawa Y, Usdin TB. 2013. Forebrain GABAergic projections to locus

coeruleus in mouse. J Comp Neurol 521(10):2373-2397.

Dodhia S, Hosanagar A, Fitzgerald DA, Labuschagne I, Wood AG, Nathan PJ, Phan KL. 2014.

Modulation of resting-state amygdala-frontal functional connectivity by oxytocin in

generalized social anxiety disorder. Neuropsychopharmacology 39(9):2061-2069.

Domes G, Heinrichs M, Kumbier E, Grossmann A, Hauenstein K, Herpertz SC. 2013. Effects of

intranasal oxytocin on the neural basis of face processing in autism spectrum disorder.

Biological psychiatry 74(3):164-171.

Dong HW, Petrovich GD, Watts AG, Swanson LW. 2001. Basic organization of projections

from the oval and fusiform nuclei of the bed nuclei of the stria terminalis in adult rat

brain. J Comp Neurol 436(4):430-455.

Page 150: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

150

Dong HW, Swanson LW. 2006. Projections from bed nuclei of the stria terminalis, dorsomedial

nucleus: implications for cerebral hemisphere integration of neuroendocrine, autonomic,

and drinking responses. J Comp Neurol 494(1):75-107.

Dooley CP, Valenzuela JE. 1984. Duodenal volume and osmoreceptors in the stimulation of

human pancreatic secretion. Gastroenterology 86(1):23-27.

Duchen LW. 1962. Effects of Ingestion of Hypertonic Saline on Pituitary Gland in Rat -

Morphological Study of Pars Intermedia and Posterior Lobe. Journal of Endocrinology

25(2):161-&.

Dunn AJ, Berridge CW. 1990. Physiological and behavioral responses to corticotropin-releasing

factor administration: is CRF a mediator of anxiety or stress responses? Brain research

reviews 15(2):71-100.

Duvernoy HM, Risold PY. 2007. The circumventricular organs: an atlas of comparative anatomy

and vascularization. Brain Res Rev 56(1):119-147.

Elias LL, Dorival Campos A, Moreira AC. 2002. The opposite effects of short- and long-term

salt loading on pituitary adrenal axis activity in rats. Horm Metab Res 34(4):207-211.

Englemann M, Wotjak CT, Ebner K, Landgraf R. 2000. Behavioural impact of intraseptally

released vasopressin and oxytocin in rats. Experimental Physiology 85(s1):125s-130s.

Epstein DE, Sherwood A, Smith PJ, Craighead L, Caccia C, Lin PH, Babyak MA, Johnson JJ,

Hinderliter A, Blumenthal JA. 2012. Determinants and consequences of adherence to the

dietary approaches to stop hypertension diet in African-American and white adults with

high blood pressure: results from the ENCORE trial. J Acad Nutr Diet 112(11):1763-

1773.

Fang Z, Carlson SH, Peng N, Wyss JM. 2000. Circadian rhythm of plasma sodium is disrupted

in spontaneously hypertensive rats fed a high-NaCl diet. Am J Physiol Regul Integr

Comp Physiol 278(6):R1490-1495.

Ferguson AV, Latchford KJ, Samson WK. 2008. The paraventricular nucleus of the

hypothalamus - a potential target for integrative treatment of autonomic dysfunction.

Expert Opin Ther Targets 12(6):717-727.

Fisher JP, Paton JF. 2012. The sympathetic nervous system and blood pressure in humans:

implications for hypertension. J Hum Hypertens 26(8):463-475.

Fitzsimons JT. 1998. Angiotensin, thirst, and sodium appetite. Physiol Rev 78(3):583-686.

Page 151: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

151

Francis DD, Young LJ, Meaney MJ, Insel TR. 2002. Naturally occurring differences in maternal

care are associated with the expression of oxytocin and vasopressin (V1a) receptors:

gender differences. J Neuroendocrinol 14(5):349-353.

Franklin KBJ, Paxinos G. 2008. The Mouse Brain in Stereotaxic Coordinates, Compact Third

Edition. Amsterdam: Boston.

Frazier CJ, Pati D, Hiller H, Nguyen D, Wang L, Smith JA, MacFadyen K, de Kloet AD, Krause

EG. 2013. Acute hypernatremia exerts an inhibitory oxytocinergic tone that is associated

with anxiolytic mood in male rats. Endocrinology 154(7):2457-2467.

Freeman ME, Kanyicska B, Lerant A, Nagy G. 2000. Prolactin: structure, function, and

regulation of secretion. Physiol Rev 80(4):1523-1631.

Fuxe K, Agnati LF, Ganten D, Lang RE, Calza L, Poulsen K, Infantellina F. 1982.

Morphometric evaluation of the coexistence of renin-like and oxytocin-like

immunoreactivity in nerve cells of the paraventricular hypothalamic nucleus of the rat.

Neuroscience letters 33(1):19-24.

Gabor A, Leenen FH. 2012. Central neuromodulatory pathways regulating sympathetic activity

in hypertension. J Appl Physiol (1985) 113(8):1294-1303.

Gaillard RC, Grossman A, Gillies G, Rees LH, Besser GM. 1981. Angiotensin II stimulates the

release of ACTH from dispersed rat anterior pituitary cells. Clin Endocrinol (Oxf)

15(6):573-578.

Ganesan R, Sumners C. 1989. Glucocorticoids potentiate the dipsogenic action of angiotensin II.

Brain Res 499(1):121-130.

Gaymann W, Martin R. 1989. Immunoreactive Galanin-Like Material in Magnocellular

Hypothalamoneurohypophysial Neurons of the Rat. Cell and Tissue Research

255(1):139-147.

Geerling JC, Loewy AD. 2006. Aldosterone-sensitive neurons in the nucleus of the solitary tract:

bidirectional connections with the central nucleus of the amygdala. J Comp Neurol

497(4):646-657.

Geerling JC, Loewy AD. 2009. Aldosterone in the brain. Am J Physiol Renal Physiol

297(3):F559-576.

Geerling JC, Shin JW, Chimenti PC, Loewy AD. 2010. Paraventricular hypothalamic nucleus:

axonal projections to the brainstem. J Comp Neurol 518(9):1460-1499.

Gillies GE, Linton EA, Lowry PJ. 1982. Corticotropin Releasing Activity of the New Crf Is

Potentiated Several Times by Vasopressin. Nature 299(5881):355-357.

Page 152: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

152

Gomez-Sanchez EP, Ahmad N, Romero DG, Gomez-Sanchez CE. 2005. Is aldosterone

synthesized within the rat brain? American journal of physiology Endocrinology and

metabolism 288(2):E342-346.

Gomez RE, Cannata MA, Milner TA, Anwar M, Reis DJ, Ruggiero DA. 1993. Vasopressinergic

mechanisms in the nucleus reticularis lateralis in blood pressure control. Brain Res

604(1-2):90-105.

Goncharuk VD, Buijs RM, Swaab DF. 2007. Corticotropin-releasing hormone neurons in

hypertensive patients are activated in the hypothalamus but not in the brainstem. J Comp

Neurol 503(1):148-168.

Goodwin FJ, Ledingham JG, Laragh JH. 1970. The effects of prolonged administration of

vasopressin and oxytocin on renin, aldosterone and sodium balance in normal man. Clin

Sci 39(5):641-651.

Gouzenes L, Desarmenien MG, Hussy N, Richard P, Moos FC. 1998. Vasopressin regularizes

the phasic firing pattern of rat hypothalamic magnocellular vasopressin neurons. J

Neurosci 18(5):1879-1885.

Grinevich V, Ma XM, Verbalis J, Aguilera G. 2001. Hypothalamic pituitary adrenal axis and

hypothalamic-neurohypophyseal responsiveness in water-deprived rats. Exp Neurol

171(2):329-341.

Grippo AJ, Francis J, Beltz TG, Felder RB, Johnson AK. 2005. Neuroendocrine and cytokine

profile of chronic mild stress-induced anhedonia. Physiol Behav 84(5):697-706.

Guastella AJ, Einfeld SL, Gray KM, Rinehart NJ, Tonge BJ, Lambert TJ, Hickie IB. 2010.

Intranasal oxytocin improves emotion recognition for youth with autism spectrum

disorders. Biological psychiatry 67(7):692-694.

Guo J, Dabrowska J, Hazra R, Rainnie DG. 2013. CRF neurons in the paraventricular nucleus

and the bed nucleus of the stria terminalis: distinct physiological properties and genetic

phenotypes. Society for Neuroscience Abstract Viewer and Itinerary Planner 43.

Guyenet PG. 2006. The sympathetic control of blood pressure. Nat Rev Neurosci 7(5):335-346.

Hafner S, Baumert J, Emeny RT, Lacruz ME, Bidlingmaier M, Reincke M, Ladwig KH. 2013.

Hypertension and depressed symptomatology: a cluster related to the activation of the

renin-angiotensin-aldosterone system (RAAS). Findings from population based KORA

F4 study. Psychoneuroendocrinology 38(10):2065-2074.

Hainsworth R. 2014. Cardiovascular control from cardiac and pulmonary vascular receptors. Exp

Physiol 99(2):312-319.

Page 153: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

153

Haque M, Wilson R, Sharma K, Mills NJ, Teruyama R. 2015. Localisation of 11beta-

Hydroxysteroid Dehydrogenase Type 2 in Mineralocorticoid Receptor Expressing

Magnocellular Neurosecretory Neurones of the Rat Supraoptic and Paraventricular

Nuclei. J Neuroendocrinol 27(11):835-849.

Hartner A, Cordasic N, Klanke B, Veelken R, Hilgers KF. 2003. Strain differences in the

development of hypertension and glomerular lesions induced by deoxycorticosterone

acetate salt in mice. Nephrology, dialysis, transplantation : official publication of the

European Dialysis and Transplant Association - European Renal Association

18(10):1999-2004.

Hattangady NG, Olala LO, Bollag WB, Rainey WE. 2012. Acute and chronic regulation of

aldosterone production. Mol Cell Endocrinol 350(2):151-162.

Hattori T, Morris M, Alexander N, Sundberg DK. 1990. Extracellular oxytocin in the

paraventricular nucleus: hyperosmotic stimulation by in vivo microdialysis. Brain

Research 506(1):169-171.

He FJ, Markandu ND, MacGregor GA. 2005. Modest salt reduction lowers blood pressure in

isolated systolic hypertension and combined hypertension. Hypertension 46(1):66-70.

Heilig CW, Stromski ME, Blumenfeld JD, Lee JP, Gullans SR. 1989. Characterization of the

major brain osmolytes that accumulate in salt-loaded rats. Am J Physiol 257(6 Pt

2):F1108-1116.

Herman JP, Cullinan WE. 1997. Neurocircuitry of stress: Central control of the hypothalamo-

pituitary-adrenocortical axis. Trends in Neurosciences 20(2):78-84.

Herman JP, McKlveen JM, Solomon MB, Carvalho-Netto E, Myers B. 2012. Neural regulation

of the stress response: glucocorticoid feedback mechanisms. Brazilian journal of medical

and biological research = Revista brasileira de pesquisas medicas e biologicas /

Sociedade Brasileira de Biofisica [et al] 45(4):292-298.

Hinks GL, Poat JA, Hughes J. 1995. Changes in Hypothalamic Cholecystokinin(a) and

Cholecystokinin(B) Receptor Subtypes and Associated Neuropeptide Expression in

Response to Salt-Stress in the Rat and Mouse. Neuroscience 68(3):765-781.

Hiroyuki K, Hirohito M, Tsutomu M, Akira N. 2013. Angiotensin II blockade and renal

protection. Current pharmaceutical design 19(17):3033-3042.

Hoffman GE, Smith MS, Verbalis JG. 1993. c-Fos and related immediate early gene products as

markers of activity in neuroendocrine systems. Front Neuroendocrinol 14(3):173-213.

Page 154: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

154

Hollis JH, McKinley MJ, D'Souza M, Kampe J, Oldfield BJ. 2008. The trajectory of sensory

pathways from the lamina terminalis to the insular and cingulate cortex: a

neuroanatomical framework for the generation of thirst. Am J Physiol Regul Integr Comp

Physiol 294(4):R1390-1401.

Huang CC, Chu CY, Yeh CM, Hsu KS. 2014. Acute hypernatremia dampens stress-induced

enhancement of long-term potentiation in the dentate gyrus of rat hippocampus.

Psychoneuroendocrinology 46:129-140.

Husain A, DeSilva P, Speth RC, Bumpus FM. 1987. Regulation of angiotensin II in rat adrenal

gland. Circ Res 60(5):640-648.

Hwang BH, Wu JY, Severs WB. 1986. Effects of chronic dehydration on angiotensin II receptor

binding in the subfornical organ, paraventricular hypothalamic nucleus and adrenal

medulla of Long-Evans rats. Neuroscience letters 65(1):35-40.

Imaki T, Katsumata H, Konishi SI, Kasagi Y, Minami S. 2003. Corticotropin-releasing factor

type-1 receptor mRNA is not induced in mouse hypothalamus by either stress or osmotic

stimulation. Journal of Neuroendocrinology 15(10):916-924.

Imaki T, Katsumata H, Miyata M, Naruse M, Imaki J, Minami S. 2001. Expression of

corticotropin releasing factor (CRF), urocortin and CRF type 1 receptors in

hypothalamic-hypophyseal systems under osmotic stimulation. Journal of

Neuroendocrinology 13(4):328-338.

Insel TR, Winslow JT, Wang ZX, Young LJ. 1998. Oxytocin, vasopressin, and the

neuroendocrine basis of pair bond formation. In: Zingg HH, Bourque CW, Bichet DG,

eds. Vasopressin and Oxytocin: Molecular, Cellular, and Clinical Advances. Vol 449.

Advances in Experimental Medicine and Biology. p 215-224.

Jacobson L. 2014. Hypothalamic-pituitary-adrenocortical axis: neuropsychiatric aspects. Compr

Physiol 4(2):715-738.

Jankord R, Herman JP. 2008. Limbic regulation of hypothalamo-pituitary-adrenocortical

function during acute and chronic stress. Ann N Y Acad Sci 1148(1):64-73.

Johnson AK, Gross PM. 1993. Sensory circumventricular organs and brain homeostatic

pathways. FASEB journal : official publication of the Federation of American Societies

for Experimental Biology 7(8):678-686.

Kawabe T, Chitravanshi VC, Kawabe K, Sapru HN. 2008. Cardiovascular function of a

glutamatergic projection from the hypothalamic paraventricular nucleus to the nucleus

tractus solitarius in the rat. Neuroscience 153(3):605-617.

Page 155: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

155

Kc P, Balan KV, Tjoe SS, Martin RJ, Lamanna JC, Haxhiu MA, Dick TE. 2010. Increased

vasopressin transmission from the paraventricular nucleus to the rostral medulla

augments cardiorespiratory outflow in chronic intermittent hypoxia-conditioned rats. The

Journal of physiology 588(Pt 4):725-740.

Kc P, Haxhiu MA, Tolentino-Silva FP, Wu M, Trouth CO, Mack SO. 2002. Paraventricular

vasopressin-containing neurons project to brain stem and spinal cord respiratory-related

sites. Respir Physiol Neurobiol 133(1-2):75-88.

Keller-Wood ME, Dallman MF. 1984. Corticosteroid inhibition of ACTH secretion. Endocrine

reviews 5(1):1-24.

Kemp AH, Quintana DS, Felmingham KL, Matthews S, Jelinek HF. 2012. Depression, comorbid

anxiety disorders, and heart rate variability in physically healthy, unmedicated patients:

implications for cardiovascular risk. Plos One 7(2):e30777.

Ken'ichi Y, Hitoshi H. 2010. Changes in vasopressin release and autonomic function induced by

manipulating forebrain GABAergic signaling under euvolemia and hypovolemia in

conscious rats. Endocrine journal 58(7):559-573.

Kilcoyne MM, Hoffman DL, Zimmerman EA. 1980. Immuno-Cytochemical Localization of

Angiotensin-Ii and Vasopressin in Rat Hypothalamus - Evidence for Production in the

Same Neuron. Clinical science 59:S57-S60.

Kim SY, Adhikari A, Lee SY, Marshel JH, Kim CK, Mallory CS, Lo M, Pak S, Mattis J, Lim

BK, Malenka RC, Warden MR, Neve R, Tye KM, Deisseroth K. 2013. Diverging neural

pathways assemble a behavioural state from separable features in anxiety. Nature

496(7444):219-223.

Kirchgessner AL, Sclafani A, Nilaver G. 1988. Histochemical identification of a PVN-hindbrain

feeding pathway. Physiol Behav 42(6):529-543.

Knobloch HS, Charlet A, Hoffmann LC, Eliava M, Khrulev S, Cetin AH, Osten P, Schwarz MK,

Seeburg PH, Stoop R, Grinevich V. 2012. Evoked axonal oxytocin release in the central

amygdala attenuates fear response. Neuron 73(3):553-566.

Kolaj M, Bai D, Renaud LP. 2004. GABAB receptor modulation of rapid inhibitory and

excitatory neurotransmission from subfornical organ and other afferents to median

preoptic nucleus neurons. J Neurophysiol 92(1):111-122.

Kolaj M, Renaud LP. 2007. Presynaptic alpha-adrenoceptors in median preoptic nucleus

modulate inhibitory neurotransmission from subfornical organ and organum vasculosum

lamina terminalis. Am J Physiol Regul Integr Comp Physiol 292(5):R1907-1915.

Koletsky S. 1958. Hypertensive vascular disease produced by salt. Laboratory investigation; a

journal of technical methods and pathology 7(4):377-386.

Page 156: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

156

Koletsky S. 1959. Role of salt and renal mass in experimental hypertension. AMA archives of

pathology 68(1):11-22.

Koletsky S. 1961. Pathogenesis of experimental hypertension induced by salt. Am J Cardiol

8(4):576-581.

Kovacs KJ, Sawchenko PE. 1993. Mediation of osmoregulatory influences on neuroendocrine

corticotropin-releasing factor expression by the ventral lamina terminalis. Proceedings of

the National Academy of Sciences of the United States of America 90(16):7681-7685.

Krause EG, de Kloet AD, Flak JN, Smeltzer MD, Solomon MB, Evanson NK, Woods SC, Sakai

RR, Herman JP. 2011a. Hydration state controls stress responsiveness and social

behavior. J Neurosci 31(14):5470-5476.

Krause EG, de Kloet AD, Scott KA, Flak JN, Jones K, Smeltzer MD, Ulrich-Lai YM, Woods

SC, Wilson SP, Reagan LP, Herman JP, Sakai RR. 2011b. Blood-borne angiotensin II

acts in the brain to influence behavioral and endocrine responses to psychogenic stress. J

Neurosci 31(42):15009-15015.

Krause EG, Melhorn SJ, Davis JF, Scott KA, Ma LY, de Kloet AD, Benoit SC, Woods SC,

Sakai RR. 2008. Angiotensin type 1 receptors in the subfornical organ mediate the

drinking and hypothalamic-pituitary-adrenal response to systemic isoproterenol.

Endocrinology 149(12):6416-6424.

Kriz W. 1967. Der Architektonische Und Funktionelle Aufbau Der Rattenniere. Zeitschrift Fur

Zellforschung Und Mikroskopische Anatomie 82(4):495-&.

Kuramochi G, Kobayashi I. 2000. Regulation of the urine concentration mechanism by the

oropharyngeal afferent pathway in man. Am J Nephrol 20(1):42-47.

Labrie F, Giguere V, Proulx L, Lefevre G. 1984. Interactions between Crf, Epinephrine,

Vasopressin and Glucocorticoids in the Control of Acth-Secretion. Journal of Steroid

Biochemistry and Molecular Biology 20(1):153-160.

Landgraf R, Neumann ID. 2004. Vasopressin and oxytocin release within the brain: a dynamic

concept of multiple and variable modes of neuropeptide communication. Front

Neuroendocrinol 25(3-4):150-176.

Laragh JH, Angers M, Kelly WG, Lieberman S. 1960. Hypotensive agents and pressor

substances. The effect of epinephrine, norepinephrine, angiotensin II, and others on the

secretory rate of aldosterone in man. Jama 174(3):234-240.

Lee SK, Ryu PD, Lee SY. 2013. Differential distributions of neuropeptides in hypothalamic

paraventricular nucleus neurons projecting to the rostral ventrolateral medulla in the rat.

Neuroscience letters 556:160-165.

Page 157: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

157

Leong DS, Terron JA, Falcon-Neri A, Armando I, Ito T, Johren O, Tonelli LH, Hoe KL,

Saavedra JM. 2002. Restraint stress modulates brain, pituitary and adrenal expression of

angiotensin II AT(1A), AT(1B) and AT(2) receptors. Neuroendocrinology 75(4):227-

240.

Leshem M. 2011. Low dietary sodium is anxiogenic in rats. Physiology & Behavior 103(5):453-

458.

Li YF, Jackson KL, Stern JE, Rabeler B, Patel KP. 2006. Interaction between glutamate and

GABA systems in the integration of sympathetic outflow by the paraventricular nucleus

of the hypothalamus. Am J Physiol Heart Circ Physiol 291(6):H2847-2856.

Li ZH, Ferguson AV. 1993. Subfornical Organ Efferents to Paraventricular Nucleus Utilize

Angiotensin as a Neurotransmitter. American Journal of Physiology 265(2):R302-R309.

Lightman SL. 2008. The neuroendocrinology of stress: a never ending story. J Neuroendocrinol

20(6):880-884.

Lightman SL, Young WS. 1987. Vasopressin, Oxytocin, Dynorphin, Enkephalin and

Corticotropin-Releasing Factor Messenger-Rna Stimulation in the Rat. Journal of

Physiology-London 394:23-39.

Llewellyn T, Zheng H, Liu X, Xu B, Patel KP. 2012. Median preoptic nucleus and subfornical

organ drive renal sympathetic nerve activity via a glutamatergic mechanism within the

paraventricular nucleus. Am J Physiol Regul Integr Comp Physiol 302(4):R424-432.

Loewy AD, McKellar S. 1980. The neuroanatomical basis of central cardiovascular control. Fed

Proc 39(8):2495-2503.

Lucas M, Chen A, Richter-Levin G. 2013. Hypothalamic corticotropin-releasing factor is

centrally involved in learning under moderate stress. Neuropsychopharmacology

38(9):1825-1832.

Ludwig M, Callahan MF, Neumann I, Landgraf R, Morris M. 1994. Systemic Osmotic

Stimulation Increases Vasopressin and Oxytocin Release Within the Supraoptic Nucleus.

Journal of Neuroendocrinology 6(4):369-373.

Ludwig M, Leng G. 1997. Autoinhibition of supraoptic nucleus vasopressin neurons in vivo: a

combined retrodialysis/electrophysiological study in rats. The European journal of

neuroscience 9(12):2532-2540.

Ludwig M, Leng G. 2006. Dendritic peptide release and peptide-dependent behaviours. Nat Rev

Neurosci 7(2):126-136.

Page 158: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

158

Mabrouk OS, Kennedy RT. 2012. Simultaneous oxytocin and arg-vasopressin measurements in

microdialysates using capillary liquid chromatography-mass spectrometry. J Neurosci

Methods 209(1):127-133.

Macdonald K, Feifel D. 2014. Oxytocin’s role in anxiety: A critical appraisal. Brain Res

1580:22-56

Madden CJ, Morrison SF. 2009. Neurons in the paraventricular nucleus of the hypothalamus

inhibit sympathetic outflow to brown adipose tissue. Am J Physiol Regul Integr Comp

Physiol 296(3):R831-843.

Majzoub JA, Rich A, Vanboom J, Habener JF. 1983. Vasopressin and Oxytocin Messenger-Rna

Regulation in the Rat Assessed by Hybridization with Synthetic Oligonucleotides.

Journal of Biological Chemistry 258(23):4061-4064.

Mak P, Broussard C, Vacy K, Broadbear JH. 2012. Modulation of anxiety behavior in the

elevated plus maze using peptidic oxytocin and vasopressin receptor ligands in the rat.

Journal of psychopharmacology (Oxford, England) 26(4):532-542.

Makara GB, Kiss A, Lolait SJ, Aguilera G. 1996. Hypothalamic-pituitary corticotroph function

after shunting of magnocellular vasopressin and oxytocin to the hypophyseal portal

circulation. Endocrinology 137(2):580-586.

Martin R, Geis R, Holl R, Schafer M, Voigt KH. 1983. Co-Existence of Unrelated Peptides in

Oxytocin and Vasopressin Terminals of Rat Neurohypophyses - Immunoreactive

Methionines-Enkephalin-Like,Leucine5-Enkephalin-Like and Cholecystokinin-Like

Substances. Neuroscience 8(2):213-227.

Martin R, Voigt KH. 1981. Enkephalins co-exist with oxytocin and vasopressin in nerve

terminals of rat neurohypophysis. Nature 289(5797):502-504.

Masaharu N, Hiraki S. 2013. Central regulation of body-fluid homeostasis. Trends Neurosci

36(11):661-673.

McKinley MJ, Gerstberger R, Mathai ML, Oldfield BJ, Schmid H. 1999. The lamina terminalis

and its role in fluid and electrolyte homeostasis. J Clin Neurosci 6(4):289-301.

McKinley MJ, Mathai ML, McAllen RM, McClear RC, Miselis RR, Pennington GL, Vivas L,

Wade JD, Oldfield BJ. 2004. Vasopressin secretion: osmotic and hormonal regulation by

the lamina terminalis. J Neuroendocrinol 16(4):340-347.

McKinley MJ, McAllen RM, Davern P, Giles ME, Penschow J, Sunn N, Uschakov A, Oldfield

BJ. 2002. The sensory circumventricular organs of the mammalian brain: subfornical

organ, OVLT and area postrema. Berlin: Springer-Verlag.

Page 159: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

159

Mecawi Ade S, Ruginsk SG, Elias LL, Varanda WA, Antunes-Rodrigues J. 2015.

Neuroendocrine Regulation of Hydromineral Homeostasis. Compr Physiol 5(3):1465-

1516.

Melander T, Hokfelt T, Rokaeus A. 1986a. Distribution of galaninlike immunoreactivity in the

rat central nervous system. J Comp Neurol 248(4):475-517.

Melander T, Hokfelt T, Rokaeus A, Cuello AC, Oertel WH, Verhofstad A, Goldstein M. 1986b.

Coexistence of galanin-like immunoreactivity with catecholamines, 5-

hydroxytryptamine, GABA and neuropeptides in the rat CNS. J Neurosci 6(12):3640-

3654.

Meneely GR, Ball CO. 1958. Experimental epidemiology of chronic sodium chloride toxicity

and the protective effect of potassium chloride. The American journal of medicine

25(5):713-725.

Menendez JA, Mcgregor IS, Healey PA, Atrens DM, Leibowitz SF. 1990. Metabolic Effects of

Neuropeptide-Y Injections into the Paraventricular Nucleus of the Hypothalamus. Brain

research 516(1):8-14.

Meynen G, Unmehopa UA, van Heerikhuize JJ, Hofman MA, Swaab DF, Hoogendijk WJ. 2006.

Increased arginine vasopressin mRNA expression in the human hypothalamus in

depression: A preliminary report. Biol Psychiatry 60(8):892-895.

Milner TA, Reis DJ, Pickel VM, Aicher SA, Giuliano R. 1993. Ultrastructural localization and

afferent sources of corticotropin-releasing factor in the rat rostral ventrolateral medulla:

implications for central cardiovascular regulation. J Comp Neurol 333(2):151-167.

Miselis RR. 1981. The Efferent Projections of the Subfornical Organ of the Rat - a

Circumventricular Organ within a Neural Network Subserving Water-Balance. Brain

research 230(1-2):1-23.

Morita M, Kita Y, Notsu Y. 2001. Mechanism of AVP release and synthesis in chronic salt-

loaded rats. The Journal of pharmacy and pharmacology 53(12):1703-1709.

Murck H, Held K, Ziegenbein M, Kunzel H, Koch K, Steiger A. 2003a. The renin-angiotensin-

aldosterone system in patients with depression compared to controls--a sleep endocrine

study. BMC psychiatry 3:15.

Murck H, Schussler P, Steiger A. 2012. Renin-angiotensin-aldosterone system: the forgotten

stress hormone system: relationship to depression and sleep. Pharmacopsychiatry

45(3):83-95.

Neumann ID, Landgraf R. 2012b. Balance of brain oxytocin and vasopressin: implications for

anxiety, depression, and social behaviors. Trends in Neurosciences 35(11):649-659.

Page 160: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

160

Niebylski A, Boccolini A, Bensi N, Binotti S, Hansen C, Yaciuk R, Gauna H. 2012.

Neuroendocrine Changes and Natriuresis in Response to Social Stress in Rats. Stress and

Health 28(3):179-185.

O'Donaughy TL, Brooks VL. 2006. Deoxycorticosterone acetate-salt rats: hypertension and

sympathoexcitation driven by increased NaCl levels. Hypertension 47(4):680-685.

O'Donaughy TL, Qi Y, Brooks VL. 2006. Central action of increased osmolality to support

blood pressure in deoxycorticosterone acetate-salt rats. Hypertension 48(4):658-663.

Oka Y, Ye M, Zuker CS. 2015. Thirst driving and suppressing signals encoded by distinct neural

populations in the brain. Nature 520(7547):349-352.

Ozaki Y, Nomura M, Saito J, Luedke CE, Muglia LJ, Matsumoto T, Ogawa S, Ueta Y, Pfaff

DW. 2004. Expression of the arginine vasopressin gene in response to salt loading in

oxytocin gene knockout mice. J Neuroendocrinol 16(1):39-44.

Papanek PE, Sladek CD, Raff H. 1997. Corticosterone inhibition of osmotically stimulated

vasopressin from hypothalamic-neurohypophysial explants. Am J Physiol 272(1 Pt

2):R158-162.

Peter K. 1927. Studies on construction and development of the kidney. Jena: G. Fischer.

Peters S, Slattery DA, Uschold-Schmidt N, Reber SO, Neumann ID. 2014. Dose-dependent

effects of chronic central infusion of oxytocin on anxiety, oxytocin receptor binding and

stress-related parameters in mice. Psychoneuroendocrinology 42:225-236.

Phillips MI, Sumners C. 1998. Angiotensin II in central nervous system physiology. Regul Pept

78(1-3):1-11.

Pirnik Z, Mravec B, Kiss A. 2004. Fos protein expression in mouse hypothalamic paraventricular

(PVN) and supraoptic (SON) nuclei upon osmotic stimulus: colocalization with

vasopressin, oxytocin, and tyrosine hydroxylase. Neurochem Int 45(5):597-607.

Pleil KE, Rinker JA, Lowery-Gionta EG, Mazzone CM, McCall NM, Kendra AM, Olson DP,

Lowell BB, Grant KA, Thiele TE, Kash TL. 2015. NPY signaling inhibits extended

amygdala CRF neurons to suppress binge alcohol drinking. Nat Neurosci 18(4):545-552.

Polito AB, 3rd, Goldstein DL, Sanchez L, Cool DR, Morris M. 2006. Urinary oxytocin as a non-

invasive biomarker for neurohypophyseal hormone secretion. Peptides 27(11):2877-

2884.

Potapenko ES, Biancardi VC, Zhou Y, Stern JE. 2013. Astrocytes modulate a postsynaptic

NMDA-GABAA-receptor crosstalk in hypothalamic neurosecretory neurons. J Neurosci

33(2):631-640.

Page 161: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

161

Poulain DA, Wakerley JB. 1982. Electrophysiology of Hypothalamic Magnocellular Neurons

Secreting Oxytocin and Vasopressin. Neuroscience 7(4):773-808.

Pretel S, Piekut DT. 1989. Mediation of changes in paraventricular vasopressin and oxytocin

mRNA content to the medullary vagal complex and spinal cord of the rat. J Chem

Neuroanat 2(6):327-334.

Pyner S. 2009. Neurochemistry of the paraventricular nucleus of the hypothalamus: implications

for cardiovascular regulation. J Chem Neuroanat 38(3):197-208.

Raadsheer FC, Hoogendijk WJ, Stam FC, Tilders FJ, Swaab DF. 1994. Increased numbers of

corticotropin-releasing hormone expressing neurons in the hypothalamic paraventricular

nucleus of depressed patients. Neuroendocrinology 60(4):436-444.

Rasmussen MS, Simonsen JA, Sandgaard NC, Hoilund-Carlsen PF, Bie P. 2004. Effects of

oxytocin in normal man during low and high sodium diets. Acta Physiol Scand

181(2):247-257.

Rivalland ET, Tilbrook AJ, Turner AI, Iqbal J, Pompolo S, Clarke IJ. 2006. Projections to the

preoptic area from the paraventricular nucleus, arcuate nucleus and the bed nucleus of the

stria terminalis are unlikely to be involved in stress-induced suppression of GnRH

secretion in sheep. Neuroendocrinology 84(1):1-13.

Rivest S, Laflamme N, Nappi RE. 1995. Immune challenge and immobilization stress induce

transcription of the gene encoding the CRF receptor in selective nuclei of the rat

hypothalamus. J Neurosci 15(4):2680-2695.

Rivier C, Vale W. 1983. Modulation of stress-induced ACTH release by corticotropin-releasing

factor, catecholamines and vasopressin. Nature 305(5932):325-327.

Rivier J, Spiess J, Thorner M, Vale W. 1982. Characterization of a Growth Hormone-Releasing

Factor from a Human Pancreatic-Islet Tumor. Nature 300(5889):276-278.

Roberts EM, Pope GR, Newson MJ, Lolait SJ, O'Carroll AM. 2011. The vasopressin V1b

receptor modulates plasma corticosterone responses to dehydration-induced stress. J

Neuroendocrinol 23(1):12-19.

Robson AC, Leckie CM, Seckl JR, Holmes MC. 1998. 11 Beta-hydroxysteroid dehydrogenase

type 2 in the postnatal and adult rat brain. Brain Res Mol Brain Res 61(1-2):1-10.

Rodaros D, Caruana DA, Amir S, Stewart J. 2007. Corticotropin-releasing factor projections

from limbic forebrain and paraventricular nucleus of the hypothalamus to the region of

the ventral tegmental area. Neuroscience 150(1):8-13.

Page 162: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

162

Rokaeus A, Young WS, Mezey E. 1988. Galanin Coexists with Vasopressin in the Normal Rat

Hypothalamus and Galanins Synthesis Is Increased in the Brattleboro (Diabetes-

Insipidus) Rat. Neuroscience letters 90(1-2):45-50.

Rood BD, Stott RT, You S, Smith CJ, Woodbury ME, De Vries GJ. 2013. Site of origin of and

sex differences in the vasopressin innervation of the mouse (Mus musculus) brain. J

Comp Neurol 521(10):2321-2358.

Saavedra JM. 2012. Angiotensin II AT(1) receptor blockers ameliorate inflammatory stress: a

beneficial effect for the treatment of brain disorders. Cell Mol Neurobiol 32(5):667-681.

Saavedra JM, Pavel J. 2005. Angiotensin II AT1 receptor antagonists inhibit the angiotensin-

CRF-AVP axis and are potentially useful for the treatment of stress-related and mood

disorders. Drug Development Research 65(4):237-269.

Sampey DB, Burrell LM, Widdop RE. 1999. Vasopressin V2 receptor enhances gain of

baroreflex in conscious spontaneously hypertensive rats. Am J Physiol 276(3 Pt 2):R872-

879.

Sanders PW. 2009. Vascular consequences of dietary salt intake. Am J Physiol Renal Physiol

297(2):F237-243.

Sanders PW, Gibbs CL, Akhi KM, MacMillan-Crow LA, Zinn KR, Chen YF, Young CJ,

Thompson JA. 2001. Increased dietary salt accelerates chronic allograft nephropathy in

rats. Kidney international 59(3):1149-1157.

Sapirstein LA, Brandt WL, Drury DR. 1950. Production of arterial hypertension in the rat by

substitution of hypertonic sodium chloride solutions for drinking water. The American

journal of medicine 8(4):525-526.

Sawchenko PE. 1987. Evidence for differential regulation of corticotropin-releasing factor and

vasopressin immunoreactivities in parvocellular neurosecretory and autonomic-related

projections of the paraventricular nucleus. Brain Res 437(2):253-263.

Schally AV. 1978. Aspects of hypothalamic regulation of the pituitary gland. Science

202(4363):18-28.

Schally AV, Arimura A, Bowers CY, Kastin AJ, Sawano S, Reeding TW. 1968. Hypothalamic

neurohormones regulating anterior pituitary function. Recent Prog Horm Res 24:497-588.

Schally AV, Arimura A, Kastin AJ. 1973. Hypothalamic regulatory hormones. Science

179(4071):341-350.

Schweda F, Kurtz A. 2010. Regulation of renin release by local and systemic factors. Reviews of

physiology, biochemistry and pharmacology 161:1-44.

Page 163: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

163

Scott Young W. 1986. Corticotropin-releasing factor mRNA in the hypothalamus is affected

differently by drinking saline and by dehydration. FEBS Letters 208(1):158-162.

Seckl JR, Johnson MR, Lightman SL. 1989. Vasopressin and oxytocin responses to hypertonic

saline infusion: effect of the opioid antagonist naloxone. Clin Endocrinol (Oxf)

30(5):513-518.

Selye H. 1943. Production of nephrosclerosis in the fowl by sodium chloride. Jour Amer Vet

Med Assoc 103((798)):140-143.

Selye H, Hall CE, Rowley EM. 1943. Malignant hypertension produced by treatment with

desoxycorticosterone acetate and sodium chloride. Canadian Medical Association Journal

49:88-92.

Selye H, Stone H. 1943. Role of sodium chloride in production of nephrosclerosis by steroids.

Proceedings of the Society for Experimental Biology and Medicine 52(3):190-193.

Sharp FR, Sagar SM, Hicks K, Lowenstein D, Hisanaga K. 1991. C-Fos Messenger-Rna, Fos,

and Fos-Related Antigen Induction by Hypertonic Saline and Stress. Journal of

Neuroscience 11(8):2321-2331.

Shelat SG, Flanagan-Cato LM, Fluharty SJ. 1999. Glucocorticoid and mineralocorticoid

regulation of angiotensin II type 1 receptor binding and inositol triphosphate formation in

WB cells. J Endocrinol 162(3):381-391.

Sivukhina EV, Poskrebysheva AS, Smurova Iu V, Dolzhikov AA, Morozov Iu E, Jirikowski GF,

Grinevich V. 2009. Altered hypothalamic-pituitary-adrenal axis activity in patients with

chronic heart failure. Horm Metab Res 41(10):778-784.

Sjoquist M, Huang W, Jacobsson E, Skott O, Stricker EM, Sved AF. 1999. Sodium excretion and

renin secretion after continuous versus pulsatile infusion of oxytocin in rats.

Endocrinology 140(6):2814-2818.

Sofroniew MV. 1980. Projections from vasopressin, oxytocin, and neurophysin neurons to neural

targets in the rat and human. The journal of histochemistry and cytochemistry : official

journal of the Histochemistry Society 28(5):475-478.

Son SJ, Filosa JA, Potapenko ES, Biancardi VC, Zheng H, Patel KP, Tobin VA, Ludwig M,

Stern JE. 2013. Dendritic peptide release mediates interpopulation crosstalk between

neurosecretory and preautonomic networks. Neuron 78(6):1036-1049.

Sonino N, Fallo F, Fava GA. 2006. Psychological aspects of primary aldosteronism. Psychother

Psychosom 75(5):327-330.

Page 164: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

164

Sonino N, Tomba E, Genesia ML, Bertello C, Mulatero P, Veglio F, Fava GA, Fallo F. 2011.

Psychological assessment of primary aldosteronism: a controlled study. J Clin Endocrinol

Metab 96(6):E878-883.

Spencer S, Saper CB, Joh T, Reis DJ, Goldstein M, Raese JD. 1985. Distribution of

catecholamine-containing neurons in the normal human hypothalamus. Brain Res

328(1):73-80.

St-Louis R, Parmentier C, Raison D, Grange-Messent V, Hardin-Pouzet H. 2012. Reactive

oxygen species are required for the hypothalamic osmoregulatory response.

Endocrinology 153(3):1317-1329.

Stachniak TJ, Trudel E, Bourque CW. 2014. Cell-specific retrograde signals mediate antiparallel

effects of angiotensin II on osmoreceptor afferents to vasopressin and oxytocin neurons.

Cell Rep 8(2):355-362.

Stellar E, Hyman R, Samet S. 1954. Gastric Factors Controlling Water-Solution-Drinking and

Salt-Solution-Drinking. Journal of Comparative and Physiological Psychology 47(3):220-

226.

Stern JE. 2015. Neuroendocrine-autonomic integration in the paraventricular nucleus: novel roles

for dendritically released neuropeptides. J Neuroendocrinol 27(6):487-497.

Stocker SD, Cunningham JT, Toney GM. 2004. Water deprivation increases Fos

immunoreactivity in PVN autonomic neurons with projections to the spinal cord and

rostral ventrolateral medulla. Am J Physiol Regul Integr Comp Physiol 287(5):R1172-

1183.

Stocker SD, Hunwick KJ, Toney GM. 2005. Hypothalamic paraventricular nucleus differentially

supports lumbar and renal sympathetic outflow in water-deprived rats. Journal of

Physiology-London 563(1):249-263.

Stocker SD, Osborn JL, Carmichael SP. 2008. Forebrain osmotic regulation of the sympathetic

nervous system. Clin Exp Pharmacol Physiol 35(5-6):695-700.

Stocker SD, Simmons JR, Stornetta RL, Toney GM, Guyenet PG. 2006. Water deprivation

activates a glutamatergic projection from the hypothalamic paraventricular nucleus to the

rostral ventrolateral medulla. J Comp Neurol 494(4):673-685.

Stocker SD, Sved AF, Stricker EM. 2000. Role of renin-angiotensin system in hypotension-

evoked thirst: studies with hydralazine. Am J Physiol Regul Integr Comp Physiol

279(2):R576-585.

Strazzullo P, D'Elia L, Kandala NB, Cappuccio FP. 2009. Salt intake, stroke, and cardiovascular

disease: meta-analysis of prospective studies. BMJ 339:b4567.

Page 165: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

165

Stricker EM, Sved AF. 2000. Thirst. Nutrition 16(10):821-826.

Stricker EM, Verbalis JG. 1986. Interaction of osmotic and volume stimuli in regulation of

neurohypophyseal secretion in rats. Am J Physiol 250(2 Pt 2):R267-275.

Sullivan RC, Fay WH, Schuler AL, Campbell M, Coleman M, Caparulo B, Cohen DJ. 1980.

Why Do Autistic-Children. J Autism Dev Disord 10(2):231-241.

Sun N, Roberts L, Cassell MD. 1991. Rat Central Amygdaloid Nucleus Projections to the Bed

Nucleus of Thestria Terminalis. Brain research bulletin 27(5):651-662.

Sunn N, McKinley MJ, Oldfield BJ. 2003. Circulating angiotensin II activates neurones in

circumventricular organs of the lamina terminalis that project to the bed nucleus of the

stria terminalis. Journal of Neuroendocrinology 15(8):725-731.

Swanson LW, Kuypers HG. 1980. The paraventricular nucleus of the hypothalamus:

cytoarchitectonic subdivisions and organization of projections to the pituitary, dorsal

vagal complex, and spinal cord as demonstrated by retrograde fluorescence double-

labeling methods. J Comp Neurol 194(3):555-570.

Swanson LW, Sawchenko PE. 1980. Paraventricular nucleus: a site for the integration of

neuroendocrine and autonomic mechanisms. Neuroendocrinology 31(6):410-417.

Swanson LW, Sawchenko PE. 1983. Hypothalamic integration: organization of the

paraventricular and supraoptic nuclei. Annu Rev Neurosci 6:269-324.

Swanson LW, Sawchenko PE, Berod A, Hartman BK, Helle KB, Vanorden DE. 1981. An

immunohistochemical study of the organization of catecholaminergic cells and terminal

fields in the paraventricular and supraoptic nuclei of the hypothalamus. J Comp Neurol

196(2):271-285.

Swanson LW, Simmons DM. 1989. Differential steroid hormone and neural influences on

peptide mRNA levels in CRH cells of the paraventricular nucleus: a hybridization

histochemical study in the rat. J Comp Neurol 285(4):413-435.

Tang Y, McKenna KE. 2001. Paraventricular nucleus projections (PVN) to spinal and brainstem

nuclei involved in the control of sexual function. Society for Neuroscience Abstracts

27(2):2227-2227.

Taniguchi H, He M, Wu P, Kim S, Paik R, Sugino K, Kvitsiani D, Fu Y, Lu J, Lin Y, Miyoshi

G, Shima Y, Fishell G, Nelson SB, Huang ZJ. 2011. A resource of Cre driver lines for

genetic targeting of GABAergic neurons in cerebral cortex. Neuron 71(6):995-1013.

Page 166: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

166

Taurines R, Schwenck C, Lyttwin B, Schecklmann M, Jans T, Reefschlager L, Geissler J,

Gerlach M, Romanos M. 2014. Oxytocin plasma concentrations in children and

adolescents with autism spectrum disorder: correlation with autistic symptomatology.

Atten Defic Hyperact Disord 6(3):231-239.

Teitelbaum P, Stellar E. 1954. Recovery from the failure to eat produced by hypothalamic

lesions. Science 120(3126):894-895.

Tilbrook AJ, Rivalland ETA, Turner AI, Pompolo S, Clarke IJ. 2006. Isolation/restraint stress in

sheep activates projections to the preoptic area arising from the paraventricular nucleus

but not those arising from the bed nucleus of the stria terminalis or the arcuate nucleus.

Frontiers in neuroendocrinology 27(1):40-41.

Tobian L, Hanlon S. 1990. High Sodium-Chloride Diets Injure Arteries and Raise Mortality

without Changing Blood-Pressure. Hypertension 15(6):900-903.

Tsuruo Y, Ceccatelli S, Villar MJ, Hokfelt T, Visser TJ, Terenius L, Goldstein M, Brown JC,

Buchan A, Walsh J. 1988. Coexistence of TRH with other neuroactive substances in the

rat central nervous system. J Chem Neuroanat 1(5):235-253.

Turkkan JS. 1994. Biobehavioral effects of extended salt loading and conflict stress in intact

baboons. J Exp Anal Behav 61(2):263-272.

Ulrich-Lai YM, Herman JP. 2009. Neural regulation of endocrine and autonomic stress

responses. Nat Rev Neurosci 10(6):397-409.

Vale W, Rivier C, Brown M. 1977. Regulatory peptides of the hypothalamus. Annu Rev Physiol

39:473-527.

Vale W, Rivier J, Vaughan J, McClintock R, Corrigan A, Woo W, Karr D, Spiess J. 1986.

Purification and characterization of an FSH releasing protein from porcine ovarian

follicular fluid. Nature 321(6072):776-779.

Vale W, Spiess J, Rivier C, Rivier J. 1981. Characterization of a 41-Residue Ovine

Hypothalamic Peptide That Stimulates Secretion of Corticotropin and Beta-Endorphin.

Science 213(4514):1394-1397.

Vale W, Vaughan J, Smith M, Yamamoto G, Rivier J, Rivier C. 1983. Effects of Synthetic Ovine

Corticotropin-Releasing Factor, Glucocorticoids, Catecholamines, Neurohypophyseal

Peptides, and Other Substances on Cultured Corticotropic Cells. Endocrinology

113(3):1121-1131.

Valentino RJ, Pavcovich LA, Hirata H. 1995. Evidence for corticotropin-releasing hormone

projections from Barrington's nucleus to the periaqueductal gray and dorsal motor

nucleus of the vagus in the rat. J Comp Neurol 363(3):402-422.

Page 167: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

167

Vanderhaeghen JJ, Lotstra F, Liston DR, Rossier J. 1983. Proenkephalin, [Met]enkephalin, and

oxytocin immunoreactivities are colocalized in bovine hypothalamic magnocellular

neurons. Proceedings of the National Academy of Sciences of the United States of

America 80(16):5139-5143.

Vanderhaeghen JJ, Lotstra F, Vandesande F, Dierickx K. 1981. Coexistence of Cholecystokinin

and Oxytocin-Neurophysin in Some Magnocellular Hypothalamo-Hypophyseal Neurons.

Cell and Tissue Research 221(1):227-231.

Vandesande F, Dierickx K. 1975. Identification of Vasopressin Producing and of Oxytocin

Producing Neurons in Hypothalamic Magnocellular Neurosecretory-System of Rat. Cell

and Tissue Research 164(2):153-162.

Varudkar A, Ramakrishnan U. 2015. Commensalism facilitates gene flow in mountains: a

comparison between two Rattus species. Heredity (Edinb) 115(3):253-261.

Verbalis JG, Mangione MP, Stricker EM. 1991. Oxytocin Produces Natriuresis in Rats at

Physiological Plasma-Concentrations. Endocrinology 128(3):1317-1322.

Vogelzangs N, Seldenrijk A, Beekman AT, van Hout HP, de Jonge P, Penninx BW. 2010.

Cardiovascular disease in persons with depressive and anxiety disorders. Journal of

affective disorders 125(1-3):241-248.

Voisin DL, Bourque CW. 2002. Integration of sodium and osmosensory signals in vasopressin

neurons. Trends Neurosci 25(4):199-205.

Vujovic N, Gooley JJ, Jhou TC, Saper CB. 2015. Projections from the subparaventricular zone

define four channels of output from the circadian timing system. J Comp Neurol

523(18):2714-2737.

Wamsteeker Cusulin JI, Fuzesi T, Watts AG, Bains JS. 2013. Characterization of corticotropin-

releasing hormone neurons in the paraventricular nucleus of the hypothalamus of Crh-

IRES-Cre mutant mice. Plos One 8(5):e64943.

Wang J, Irnaten M, Venkatesan P, Evans C, Mendelowitz D. 2002. Arginine vasopressin

enhances GABAergic inhibition of cardiac parasympathetic neurons in the nucleus

ambiguus. Neuroscience 111(3):699-705.

Wang J, Qin W, Liu B, Zhou Y, Wang D, Zhang Y, Jiang T, Yu C. 2013a. Neural mechanisms

of oxytocin receptor gene mediating anxiety-related temperament. Brain Struct Funct

219(5):1543-1554.

Wang YF, Sun MY, Hou Q, Hamilton KA. 2013b. GABAergic inhibition through synergistic

astrocytic neuronal interaction transiently decreases vasopressin neuronal activity during

hypoosmotic challenge. The European journal of neuroscience 37(8):1260-1269.

Page 168: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

168

Watson SJ, Akil H, Fischli W, Goldstein A, Zimmerman E, Nilaver G, van wimersma Griedanus

TB. 1982. Dynorphin and vasopressin: common localization in magnocellular neurons.

Science 216(4541):85-87.

Watts AG. 1992a. Disturbance of Fluid Homeostasis Leads to Temporally and Anatomically

Distinct Responses in Neuropeptide and Tyrosine-Hydroxylase Messenger-Rna Levels in

the Paraventricular and Supraoptic Nuclei of the Rat. Neuroscience 46(4):859-879.

Watts AG. 1992b. Osmotic Stimulation Differentially Affects Cellular-Levels of Corticotropin-

Releasing Hormone and Neurotensin Neuromedin-N Messenger-Rnas in the Lateral

Hypothalamic Area and Central Nucleus of the Amygdala. Brain research 581(2):208-

216.

Watts AG. 1996. The impact of physiological stimuli on the expression of corticotropin-releasing

hormone (CRH) and other neuropeptide genes. Front Neuroendocrinol 17(3):281-326.

Watts AG. 1999. Dehydration-associated anorexia: Development and rapid reversal. Physiology

& Behavior 65(4-5):871-878.

Watts AG, Kelly AB, Sanchez-Watts G. 1995. Neuropeptides and thirst: the temporal response

of corticotropin-releasing hormone and neurotensin/neuromedin N gene expression in rat

limbic forebrain neurons to drinking hypertonic saline. Behav Neurosci 109(6):1146-

1157.

Watts AG, Sanchezwatts G. 1995. A Cell-Specific Role for the Adrenal-Gland in Regulating Crh

Messenger-Rna Levels in Rat Hypothalamic Neurosecretory Neurons after Cellular

Dehydration. Brain Research 687(1-2):63-70.

Watts AG, Swanson LW, Sanchez-Watts G. 1987. Efferent projections of the suprachiasmatic

nucleus: I. Studies using anterograde transport of Phaseolus vulgaris leucoagglutinin in

the rat. J Comp Neurol 258(2):204-229.

Whitnall MH, Gainer H, Cox BM, Molineaux CJ. 1983. Dynorphin-a-(1-8) Is Contained within

Vasopressin Neurosecretory Vesicles in Rat Pituitary. Science 222(4628):1137-1139.

Wiedenmayer CP, Magarinos AM, McEwen BS, Barr GA. 2005. Age-specific threats induce

CRF expression in the paraventricular nucleus of the hypothalamus and hippocampus of

young rats. Horm Behav 47(2):139-150.

Williams RW, Rakic P. 1988. Three-dimensional counting: an accurate and direct method to

estimate numbers of cells in sectioned material. J Comp Neurol 278(3):344-352.

Williams TD, Abel DC, King CM, Jelley RY, Lightman SL. 1985. Vasopressin and oxytocin

responses to acute and chronic osmotic stimuli in man. The Journal of endocrinology

108(1):163-168.

Page 169: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

169

Windle RJ, Kershaw YM, Shanks N, Wood SA, Lightman SL, Ingram CD. 2004. Oxytocin

attenuates stress-induced c-fos mRNA expression in specific forebrain regions associated

with modulation of hypothalamo-pituitary-adrenal activity. J Neurosci 24(12):2974-2982.

Windle RJ, Shanks N, Lightman SL, Ingram CD. 1997. Central oxytocin administration reduces

stress-induced corticosterone release and anxiety behavior in rats. Endocrinology

138(7):2829-2834.

Wittmann G, Fuzesi T, Singru PS, Liposits Z, Lechan RM, Fekete C. 2009. Efferent projections

of thyrotropin-releasing hormone-synthesizing neurons residing in the anterior

parvocellular subdivision of the hypothalamic paraventricular nucleus. J Comp Neurol

515(3):313-330.

Yang G, Xi ZX, Wan Y, Wang H, Bi G. 1993. Changes in circulating and tissue angiotensin II

during acute and chronic stress. Biological signals 2(3):166-172.

Yang Z, Coote JH. 1998. Influence of the hypothalamic paraventricular nucleus on

cardiovascular neurones in the rostral ventrolateral medulla of the rat. The Journal of

physiology 513 ( Pt 2)(2):521-530.

Yao ST, Antunes VR, Bradley PM, Kasparov S, Ueta Y, Paton JF, Murphy D. 2011. Temporal

profile of arginine vasopressin release from the neurohypophysis in response to

hypertonic saline and hypotension measured using a fluorescent fusion protein. J

Neurosci Methods 201(1):191-195.

Yu HCM, Burrell LM, Black MJ, Wu LL, Dilley RJ, Cooper ME, Johnston CI. 1998. Salt

induces myocardial and renal fibrosis in normotensive and hypertensive rats. Circulation

98(23):2621-2628.

Yue C, Mutsuga N, Sugimura Y, Verbalis J, Gainer H. 2008. Differential kinetics of oxytocin

and vasopressin heteronuclear RNA expression in the rat supraoptic nucleus in response

to chronic salt loading in vivo. J Neuroendocrinol 20(2):227-232.

Zhai XY, Thomsen JS, Birn H, Kristoffersen IB, Andreasen A, Christensen EI. 2006. Three-

dimensional reconstruction of the mouse nephron. Journal of the American Society of

Nephrology : JASN 17(1):77-88.

Ziegler DR, Herman JP. 2002. Neurocircuitry of stress integration: anatomical pathways

regulating the hypothalamo-pituitary-adrenocortical axis of the rat. Integr Comp Biol

42(3):541-551.

Page 170: © 2016 Justin Andrew Smithufdcimages.uflib.ufl.edu/UF/E0/05/01/86/00001/SMITH_J.pdf · 1-1 Schematic of the major systems that interact with neurons in the paraventricular ... responses

170

BIOGRAPHICAL SKETCH

Justin entered academia later in life, attending Clark State Community College in

Springfield, OH in January 2008 at the age of 32. After one year of coursework, he transferred

to the University of Cincinnati where he completed premedical coursework and explored

professions related to medicine and basic science research. In December of 2010, he accepted a

position working in the laboratory of Dr. Randall Sakai under the mentorship of Dr. Eric Krause.

This experience exposed him to the basics of neuroanatomical methods and research aims

exploring the central actions of neuropeptides and rodent models of behavior. Graduating with a

BS in neuroscience and a minor in mathematics in March of 2012, Justin then moved to the

University of Florida to work with Dr. Krause before beginning graduate studies in August of the

same year. During the course of his training, Justin was involved with several research projects

producing eight published peer-reviewed manuscripts with four more manuscripts either

submitted for review or in preparation at the time of this publication. Following graduation,

Justin plans to move to Boston College and Michigan State University to continue his training as

a postdoctoral fellow under the mentorship of Dr. Alexa Veenema.