Neural Basis for Regulation of Secretion by Presystemic Anticipated Disturbances in Osmolality, and by Hypoglycemia

The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters

Citation Kim, Angela. 2020. Neural Basis for Regulation of Vasopressin Secretion by Presystemic Anticipated Disturbances in Osmolality, and by Hypoglycemia. Doctoral dissertation, Harvard University, Graduate School of Arts & Sciences.

Citable link https://nrs.harvard.edu/URN-3:HUL.INSTREPOS:37365939

Terms of Use This article was downloaded from Harvard University’s DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http:// nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA

Neural Basis for Regulation of Vasopressin Secretion by Presystemic Anticipated Disturbances

in Osmolality, and by Hypoglycemia

A dissertation presented

by

Angela Kim

to

The Division of Medical Sciences

in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

in the subject of

Neuroscience

Harvard University

Cambridge, Massachusetts

April 2020

© 2020 Angela Kim

All rights reserved.

Dissertation Advisor: Bradford B. Lowell Angela Kim

Neural Basis for Regulation of Vasopressin Secretion by Presystemic Anticipated

Disturbances in Osmolality, and by Hypoglycemia

Abstract

Vasopressin (AVP), released by magnocellular AVP neurons of the hypothalamus, is a multifaceted hormone. AVP regulates many physiological processes including blood osmolality, blood glucose, and blood pressure. In this dissertation, I present two independent studies investigating the neural mechanism for role of AVP in fluid and glucose homeostasis.

In regards to regulating water excretion and hence blood osmolality, magnocellular AVP neurons are regulated by two temporally distinct signals: 1) slow systemic signals that convey systemic osmolality information, and 2) rapid ‘presystemic’ signals that anticipate future osmotic challenges. Magnocellular AVP neurons show bidirectional presystemic responses to feeding and drinking. In our first study, we identified and characterized the neural circuits mediating presystemic regulation of AVP release. Using in vivo calcium imaging and opto- and chemo- genetic approaches, we demonstrated that presystemic regulation of magnocellular AVP neurons is mediated by two functionally-distinct neural circuits that transmit information related ingestion of either water or food, which decrease or increase osmolality, respectively. We also validated a brainstem circuit mediating blood pressure information to magnocellular AVP neurons that is completely separate from water- and food-related presystemic circuits.

iii

Next, we focused on the glucoregulatory function of AVP. We adopted a wide range of in vitro and in vivo techniques in mice and human to provide a comprehensive analysis of central and peripheral pathways mediating AVP’s glucoregulatory effect. We performed the very first in vivo simultaneous real-time monitoring of blood glucose and magnocellular AVP neuron activity in mice to demonstrate robust activation of magnocellular AVP neurons by hypoglycemia. We further demonstrated that A1/C1 neurons in the brainstem mediate this activation. We showed that AVP acts on V1b subtype of AVP receptors specifically expressed on pancreatic alpha cells to stimulate glucagon release, which then stimulates hepatic glucose production.

Taken together, our data shows that activity of magnocellular AVP neurons is tightly regulated by multiple functionally-distinct neural circuits. We propose that proper regulation of

AVP release is crucial for homeostasis of multiple physiological processes, and dysregulation of

AVP release might underlie certain diseases with impaired water and glucose balance.

iv

Table of Contents

Title Page………………………………………………………………………………………….i Copyright Page……………………………………………………….…………………………..ii Abstract……………………………………………...... ….iii Table of Contents………………………………...... ……..….v List of Figures…………………………………………………………………...... viii Acknowledgement……………………………………………………………...... x Chapter 1. Introduction 1.1 Introduction to Vasopressin...... 2 1.2 Systemic regulation of vasopressin release……………………...... 4 1.3 New insight into hypothalamic function…...... …....6 1.4 Presystemic regulation of Vasopressin neurons...... ……8 1.5 Osmotic homeostasis and significance of presystemic regulation...... 12 References…...... 16

Chapter 2. Regulation of Vasopressin Secretion by Presystemic Anticipated Disturbances in Osmolality 2.1 Introduction…...... 22 2.2 Results 2.2.1 Vasopressin neurons receive inhibitory and excitatory inputs from the LT...... 24 2.2.2 Water-related presystemic regulation of vasopressin neurons is mediated by the LT...... ……27 2.2.3 Presystemic signals enter the LT at the level of MnPO...... 31 2.2.4 SON-projecting LT neurons show presystemic response to water-predicting cues and drinking...... 34 2.2.5 The LT does not mediate food-related presystemic regulation of vasopressin Neurons...... 37 2.2.6 Vasopressin neurons receive excitatory inputs from A1/C1 neurons in the VLM…..40

v

2.2.7 A1/C1 neurons in the VLM mediate hypotension-induced, but not feeding-induced, activation of vasopressin neurons...... 43 2.2.8 Perinuclear zone GABAergic neurons do not mediate food-related presystemic regulation of vasopressin neurons...... 45 2.2.9 AgRP and POMC neurons do not mediate food-related presystemic regulation of vasopressin neurons...... 48 2.2.10 Food-related presystemic regulation of vasopressin neurons is mediated by unknown neurons in the ARC...... 51 2.3 Discussion…...... 54 2.4 Methods...... 55 References...... 62

Chapter 3. Regulation of Vasopressin Secretion by Hypoglycemia 3.1 Introduction...... 66 3.2 Results 3.2.1 Glucagon secretion in response to an insulin tolerance test is driven by Vasopressin…...... 68 3.2.2 Vasopressin evokes hyperglycemia and hyperglucagonemia...... 75 3.2.3 Hypoglycemia evokes Vasopressin secretion via activation of A1/C1 neurons...... 79 3.2.4 Vasopressin -induced glucagon secretion maintains glucose homeostasis during dehydration…...... 83 3.2.5 Insulin-induced Vasopressin secretion may underlie counter-regulatory glucagon secretion in human, and is diminished in subjects with Type 1 Diabetes...... 85 3.3 Discussion...... 88 3.4 Methods...... 90 References...... 109

Chapter 4. Conclusion 4.1 Summary of findings...... 116 4.2 Function of presystemic regulation in Vasopressin neurons...... 118 4.3 Significance of functionally distinct circuits in regulation of Vasopressin release...... 120 4.4 Neural mechanism for stress-induced Vasopressin release...... 122

vi

4.5 Lack of food-predicting cue response in magnocellular Vasopressin neurons...... 123 4.6 Technical consideration: rabies retrograde tracing...... 125 4.7 ARC population mediating food-related presystemic regulation...... 128 4.8 Neural mechanism for hypoglycemia-induced Vasopressin release...... 129 4.9 Relevance of findings to diabetes and diabetes complications...... 131 References...... 133

Appendix A: Supplementary Figures for Chapter 2...... 139 Appendix B: Supplementary Figures for Chapter 3...... 144

vii

List of Figures Chapter 1 Figure 1.1. Location of AVP neurons Figure 1.2. Axon endings of magnocellular AVP neurons in the posterior pituitary Figure 1.3. Systemic regulation of AVP release Figure 1.4. Regulation of AVP release: Systemic and Presystemic regulation Figure 1.5. Changes in plasma osmolality after water and food intake

Chapter 2 Figure 2.1. Magnocellular AVP neurons receive excitatory and inhibitory input from the LT Figure 2.2. The LT mediates water-related presystemic regulation of SONAVP neurons Figure 2.3. Organization of water-related presystemic neural circuit upstream and downstream of the LT Figure 2.4. SON-projecting SFOVglut2, MnPO/OVLTVglut2, and MnPO/OVLTVgat neurons show presystemic responses to water bowl placement and drinking Figure 2.5. The MnPO/OVLT is not involved in food-related presystemic regulation of SONAVP neurons Figure 2.6. SONAVP neurons receive direct glutamatergic inputs from A1/C1 neurons of the VLM Figure 2.7. A1/C1 neurons mediate hypotension-induced activation of SONAVP neurons, but are not required for presystemic regulation Figure 2.8. PNZGABA neurons show presystemic responses to feeding but are not required for food-related presystemic regulation of SONAVP neurons Figure 2.9. non-AgRP/POMC neurons in the ARC mediates food-related presystemic regulation of SONAVP neurons

Chapter 3 Figure 3.1. Insulin-induced hypoglycaemia enhances population activity of AVP neurons in the SON, driving glucagon secretion Figure 3.2. Simultaneous continuous glucose monitoring (CGM) and in vivo fibre photometry of AVP neurons Figure 3.3. The vasopressin 1b receptor mediates insulin-induced glucagon secretion Figure 3.4. Metabolic response to elevating AVP in vivo Figure 3.5. Insulin-induced AVP secretion is mediated by A1/C1 neurons Figure 3.6. AVP-induced glucagon secretion during dehydration Figure 3.7. Insulin-induced hypoglycaemia evokes copeptin and glucagon secretion in human subjects Figure 3.8. Insulin-induced copeptin and glucagon secretion is diminished in type 1 diabetic subjects

Chapter 4 Figure 4.1. Regulation of AVP release

Appendix A Supplementary Figure 2.1. Lack of water- and food-induced responses in GFP-expressing AVP- IRES-Cre mice

viii

Supplementary Figure 2.2. Monosynaptic rabies tracing from SON-projecting SFONos-1 neurons showing lack of extra LT afferents Supplementary Figure 2.3. A1/C1 neurons do not mediate non blood pressure-related, aversive stimuli-induced activation of SONAVP neurons Supplementary Figure 2.4. PVHAVP and SONAVP neurons do not receive direct synaptic inputs from AgRP and POMC neurons

Appendix B Supplementary Figure 3.1. alpha-cell output is not increased by systemic concentrations of cortisol or adrenaline Supplementary Figure 3.2. AVP mediates hyperglycemia and hyperglucagonemia in response to 2DG Supplementary Figure 3.3. Effects of CNO in animals expressing mCherry in SON AVP neurons Supplementary Figure 3.4. Expression of the vasopressin 1b receptor in mouse Supplementary Figure 3.5. AVP increases glucagon secretion ex vivo and in situ Supplementary Figure 3.6. AVP evokes glucagon secretion via the canonical Gq pathway Supplementary Figure 3.7. Viral tracing of A1/C1 terminals Supplementary Figure 3.8. c-fos expression in A1/C1 neurons during an ITT Supplementary Figure 3.9. Raw glucagon and copeptin data from human studies

ix

Acknowledgement

I wish to express my deepest gratitude to all the people who provided invaluable assistance during my study.

First and foremost, I would like to thank my advisor, Dr. Brad Lowell, for his guidance

and encouragement over the past six years. I thank him for having confidence in me and giving

me an opportunity to be his first graduate student. He taught me all the essential aspects of being

a scientist. He is a ‘true scientist’ in my mind, a role model that I will continuously strive for.

I also wish to show my gratitude to my past advisor, Dr. Hee-Sup Shin, who allowed me

to embark on my journey as a scientist. I admire his passion and dedication for science. He

taught me to be hard-working, creative, and ambitious. I would not have achieved this far

without his presence in my life.

I want to extend my thanks to current and former members of the Lowell lab. Thank you

for your emotional and intellectual support throughout my study. Thank you for creating an

exciting and motivating environment to work in. I would not have survived miserable and

frustrating moments of my graduate study without you around.

Next, I would like to express my gratitude to my Dissertation Advisory Committee, Drs.

Clifford Saper, Joseph Majzoub, and Stephen Liberles, for their time and scientific guidance over

x the years. I also want to thank my Dissertation Exam Committee, Drs. Naoshige Uchida, Stephen

Liberles, Matt Carter, and Dong Kong, for kindly agreeing to participate.

I am also extremely grateful to my collaborator, Dr. Linford Briant, for giving me an opportunity to delve into a completely new field of science. I have thoroughly enjoyed working with you and look forward to continuing our scientific relationship.

Finally, I want to thank my parents, my sister, and my beloved dog, Tori, for their unconditional love and support. Thank you from the bottom of my heart.

xi

Chapter 1

Introduction

1

1.1 Introduction to Vasopressin

Vasopressin (AVP), also PVH known as antidiuretic hormone, is a nine-amino acid peptide hormone synthesized by AVP SON neurons in the hypothalamic paraventricular (PVH) and Figure 1.1. Location of AVP neurons. supraoptic (SON) nuclei of the AVP neurons are visualized by immunostaining for AVP brain (Figure 1.1) (Bolignano et (green). Blue: DAPI, Scale bar: 500µm. al., 2014; Brown et al., 2013).

AVP neurons are divided into two subpopulations based on their projections. Parvocellular AVP

neurons are located in the PVH and send projections to the median eminence or other regions in

the brain. Magnocellular AVP neurons are located in the PVH and SON and send projections to

the posterior pituitary. Magnocellular AVP neurons are the sole source of circulating AVP.

Axons of magnocellular AVP neurons are

equipped with 2000-10000 neurosecretary

endings packed with secretory granules

containing AVP (Figure 1.2) (Brown et al., 2013;

Leng et al., 1999). Upon activation of Figure 1.2. Axon endings of magnocellular AVP neurons in the posterior pituitary. magnocellular AVP neurons, secretory granules

Green: axon endings in the posterior pituitary are exocytosed to release AVP into the (center). Blue: DAPI, Scale bar: 500µm. extracellular space. Released AVP then diffuses

2

into fenestrated capillaries of the pituitary and enters the circulation. Magnocellular AVP neurons are also capable of releasing AVP from their dendrites, known as dendritic release.

Locally-released AVP serves autoregulatory function to maintain phasic firing of magnocellular

AVP neurons (Ludwig and Leng, 2006).

AVP is generated from its precursor, prepro-vasopressin, synthesized in the cell bodies

of magnocellular AVP neurons. Prepro-vasopressin is further processed into three-domain prohormone by a cleavage of signal peptide and transported in secretory granules to the axonal terminals located at the posterior pituitary (Ferguson et al., 2003). Secretory granules contain several enzymes, namely, endopeptidase, exopeptidase, mono-oxygenase, and lyase, which act in sequence to produce AVP, neurophysin II and copeptin (Acher, 1993). The whole synthesis process takes approximately 2 hours (Danziger and Zeidel, 2014). Once released into the circulation, AVP is immediately metabolized by vasopressinases in the liver and kidney, resulting in a short half-life of 5-12 minutes (Fenske et al., 2018; Janaky et al., 1982).

AVP is a versatile hormone involved in homeostatic maintenance of diverse

physiological processes. Circulating AVP, originating from magnocellular AVP neurons,

regulates water excretion and thus blood osmolality, blood pressure, and blood glucose. AVP

released in the median eminence by parvocellular AVP neurons aids in stress hormone release by

causing release from the pituitary of ACTH which then causes the adrenal gland to secrete

glucocorticoids. Recent evidences focusing on the effect centrally released AVP, mainly by

parvocellular AVP neurons, suggest that AVP is also involved in diverse social behaviors,

including pair bonding, parental behavior, and aggression (Caldwell, 2017). Known stimuli for

3

AVP release by the magnocellular AVP neurons include hypoosmolality, hypovolemia,

hypotension, hypoglycemia, stress and nausea.

Osmoregulatory function of AVP is mainly attributed to its action on the kidney

(Danziger and Zeidel, 2014; Koshimizu et al., 2012). AVP binds V2 subtype receptors in the

kidney and limits renal water excretion by altering water reabsorption properties of distal tubule

and collecting duct. When released in response to hypotension, AVP maintains blood pressure by

causing vasoconstriction. Binding of AVP to V1a subtype receptors in the vasculature causes

constriction of vascular smooth muscles, leading to an increase in blood pressure (Japundzic-

Zigon et al., 2020). The mechanism of AVP’s glucoregulatory function is less well-understood.

However, our recent study provides clear evidence for involvement of V1b subtype receptors in

the pancreas (Chapter 3). V1b receptors are specifically expressed in the pancreatic alpha cells

and binding of AVP causes release of glucagon. Elevated glucagon stimulates hepatic glucose production, leading to an increase in blood glucose. AVP is also involved in ACTH release in response to stress. AVP is released into the median eminence by parvocellular AVP neurons upon stress. AVP will then travel through the hypophyseal portal system to the anterior pituitary and act synergistically with corticotrophin-releasing hormone (CRH) to stimulate corticotrophs through the V1b receptor (Caldwell et al., 2017). Behavioral effects of centrally-released AVP is

thought to be mediated by its action on V1a and V1b subtypes expressed throughout the brain

(Caldwell, 2017).

1.2 Systemic regulation of vasopressin release

4

AVP release is tightly regulated by changes in systemic osmolality. Systemic regulation

of AVP release is mediated by three structures in the lamina terminalis (LT), namely, the subfornical organ (SFO), median preoptic nucleus (MnPO), and organum vasculosum lamina terminalis (OVLT). The SFO and OVLT are thought to function as central osmoreceptors

(McKinley et al., 1999). Location of central osmoreceptors was first suggested by a lesion study in goats, in which the authors found that a lesion involving anterior wall of the third ventricle caused a significant attenuation in thirst and AVP response to systemic hyperosmolality

(Andersson et al., 1975). The result was replicated in many studies using intracerebroventricular injections, electrophysiological recordings, c-fos expression, and functional MRI, identifying two regions as central osmoreceptors (Buggy et al., 1979; Ciura and Bourque, 2006; Ji et al.,

2005; Morita et al., 2004; Oldfield et al., 1994). The SFO and OVLT belong to a group of circumventricular organs (CVOs), which are periventricular regions of the brain that lack a blood-brain barrier (BBB) (Fry and Ferguson, 2007). The SFO and OVLT are specifically referred to as sensory CVOs, as they contain neuronal cell bodies and send efferent projections to multiple brain regions (Fry et al., 2007). The area postrema located above the nucleus of the solitary tract is the only other sensory CVO. The median eminence is referred to as secretory

CVO as it is comprised of non-neuronal cells, such as glia and epithelial cells (Johnson and

Gross, 1993). Located within extensive vasculature and highly fenestrated capillaries, neurons in the SFO and OVLT have direct access to the circulation and are able to sense changes in blood osmolality and circulating hormones that cannot pass BBB, such as Angiotensin II. The SFO and

OVLT are strongly interconnected with the MnPO, and the three structures collectively play a key role in systemic regulation of AVP release.

5

The SFO, MnPO, and OVLT provide both direct and indirect synaptic inputs to

magnocellular VP neurons located in the PVH and SON (Noda and Sakuta, 2013). By integrating

a multitude of osmosensory inputs provided by different areas, magnocellular VP neurons can

precisely regulate AVP release within a dynamic range of systemic osmolality. A recent study

demonstrated that the SFO and MnPO, through its glutamatergic and GABAergic projections,

exerts bidirectional control over drinking behavior, highlighting a close relationship between

behavioral and neuroendocrine responses in osmotic homeostasis (Abbott et al., 2016; Augustine

et al., 2018; Oka et al., 2015).

Figure 1.3. Systemic regulation of AVP release. Lamina terminalis (orange), comprised of the SFO, MnPO, and OVLT, relays systemic osmolality information to magnocellular AVP neurons in the PVH and SON (blue).

6

1.3 New insight into hypothalamic function

Hypothalamus is an evolutionarily conserved brain structure. The hypothalamus harbors

unique sets of neurons and neural circuits that govern nearly every aspects of bodily function

(Andermann and Lowell, 2017; Clarke, 2015; Gizowski and Bourque, 2018; Guyenet, 2006;

Morrison and Nakamura, 2019; Saper and Fuller, 2017): energy/fluid homeostasis,

thermoregulation, cardiovascular regulation, sleep-wake cycles, stress response, reproductive behavior and ingestive behavior. Despite an apparent need to understand such elemental physiological and behavioral processes, structural and molecular heterogeneity of the hypothalamus has remained a major obstacle in the research. The hypothalamus does not show clear histological organization. Anatomical denominations of the nuclei often lack functional relevance as a wide range of neuronal subtypes with unrelated or opposite functions are tightly intermingled in the structure. In the last few decades, development of genetic and molecular technology has allowed cell-type specific access to the brain and enabled systematic deconstruction of its functional and structural complexity (Haery et al., 2019). More recently,

with an advent of CRISPR (clustered regularly interspaced short palindromic repeats) technology

and highly-efficient and minimally-invasive in vivo imaging techniques, studies have brought

excitements and surprises to the hypothalamic field (Heidenreich and Zhang, 2016; Lowell,

2019).

Notable findings came from studies utilizing the newest in vivo imaging techniques such

as fiber photometry (Li et al., 2019) and endoscopic imaging (Ghosh et al., 2011). Combined

with genetically encoded neuronal activity indicators (Shen et al., 2020), such as GCaMP, these techniques allowed investigation of neural dynamics of molecularly-defined neuronal subgroups

7

in vivo while animals are freely behaving or performing tasks. One of the most groundbreaking observations was made from key hunger neurons residing in the arcuate nucleus (ARC) of the hypothalamus. These neurons, characterized by expression of Agouti-related peptide (AgRP),

Neuropeptide Y (NPY), and γ-aminobutyric acid (GABA), are a central regulator of energy homeostasis and feeding behavior (Andermann and Lowell, 2017; Augustine et al., 2020;

Sternson and Eiselt, 2017; Waterson and Horvath, 2015).

Strategically located near the median eminence, AgRP neurons have privileged access to circulating hormones, and, therefore, were thought to function as interoceptors, monitoring slow fluctuations in the energy balance. However, recent studies have proved that information carried by AgRP neurons is far more complex and diverse than those relayed by humoral cues (Betley et al., 2015; Beutler et al., 2017; Chen et al., 2015; Mandelblat-Cerf et al., 2015): AgRP neurons can encode sensory detection of food or food-predicting cues and gastrointestinal nutritional state to predict upcoming changes in energy balance. This finding, along with improved understanding of neural connections and functional diversity of AgRP neurons, challenged a dogmatic view of hypothalamus as an interoceptive structure merely responding to slow systemic fluctuations.

Presence of rapid, anticipatory responses in AgRP neurons suggests that these neurons are able to learn and foresee future systemic changes induced by food ingestion, and readily adjust animal’s behavior to prevent overconsumption. For this reason, anticipatory regulation is referred to as ‘feedforward’ homeostatic regulation, as a counterpart to ‘feedback’ regulation that comes into play after systemic perturbations develop.

Following this initial discovery, numerous studies have proven that rapid, anticipatory regulation is a common feature in many homeostatic circuits (Augustine et al., 2018; Betley et al.,

8

2015; Zimmerman et al., 2016): neurons for energy/fluid homeostasis and feeding/drinking

behaviors show rapid, anticipatory responses to food-/water-predicting cues, respectively. It is interesting to note that such anticipatory responses are highly tuned for the stimuli that are functionally relevant to each neuronal subgroup; hunger neurons respond exclusively to food- predicting cues and thirst neurons respond exclusively to water-predicting cues. This attests to highly complex and intricate wiring of the brain.

1.4 Presystemic regulation of Vasopressin neurons

Although significance of anticipatory feedforward regulation of hunger and thirst circuits is just beginning to be recognized, an idea of such regulation in the hypothalamus existed since late in the 1900s. A conceptual framework was provided by studies looking at the effect of water intake on blood vasopressin levels, from which the term ‘presystemic regulation’ originates.

‘Presystemic’ refers to a period before systemic change is triggered by a stimulus, such as changes in energy/fluid balance by food/water intake (Stricker and Hoffmann, 2007).

Presystemic regulation was first described in 1980, in a study using dehydrated dogs

(Thrasher et al., 1987; Thrasher et al., 1981). Water intake caused a significant reduction in plasma AVP concentration within 6 minutes of water intake whereas systemic osmolality was unaffected and remained elevated for 15 minutes. This phenomenon was later replicated in humans, monkeys, sheep, and rats by different groups (Arnauld and du Pont, 1982; Blair-West et al., 1985; Geelen et al., 1984; Huang et al., 2000). However, these studies were less influential than any of the recent findings in hunger and thirst circuits, most likely due to the lack of temporal resolution. Frequency of AVP measurement was limited by availability of blood

9

samples, which is every several minutes at best. Lack of temporal resolution prevented further

dissection of the characteristics of distinct stimuli causing the presystemic reductions in AVP.

Despite numerous attempts searching for the source of presystemic regulation, the

question was left unresolved for decades with puzzling findings that suggested species-specific

differences in dogs and rats (Stricker and Stricker, 2011). By directly infusing/removing water and different concentrations of saline solutions from the stomach or different locations of the gastrointestinal tracts, studies came to a rather perplexing conclusion that 1) act of swallowing

was required for the drop in AVP levels in dogs, regardless of the osmolality of ingested fluid

(Thrasher et al., 1987), and 2) osmolality of ingested fluid, which was sensed at the distal part of

gastrointestinal tract or hepatic portal system, was a major cause for AVP decrease in rats, with

act of swallowing or sensation of fluid in the oropharyngeal tract having no effect (Baertschi and

Pence, 1995; Baertschi and Vallet, 1981; Davis et al., 2002; Huang et al., 2000).

We recently undertook this challenge by replicating the experiment in mice (Mandelblat-

Cerf et al., 2017). Instead of measuring blood AVP levels, we monitored the activity of

magnocellular AVP neurons, which are the sole source of circulating AVP. By using fiber

photometry, we were able to continuously follow magnocellular AVP neuron activity while mice

are consuming food and water. The result from this study was striking (Figure 1.3): we found

that 1) magnocellular AVP neuron activity immediately decreased upon presentation of water-

predicting cues, the rapid nature of which could not have been predicted from studies using AVP

measurements , 2) magnocellular AVP neurons show bidirectional presystemic feedforward

regulation by food and water; food causes a rapid increase in magnocellular AVP neuron activity

while water decreases it, 3) presystemic regulation by food and water happens asymmetrically;

10

Figure 1.4. Regulation of AVP release: Systemic and Presystemic regulation. AVP release is regulated by presystemic and systemic mechanisms. Systemic osmolality information is detected by osmoreceptors in the SFO and OVLT and relayed to magnocellular AVP neurons directly or indirectly via the MnPO. Presystemic information is further divided into pre-ingestive and post-ingestive information. Both pre- and post- ingestive signals related to water cause decrease in magnocellular AVP neuron activity. In case of food, only post-ingestive signal is capable of increasing magnocellular AVP neuron activity. Modified from Mandelblat-cerf et al., 2017.

11

water-predicting cues and drinking are the main source of the presystemic drop in magnocellular

AVP neuron activity while the act of eating or contact with food is what initiates the presystemic increase in magnocellular AVP neuron activity, with food-predicting external sensory cues

having no effect. We have validated that the changes we observed in magnocellular AVP neuron

activity were bona fide ‘presystemic’ responses by comparing their timing of onset to that of

systemic osmolality changes (Figure 1.4). The changes in AVP neuron activity were initiated

within seconds of the stimulus onset while it required 5 and 15 min for food and water intake,

respectively, to cause a significant change in blood osmolality.

1.5 Osmotic homeostasis and significance of presystemic regulation

Osmotic homeostasis refers to biological processes that maintain body’s water content

and systemic osmolality within a narrow, physiologic range (Bourque, 2008; Danziger and

Zeidel, 2014). Water is an integral part of human body, constituting more than 70% of body

weight. Despite large fluctuations in water intake and loss during daily life, systemic osmolality

remains extremely stable in healthy individuals owing to multitudes of regulatory mechanisms

that are activated upon osmotic perturbations. Systemic hyper- or hypo-osmolality can result

from diverse situations common in everyday setting. Excessive sweating after exercise or

exposure to dry and hot environments can lead to reduced body water content and systemic

hyper-osmolality. On the other hand, excessive voluntary drinking after strenuous exercise, such

as marathon, can lead to water overload and systemic hypoosmolality. Osmotic perturbation is

12

Figure 1.5. Changes in plasma osmolality after water and food intake. Left, 24h water restriction caused a significant increase in plasma osmolality. It took 15 min for ingested water to restore baseline plasma osmolality. Right, mice were chronically food-restricted. Effect of food intake on plasma osmolality was rapid. Increase was observed as early as 2 min, and by 5 min of food intake, maximum plasma osmolality was reached.

13

also frequently observed in pathological states. Vomiting and diarrhea can lead to severe

dehydration and systemic hyperosmolality while compulsive drinking in some psychiatric

patients can lead to over-hydration and systemic hypo-osmolality (Bourque, 2008; Danziger and

Zeidel, 2014).

Maintaining osmotic homeostasis is especially important for brain function. The brain is encased in a rigid skull and the volume of the brain has to be tightly controlled to retain its morphology and functional integrity with surrounding structures. Swelling of the brain or cerebral edema causes the brain to overfill the cranial cavity causing compression of the ventricles and disruption of functional structures, such as gyri and sulci. In the case of brain shrinkage, the brain will detach from the skull and capillary structures will be torn off, leading to brain hemorrhages (Danziger and Zeidel, 2014). Therefore, even slight changes in a brain volume resulting from fluid imbalance can lead to deleterious consequences, ranging from headache, nausea, and vomiting to coma, seizure and even death (Bourque, 2008). Any situation that results in deviation of systemic osmolality from the desired physiological range initiates behavioral and physiological responses that act concurrently to bring back the water balance. For example, dehydration induces a sensation of thirst, which then initiates a set of behavioral sequences that promote water intake. Another crucial response to dehydration is release of AVP into the bloodstream, which prevents additional water loss through renal excretion.

It is now well-established that presystemic feedforward regulation exists in both nodes of osmoregulatory processes. Thirst neurons, residing in the subfornical organ (SFO) or median preoptic nucleus (MnPO) (Abbott et al., 2016; Oka et al., 2015), display a presystemic drop in

neural activity upon presentation of a water-predicting cue (Augustine et al., 2018; Betley et al.,

14

2015; Zimmerman et al., 2016). This is also true for magnocellular AVP neurons except that they also display presystemic activation in response to food intake (Figure 1.3) (Mandelblat-Cerf et

al., 2017). Systemic signals are also effective in regulating the activity of thirst and

magnocellular AVP neurons. Indeed, the SFO and MnPO constitute a larger structure called the

lamina terminalis (LT), which is a key brain region that senses and transmits systemic osmolality

signals to other brain regions (McKinley et al., 1999). The LT provides strong input to AVP

neurons, thus, mediating systemic regulation of AVP release.

Systemic and presystemic regulation work together to maintain fluid balance. Systemic changes happen slowly; it takes at least15 min for ingested water to restore blood osmolality in

dehydrated mice (Figure 1.4) (Mandelblat-Cerf et al., 2017). If thirst and AVP neurons are solely under systemic feedback control, and are inhibited at 15min, there will be an excessive amount of water from drinking and renal water reabsorption, which can potentially lead to severe

hypoosmolality. The presystemic feedforward circuit takes advantage of high speed of neural

transduction and provides the brain with an anticipatory signal well before ingested fluid is

absorbed into the system. By exerting a potent inhibitory effect on AVP secretion and sensation

of thirst well before systemic signals come into play, presystemic feedforward signals prevent fluid imbalance that might arise from temporal delay in systemic processes.

In the following chapters, I describe the results from two independent studies investigating neural basis for regulation of AVP release by presystemic anticipated disturbances

in osmolality (Chapter 2), and by hypoglycemia (Chapter 3).

15

References

Abbott, S.B., Machado, N.L., Geerling, J.C., and Saper, C.B. (2016). Reciprocal Control of Drinking Behavior by Median Preoptic Neurons in Mice. The Journal of neuroscience : the official journal of the Society for Neuroscience 36, 8228-8237.

Acher, R. (1993). Neurohypophysial peptide systems: processing machinery, hydroosmotic regulation, adaptation and evolution. Regulatory peptides 45, 1-13.

Andermann, M.L., and Lowell, B.B. (2017). Toward a Wiring Diagram Understanding of Appetite Control. Neuron 95, 757-778.

Andersson, B., Leksell, L.G., and Lishajko, F. (1975). Perturbations in fluid balance induced by medially placed forebrain lesions. Brain research 99, 261-275.

Arnauld, E., and du Pont, J. (1982). Vasopressin release and firing of supraoptic neurosecretory neurones during drinking in the dehydrated monkey. Pflugers Archiv : European journal of physiology 394, 195-201.

Augustine, V., Gokce, S.K., Lee, S., Wang, B., Davidson, T.J., Reimann, F., Gribble, F., Deisseroth, K., Lois, C., and Oka, Y. (2018). Hierarchical neural architecture underlying thirst regulation. Nature 555, 204-209.

Augustine, V., Lee, S., and Oka, Y. (2020). Neural Control and Modulation of Thirst, Sodium Appetite, and Hunger. Cell 180, 25-32.

Baertschi, A.J., and Pence, R.A. (1995). Gut-brain signaling of water absorption inhibits vasopressin in rats. The American journal of physiology 268, R236-247.

Baertschi, A.J., and Vallet, P.G. (1981). Osmosensitivity of the hepatic portal vein area and vasopressin release in rats. The Journal of physiology 315, 217-230.

Betley, J.N., Xu, S., Cao, Z.F., Gong, R., Magnus, C.J., Yu, Y., and Sternson, S.M. (2015). Neurons for hunger and thirst transmit a negative-valence teaching signal. Nature.

Beutler, L.R., Chen, Y.M., Ahn, J.S., Lin, Y.C., Essner, R.A., and Knight, Z.A. (2017). Dynamics of Gut-Brain Communication Underlying Hunger. Neuron 96, 461-+.

Blair-West, J.R., Gibson, A.P., Woods, R.L., and Brook, A.H. (1985). Acute reduction of plasma vasopressin levels by rehydration in sheep. The American journal of physiology 248, R68-71.

Bolignano, D., Cabassi, A., Fiaccadori, E., Ghigo, E., Pasquali, R., Peracino, A., Peri, A., Plebani, M., Santoro, A., Settanni, F., et al. (2014). Copeptin (CTproAVP), a new tool for understanding the role of vasopressin in pathophysiology. Clinical chemistry and laboratory medicine : CCLM / FESCC 52, 1447-1456.

16

Bourque, C.W. (2008). Central mechanisms of osmosensation and systemic osmoregulation. Nature reviews. Neuroscience 9, 519-531.

Brown, C.H., Bains, J.S., Ludwig, M., and Stern, J.E. (2013). Physiological regulation of magnocellular neurosecretory cell activity: integration of intrinsic, local and afferent mechanisms. Journal of neuroendocrinology 25, 678-710.

Buggy, J., Hoffman, W.E., Phillips, M.I., Fisher, A.E., and Johnson, A.K. (1979). Osmosensitivity of rat third ventricle and interactions with angiotensin. The American journal of physiology 236, R75-82.

Caldwell, H.K. (2017). and Vasopressin: Powerful Regulators of Social Behavior. Neuroscientist 23, 517-528.

Caldwell, H.K., Aulino, E.A., Rodriguez, K.M., Witchey, S.K., and Yaw, A.M. (2017). Social Context, Stress, Neuropsychiatric Disorders, and the Vasopressin 1b Receptor. Frontiers in neuroscience 11, 567.

Chen, Y., Lin, Y.C., Kuo, T.W., and Knight, Z.A. (2015). Sensory detection of food rapidly modulates arcuate feeding circuits. Cell 160, 829-841.

Ciura, S., and Bourque, C.W. (2006). Transient receptor potential vanilloid 1 is required for intrinsic osmoreception in organum vasculosum lamina terminalis neurons and for normal thirst responses to systemic hyperosmolality. The Journal of neuroscience : the official journal of the Society for Neuroscience 26, 9069-9075.

Clarke, I.J. (2015). Hypothalamus as an Endocrine Organ. Compr Physiol 5, 217-253.

Danziger, J., and Zeidel, M.L. (2014). Osmotic Homeostasis. Clinical journal of the American Society of Nephrology : CJASN.

Davis, J.D., Smith, G.P., and McCann, D.P. (2002). The control of water and sodium chloride intake by postingestional and orosensory stimulation in water-deprived rats. Physiology & behavior 75, 7-14.

Fenske, W.K., Schnyder, I., Koch, G., Walti, C., Pfister, M., Kopp, P., Fassnacht, M., Strauss, K., and Christ-Crain, M. (2018). Release and Decay Kinetics of Copeptin vs AVP in Response to Osmotic Alterations in Healthy Volunteers. The Journal of clinical endocrinology and metabolism 103, 505-513.

Ferguson, J.W., Therapondos, G., Newby, D.E., and Hayes, P.C. (2003). Therapeutic role of antagonism in patients with liver cirrhosis. Clinical science 105, 1-8.

Fry, M., and Ferguson, A.V. (2007). The sensory circumventricular organs: brain targets for circulating signals controlling ingestive behavior. Physiology & behavior 91, 413-423.

17

Fry, M., Hoyda, T.D., and Ferguson, A.V. (2007). Making sense of it: roles of the sensory circumventricular organs in feeding and regulation of energy homeostasis. Exp Biol Med (Maywood) 232, 14-26.

Geelen, G., Keil, L.C., Kravik, S.E., Wade, C.E., Thrasher, T.N., Barnes, P.R., Pyka, G., Nesvig, C., and Greenleaf, J.E. (1984). Inhibition of plasma vasopressin after drinking in dehydrated humans. The American journal of physiology 247, R968-971.

Ghosh, K.K., Burns, L.D., Cocker, E.D., Nimmerjahn, A., Ziv, Y., Gamal, A.E., and Schnitzer, M.J. (2011). Miniaturized integration of a fluorescence microscope. Nature methods 8, 871-878.

Gizowski, C., and Bourque, C.W. (2018). The neural basis of homeostatic and anticipatory thirst. Nat Rev Nephrol 14, 11-25.

Guyenet, P.G. (2006). The sympathetic control of blood pressure. Nature reviews. Neuroscience 7, 335-346.

Haery, L., Deverman, B.E., Matho, K.S., Cetin, A., Woodard, K., Cepko, C., Guerin, K.I., Rego, M.A., Ersing, I., Bachle, S.M., et al. (2019). Adeno-Associated Virus Technologies and Methods for Targeted Neuronal Manipulation. Frontiers in neuroanatomy 13, 93.

Heidenreich, M., and Zhang, F. (2016). Applications of CRISPR-Cas systems in neuroscience. Nature reviews. Neuroscience 17, 36-44.

Huang, W., Sved, A.F., and Stricker, E.M. (2000). Water ingestion provides an early signal inhibiting osmotically stimulated vasopressin secretion in rats. American journal of physiology. Regulatory, integrative and comparative physiology 279, R756-760.

Janaky, T., Laszlo, F.A., Sirokman, F., and Morgat, J.L. (1982). Biological half-life and organ distribution of [3H]8-arginine-vasopressin in the rat. The Journal of endocrinology 93, 295-303.

Japundzic-Zigon, N., Lozic, M., Sarenac, O., and Murphy, D. (2020). Vasopressin & Oxytocin in Control of the Cardiovascular System: An Updated Review. Current neuropharmacology 18, 14- 33.

Ji, L.L., Fleming, T., Penny, M.L., Toney, G.M., and Cunningham, J.T. (2005). Effects of water deprivation and rehydration on c-Fos and FosB staining in the rat supraoptic nucleus and lamina terminalis region. American journal of physiology. Regulatory, integrative and comparative physiology 288, R311-321.

Johnson, A.K., and Gross, P.M. (1993). Sensory circumventricular organs and brain homeostatic pathways. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 7, 678-686.

Koshimizu, T.A., Nakamura, K., Egashira, N., Hiroyama, M., Nonoguchi, H., and Tanoue, A. (2012). Vasopressin V1a and V1b receptors: from molecules to physiological systems. Physiological reviews 92, 1813-1864.

18

Leng, G., Brown, C.H., and Russell, J.A. (1999). Physiological pathways regulating the activity of magnocellular neurosecretory cells. Progress in neurobiology 57, 625-655.

Li, Y., Liu, Z., Guo, Q., and Luo, M. (2019). Long-term Fiber Photometry for Neuroscience Studies. Neurosci Bull 35, 425-433.

Lowell, B.B. (2019). New Neuroscience of Homeostasis and Drives for Food, Water, and Salt. The New England journal of medicine 380, 459-471.

Ludwig, M., and Leng, G. (2006). Dendritic peptide release and peptide-dependent behaviours. Nature reviews. Neuroscience 7, 126-136.

Mandelblat-Cerf, Y., Kim, A., Burgess, C.R., Subramanian, S., Tannous, B.A., Lowell, B.B., and Andermann, M.L. (2017). Bidirectional Anticipation of Future Osmotic Challenges by Vasopressin Neurons. Neuron 93, 57-65.

Mandelblat-Cerf, Y., Ramesh, R.N., Burgess, C.R., Patella, P., Yang, Z., Lowell, B.B., and Andermann, M.L. (2015). Arcuate hypothalamic AgRP and putative POMC neurons show opposite changes in spiking across multiple timescales. eLife 4.

McKinley, M.J., Gerstberger, R., Mathai, M.L., Oldfield, B.J., and Schmid, H. (1999). The lamina terminalis and its role in fluid and electrolyte homeostasis. Journal of clinical neuroscience : official journal of the Neurosurgical Society of Australasia 6, 289-301.

Morita, H., Ogino, T., Fujiki, N., Tanaka, K., Gotoh, T.M., Seo, Y., Takamata, A., Nakamura, S., and Murakami, M. (2004). Sequence of forebrain activation induced by intraventricular injection of hypertonic NaCl detected by Mn2+ contrasted T1-weighted MRI. Autonomic neuroscience : basic & clinical 113, 43-54.

Morrison, S.F., and Nakamura, K. (2019). Central Mechanisms for Thermoregulation. Annu Rev Physiol 81, 285-308.

Noda, M., and Sakuta, H. (2013). Central regulation of body-fluid homeostasis. Trends in neurosciences 36, 661-673.

Oka, Y., Ye, M., and Zuker, C.S. (2015). Thirst driving and suppressing signals encoded by distinct neural populations in the brain. Nature.

Oldfield, B.J., Badoer, E., Hards, D.K., and McKinley, M.J. (1994). Fos production in retrogradely labelled neurons of the lamina terminalis following intravenous infusion of either hypertonic saline or angiotensin II. Neuroscience 60, 255-262.

Saper, C.B., and Fuller, P.M. (2017). Wake-sleep circuitry: an overview. Current opinion in neurobiology 44, 186-192.

19

Shen, Y., Nasu, Y., Shkolnikov, I., Kim, A., and Campbell, R.E. (2020). Engineering genetically encoded fluorescent indicators for imaging of neuronal activity: Progress and prospects. Neuroscience research.

Sternson, S.M., and Eiselt, A.K. (2017). Three Pillars for the Neural Control of Appetite. Annu Rev Physiol 79, 401-423.

Stricker, E.M., and Hoffmann, M.L. (2007). Presystemic signals in the control of thirst, salt appetite, and vasopressin secretion. Physiology & behavior 91, 404-412.

Stricker, E.M., and Stricker, M.L. (2011). Pre-systemic controls of fluid intake and vasopressin secretion. Physiology & behavior 103, 86-88.

Thrasher, T.N., Keil, L.C., and Ramsay, D.J. (1987). Drinking, oropharyngeal signals, and inhibition of vasopressin secretion in dogs. The American journal of physiology 253, R509-515.

Thrasher, T.N., Nistal-Herrera, J.F., Keil, L.C., and Ramsay, D.J. (1981). Satiety and inhibition of vasopressin secretion after drinking in dehydrated dogs. The American journal of physiology 240, E394-401.

Waterson, M.J., and Horvath, T.L. (2015). Neuronal Regulation of Energy Homeostasis: Beyond the Hypothalamus and Feeding. Cell metabolism 22, 962-970.

Zimmerman, C.A., Lin, Y.C., Leib, D.E., Guo, L., Huey, E.L., Daly, G.E., Chen, Y., and Knight, Z.A. (2016). Thirst neurons anticipate the homeostatic consequences of eating and drinking. Nature 537, 680-684.

20

Chapter 2

Regulation of Vasopressin Secretion by Presystemic Anticipated Disturbances in Osmolality

Attributions: Angela Kim performed all experiments for this chapter unless otherwise stated.

Joseph C. Madara performed brain slice electrophysiology.

21

2.1 Introduction

The body’s main response to dehydration or elevated systemic osmolality is to increase circulating vasopressin (AVP). AVP, also known as an antidiuretic hormone, is synthesized by magnocellular AVP neurons of the paravnetricular (PVH) and supraoptic nucleus (SON) of the hypothalamus. Upon stimulation of magnocellular AVP neurons, AVP is released from the axon terminals in the posterior pituitary and enter the bloodstream (Bolignano et al., 2014; Ferguson et al., 2003). Once released, AVP binds to receptors in the kidney and stimulates water reabsorption to restore water balance (Nielsen et al., 1995). Activity of magnocellular AVP neurons is tightly regulated by two temporally distinct mechanisms. Systemic regulation is mainly mediated by the lamina terminalis (LT), which is comprised of three structures, the subfornical organ (SFO), median preoptic nucleus (MnPO), and organum vasculosum lamina terminalis (OVLT)

(McKinley et al., 2004). The SFO and OVLT are circumventricular organs that lack a blood- brain barrier, and are capable of sensing systemic osmolality and blood-borne factors, such as angiotensin II. Information sensed by the SFO and OVLT is then sent to magnocellular AVP neurons directly or indirectly through the MnPO to regulate AVP release. Additional fine tuning of AVP release is achieved by a mechanism termed presystemic regulation. As its name suggests, presystemic regulation provides feedforward information about anticipated future osmotic perturbations and readily adjust the activity of magnocellular AVP neurons. Presystemic regulation is thought to function as a safety measure to protect animals from harmful effects of overconsumption caused by the temporal delay in systemic regulation (Stricker and Hoffmann,

2007; Stricker and Stricker, 2011).

22

Using optetrode recordings and fiber photometry, we recently demonstrated that magnocellular AVP neurons are under bidirectional presystemic regulation by water and food

(Mandelblat-Cerf et al., 2017). Interestingly, presystemic regulation by water and food happened asymmetrically; water-predicting cue and drinking cause decreases in AVP neuron activity while feeding, but not food-predicting cues, cause an increase in AVP neuron activity. Presystemic regulation has been also described in the feeding and thirst circuit. It is now well-established that food- and water-predicting cue causes a rapid inhibition of AgRP neurons (Betley et al., 2015;

Chen et al., 2015; Mandelblat-Cerf et al., 2015) and thirst neurons (Augustine et al., 2018; Betley et al., 2015; Zimmerman et al., 2016), respectively. Despite clear demonstration of presystemic regulation in different homeostatic circuits, the underlying mechanisms remains largely unknown.

In the present study, we identified and characterized the neural circuits mediating

presystemic regulation of magnocellular AVP neurons using in vivo calcium imaging and opto-

and chemo-genetic approaches. We demonstrate that presystemic regulation of magnocellular

AVP neurons is mediated by two functionally-distinct neural circuits that exclusively transmit

the information related to water or food. We also validate a brainstem circuit mediating blood

pressure information to magnocellular AVP neurons that is completely separate from water- and food-related presystemic circuits. Taken together, our data show that activity of magnocellular

AVP neurons is tightly regulated by multiple functionally-distinct neural circuits that convey specific systemic and presystemic information.

23

2.2 Results

2.2.1 Vasopressin neurons receive inhibitory and excitatory inputs from the LT

To identify regions providing direct input to magnocellular AVP neurons, we performed retrograde rabies mapping from magnocellular PVHAVP and SONAVP neurons using AVP-IRES-

Cremice. In order to target magnocellular SONAVP neurons, we injected rabies virus directly into

the SON (Figure 2.1a). This was possible because entire SON neuronal population is

magnocellular. In contrast, the PVH contains parvocellular and magnocellular AVP neurons,

which are both marked by AVP-IRES-Cre. Due to heterogeneity of the AVP neurons in the PVH,

we devised a retrograde approach that allows us to target magnocellular neurons specifically

(Figure 2.1e). Rabies virus was injected into the posterior pituitary instead of the PVH so that

only posterior pituitary-projecting, TVA-expressing PVHAVP neurons are infected. Starter magnocellular PVHAVP and SONAVP neurons were identified by co-expression of rabies-GFP and

TVA-mCherry (Figure 2.1b, f). Despite their distinct anatomical locations, results from magnocellular PVHAVP and SONAVP neurons were strikingly similar. Rabies tracing showed that magnocellular PVHAVP and SONAVP neurons receive inputs from all three structures of the lamina terminalis (LT), namely, the subfornical organ (SFO), the median preoptic nucleus

(MnPO), and the organum vasculosum lamina terminalis (OVLT), which are key sites for central osmosensing and regulation of thirst (Figure 2. 1d, h). It is worth noting that we found dense rabies labeling in the PVH when inputs to SONAVP neurons were examined, and vice versa

(Figure 2. 1c, g). Following series of antero- and retrograde tracing experiments (not shown), we

24

concluded that these are artifacts of rabies tracing presumably resulting from non-synaptic

traveling of the rabies, and there are no functional connections between the PVH and SON.

Figure 2.1. Magnocellular AVP neurons receive excitatory and inhibitory input from the LT. a, e, Schematic of monosynaptic rabies tracing from magnocellular PVHAVP (a) and SONAVP (e) neurons. In order to target magnocellular PVHAVP neurons, rabies virus was injected into the posterior pituitary. b, f, Representative images showing magnocellular PVHAVP (c) and SONAVP (f) starter neurons as identified by co-expression of GFP and mCherry. c, g, Representative images showing magnocellular PVHAVP (c) and SONAVP (g) starter neurons and dense rabies labeling in the SON (c) and PVH (g), which were identified to be technical artifacts. d, h, Representative images showing sites containing rabies-labeled neurons in the LT that are monosynaptically connected to magnocellular PVHAVP (d) and SONAVP (h) neurons. i, Representative images showing expression of Syn-YFP in excitatory and inhibitory populations of the SFO and MnPO/OVLT (left box), and their efferent projections in the SFO, MnPO, OVLT, PVH and SON (right box). Note lack of YFP-labeled fibers from SFOVgat neurons in the PVH and SON.

25

Figure 2.1 (Continued) j, Representative images showing co-localization of GFP and AVP immunofluorescence (red) in the PVH (top) and SON (bottom) of AVP-GFP mice. k, l, Number of PVHAVP and SONAVP neurons (e) and non-GFP PVH and SON neurons (f) receiving direct synaptic inputs from MnPO/OVLTVglut2, SFOVglut2, and MnPO/OVLTVgat neurons as identified by CRACM. Mice used include AVP-GFP;Vglut2-IRES-Cre (MnPO/OVLTVglut2 and SFOVglut2) and AVP-GFP;Vgat-IRES-Cre (MnPO/OVLTVgat) Scale bar, 200µm.

In order to identify the subtypes of LT neurons providing inputs to magnocellular

PVHAVP and SONAVP neurons, we performed anterograde tracing from glutamatergic and

GABAergic populations of the LT. We injected cre-dependent anterograde tracer, synaptophysin

(AAV-DIO-Syn-YFP) into the SFO, and MnPO/OVLT of Vglut2-IRES-Cre and Vgat-IRES-Cre

to individually trace glutamatergic and GABAergic projections. The MnPO and OVLT are

considered as one group in this study because targeting of individual structures was extremely

challenging due to close proximity of two structures. We found that the MnPO/OVLT sends both

excitatory and inhibitory projections to the PVH and SON where AVP neurons reside while the

SFO only sends excitatory projections (Figure 2. 1i). Lack of long-range inhibitory projections of the SFO has been reported previously (Oka et al., 2015).

26

Next, we performed Channelrhodopsin (ChR2)-assisted circuit mapping (CRACM) to

test synaptic connectivity of these projections to PVHAVP and SONAVP neurons. AVP-GFP mice were crossed to either Vglut2-IRES-Cre or Vgat-IRES-Cre mice to test the excitatory and inhibitory LT inputs individually. AAV-DIO-ChR2-mCherry was injected into the SFO or

MnPO/OVLT of these mice and light-evoked postsynaptic currents were recorded from PVHAVP

and SONAVP neurons, identified by GFP expression (Figure 2. 1j). It is worth noting that we did

not distinguish parvocellular and magnocellular PVHAVP neurons in this experiment because

GFP marks both AVP neurons in AVP-GFP mice. We found that SFOVglut2 and

MnPO/OVLTVglut2 neurons provided robust glutamatergic input to PVHAVP and SONAVP neurons, with nearly 100% connectivity (Figure 2. 1k). In contrast, GABAergic input from

MnPO/OVLTVgat neurons was detected only in half of PVHAVP and SONAVP neurons we recorded. Next, we tested the synaptic connectivity of the LT to non-GFP neurons in the PVH and SON (Figure 2. 1l). The MnPO/OVLT provided glutamatergic and GABAergic input to the majority of non-GFP neurons in the PVH and SON. Interestingly, we did not find any connection from SFO Vglut2 neurons to non-GFP neurons in the PVH. This suggests that SFOVglut2 neurons

provide restricted inputs to a defined PVH population but further study is needed to validate this

finding. Together, these results demonstrate that PVHAVP and SONAVP neurons receive both

excitatory and inhibitory inputs from the LT.

2.2.2 Water-related presystemic regulation of vasopressin neurons is mediated by the LT

Next, we examined the involvement of the LT in presystemic regulation of SONAVP

neurons using fiber photometry. We decided to focus on SONAVP neurons in the following

27

photometry experiments as the SON is exclusively comprised of magnocellular AVP neurons.

We measured the in vivo calcium dynamics of GCaMP6s-expressing SONAVP neurons while inhibiting the MnPO/OVLT or SFO with hM4Di (Figure 2. 2a and i). When recording was performed with mice expressing calcium-independent fluorescent marker, EYFP, instead of

GCaMP6s in SONAVP neurons, we did not observe any consistent time-locked responses to drinking or feeding. This validates the lack of drinking- or feeding-related movement artifact in our system (Supplementary Figure 2. 1). To monitor water-related presystemic responses in

SONAVP neurons, mice were chronically water-restricted and habituated to an experimental paradigm. Experiments were performed in the home cage and mice were expected to drink from a water bowl that was placed in the cage at the beginning of each session with random latency.

Mice were allowed to drink freely for ~15 min. Each mouse had one experimental session per day, in which they were pre-injected with either saline or CNO. In the control group, SONAVP

neurons were rapidly inhibited by the presentation of the water bowl (Figure 2. 2b) as reported

previously (Mandelblat-Cerf et al., 2017). SONAVP activity continuously declined as drinking continued until it reached a stable value. Inhibition of the MnPO/OVLT abolished or even reversed the initial pre-ingestive drop in SONAVP neurons activity caused by water bowl

presentation (Figure 2. 2b-f). Drinking-induced inhibition was also significantly attenuated. The

MnPO contains thirst neurons that are necessary and sufficient to induce drinking behavior

(Abbott et al., 2016; Augustine et al., 2018). MnPO/OVLT inhibition in our study did not prevent drinking but delayed drinking onset as shown by a significant increase in the latency

(Figure 2. 2f-g). The number of drinking bouts was not affected. The MnPO/OVLT, along with

28

Figure 2.2. The LT mediates water-related presystemic regulation of SONAVP neurons. a, Schematic of SONAVP photometry experiment with hM4Di-mediated non-specific inhibition of the MnPO/OVLT.

AVP b, Heatmap showing single-trial timecourses of SON population activity in response to water bowl placement in saline and CNO trials. Trials are sorted according to latency from water bowl placement to drinking onset (black ticks). n= 6 mice.

AVP c, Average SON population activity in response to water bowl placement in saline and CNO trials. n= 6 mice. d, Average of pre-ingestive responses in saline and CNO trials. n= 6 mice. e, Average SONAVP population activity binned across drinking periods. ***p < 0.001; repeated measures (RM) 2-way ANOVA, n= 6 mice. f-h, Average pre- and post-ingestive responses (f), average latency to drinking onset, number of drinking bouts (g), and average change in baseline activity (h) in saline and CNO trials. ns, p > 0.05; *p < 0.05; **p<0.01; paired t-test, n=6 mice.

29

Figure 2.2. (Continued) i, Schematic of SONAVP photometry experiment with hM4Di-mediated inhibition of the MnPO/OVLT. j, Heatmap showing single-trial timecourses of SONAVP population activity in response to water bowl placement in saline and CNO trials. Trials are sorted according to latency from water bowl placement to drinking onset (black ticks). n= 5 mice. k, Average SONAVP population activity in response to water bowl placement in saline and CNO trials. n= 5 mice. l, Average SONAVP population activity binned across drinking periods. ns, p > 0.05; RM 2-way ANOVA; n= 5 mice. Values are means ± SEMs.

the SFO, is the main source of excitatory input that activates SONAVP neurons during systemic hyperosmolality (McKinley et al., 2004). We found that MnPO/OVLT inhibition caused a drop in baseline activity of SONAVP neurons (Figure 2. 2h). However, SONAVP neurons were not

completely silenced by MnPO/OVLT inhibition as they still showed a drop in activity in

30

response to water bowl presentation and drinking (Figure 2. 2b-f). Residual activity is likely provided by the SFO, which is not inhibited by hM4Di in these mice. In a separate group of mice, we investigated the effect of SFO inhibition (Figure 2. 2i). Inhibition of the SFO caused a slight attenuation of the water bowl-induced inhibition of SONAVP neurons but was not effective enough to cause a significant change (Figure 2. 2j-l). Together, these data indicate that the

MnPO/OVLT is a main mediator of water-related presystemic signals to AVP neurons, with the

SFO playing a minor role.

2.2.3 Presystemic signals enter the LT at the level of MnPO

Next, we sought to understand the organization of presystemic circuit mediating water-

related information within and upstream of the LT. Both the SFO and MnPO/OVLT provide

strong excitatory inputs to AVP neurons and are capable of regulating AVP neuron activity in

response to changes in systemic osmolality. However, we found that only the MnPO/OVLT was

required for water-related presystemic regulation of SONAVP neurons (Figure 2. 2), suggesting that the presystemic circuit might be organized similarly to the hierarchical thirst circuit

described previously (Augustine et al., 2018). To test this hypothesis, we mapped inputs to SON-

projecting MnPO/OVLTVglut2 and SFO Vglut2 neurons using monosynaptic rabies tracing. Similar to our retrograde approach targeting magnocellular AVP neurons (Figure 2. 1a), we injected the rabies virus into the SON of Vglut2-IRES-Cre mice that already had TVA-mCherry and RG expressed in the MnPO/OVLT or SFO (Figure 2. 3a and d). TVA is known to be trafficked to the axon terminals (Betley et al., 2013; Livneh et al., 2017). Rabies virus can be taken up by TVA-

expressing axons of MnPO/OVLTVglut2 and SFO Vglut2 neurons in the SON, allowing projection-

31

Figure 2.3. Organization of water-related presystemic neural circuit upstream and downstream of the LT. a, Schematic of monosynaptic rabies tracing from MnPO/OVLTVglut2 neurons. b, c, Representative images showing MnPO/OVLTVglut2 starter neurons as identified by co- expression of GFP and mCherry (b), and sites containing rabies-labeled neurons that are monosynaptically connected to MnPO/OVLTVglut2 neurons (c). d, Schematic of monosynaptic rabies tracing from SFOVglut2 neurons. e, f, Representative image showing SFOVglut2 starter neurons as identified by co-expression of GFP and mCherry (e), and sites containing rabies-labeled neurons that are monosynaptically connected to SFOVglut2 neurons (f). g, h, Representative images showing SON-projecting SFOVglut2 starter neurons as identified by co-expression of GFP and mCherry (g), and rabies-labeled collateral projections of SON- projecting SFOVglut2 neurons in the MnPO, OVTL, PVH and SON (h). i, h in higher magnification. j, k, Representative images showing SON-projecting MnPO/OVLTVglut2 starter neurons as identified by co-expression of GFP and mCherry (g), and rabies-labeled collateral projections of SON-projecting MnPO/OVLTVglut2 neurons in the MnPO, OVTL, PVH and SON (h). l, h in higher magnification. Scale bar, 200µm.

32

specific rabies infection. The MnPO/OVLT and SFO contained large numbers of retrogradely

infected starter neurons that were identified by colocalization of TVA-mCherry and rabies-GFP

signals (Figure 2. 3b and e). Results from the MnPO/OVLT and SFO were strikingly different.

SON-projecting MnPO/OVLT Vglut2 neurons received inputs from multiple areas (Figure 2. 3c),

including the SFO, PVH, ventromedial (VMH), and dorsomedial nuclei (DMH) of the

hypothalamus, and the lateral parabrachial nucleus (LPBN). Among these areas, the SFO

contained the greatest number of rabies-labeled neurons, demonstrating strong interconnectivity

between the LT structures. No other sites contained rabies-labeled neurons. In contrast, SON-

projecting SFO Vglut2 neurons received inputs only from the other two LT structures (Figure 2. 3f),

the MnPO and OVLT. Lack of inputs from regions outside the LT was validated using Nos-1-

IRES-Cre, another cre line that specifically marks excitatory SFO neurons (Supplementary

Figure 2. 2) (Zimmerman et al., 2016). These data demonstrate that the MnPO/OVLT is the only

LT region receiving inputs from the regions outside the LT and, thus, a main entry point of presystemic information to the LT.

Next, we sought to understand the organization of LT projections to AVP neurons. We performed rabies collateral mapping from SFOVglut2 neurons to examine whether these neurons

send collateral projections to the MnPO/OVLT, PVH, and SON. Rabies collateral mapping was

performed as described previously (Livneh et al., 2020). Expression of TVA only in SFOVglut2

neurons and subsequent injection of rabies virus into one of the downstream regions allowed

infection of discrete SFOVglut2 population projecting to the rabies injection site (Figure 2. 3 g and j). Rabies infection then drove intense GFP expression, visualizing the entire projection pattern of SFO Vglut2 neurons projecting to the rabies-injected site. Injection of the rabies into the SON caused a dense labeling of the fibers in the SON validating successful infection of SON-

33

projecting SFOVglut2 population (Figure 2. 3g, h). Rabies-labeled fibers were also present in the

PVH, MnPO, and OVLT but the density was much lower than those in the SON (Figure 2. 3i).

Rabies injection into the MnPO/OVLT resulted in comparable amount of rabies-labeled fibers in the PVH and SON (Figure 2. 3j-l). A similar study performed by another group reported extensive collateralization of paraventricular thalamus-projecting MnPO/OVLTVglut2 neurons

(Livneh et al., 2020). These results demonstrate that the LT sends collateral projections to magnocellular AVP neurons in the PVH and SON.

2.2.4 SON-projecting LT neurons show presystemic response to water-predicting cues and drinking

Next, we measured in vivo calcium dynamics of SON-projecting MnPO/OVLT and SFO neurons in response to water bowl presentation. GCaMP6s was expressed in SON-projecting

excitatory or inhibitory neurons of the MnPO/OVLT and SFO using modified herpes simplex

virus (HSV) that cre-dependently expresses GCaMP6s (HSV-GCaMP6s). This virus is capable

of retrogradely infecting neurons from the axon terminals and has been used to study the activity

of projection-defined neurons in the other region (Kohl et al., 2018). HSV-GCaMP6s was injected into the SON of Vglut2-IRES-Cre or Vgat-IRES-Cre mice and an optic fiber was place above the MnPO/OVLT or SFO (Figure 2. 4a, g). This approach was chosen for two reasons: 1) it allows simple, fast, and efficient targeting of SON-projecting LT populations compared to other retrograde methods that require multiple viral injections and often accompany neurotoxicity, and 2) it gives highly selective access to LT populations directly upstream of

SONAVP neurons as the SON contains only two known neuronal subtypes, AVP- and oxytocin-

34

Figure 2.4. SON-projecting SFOVglut2, MnPO/OVLTVglut2, and MnPO/OVLTVgat neurons show presystemic responses to water bowl placement and drinking. a, Schematic of photometry experiment from SON-projecting SFOVglut2 and MnPO/OVLTVglut2 neurons. b, Heatmap showing single-trial timecourses of SON-projecting SFOVglut2 (top) and Vglut2 MnPO/OVLT (bottom) population activity in response to water bowl placement. Trials are sorted according to latency from water bowl placement to drinking onset (black ticks). n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2) mice.

Vglut2 Vglut2 c, Average population activity of SON-projecting SFO (light blue) and MnPO/OVLT (dark blue) neurons in response to water bowl placement. n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2) mice. d, Average pre-ingestive responses of SON-projecting SFOVglut2 (light blue) and MnPO/OVLTVglut2 (dark blue) neurons. n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2) mice. e, Average SON-projecting SFOVglut2 (light blue) and MnPO/OVLTVglut2 (dark blue) population activity binned across drinking periods. n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2) mice. f, Normalized average SON-projecting SFOVglut2 and MnPO/OVLTVglut2 population activity binned scross drinking periods. Values are normalized to the total change. ***p < 0.001; 2-way ANOVA, n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2) mice.

35

Figure 2.4. (Continued) g, Schematic of photometry experiment from SON-projecting MnPO/OVLTVgat neurons.

Vgat h, Single-trial timecourses of SON-projecting MnPO/OVLT population activity in response to water bowl placement. Trials are sorted according to latency from water bowl placement to drinking onset (black ticks). n= 9 mice.

Vgat i, Average population activity of SON-projecting MnPO/OVLT neurons in response to water bowl placement. n= 9 mice. j, Average pre-ingestive responses of SON-projecting MnPO/OVLTVgat neurons. n= 9 mice. k, Average SON-projecting MnPO/OVLTVgat population activity across drinking periods. n= 9 mice. Scale bar, 200µm. Values are means ± SEMs.

36

expressing neurons, and LT projections to surrounding structures are minimal. We found that

SON-projecting MnPO/OVLTVglut2 and SFOVglut2 neurons showed an immediate drop in activity

upon water bowl placement and a further drop was observed as drinking continued (Figure 2. 4b-

e). This resembled the response of SONAVP neurons, suggesting that inputs from these neurons

underlie water-related presystemic regulation of SONAVP neurons. When responses of

MnPO/OVLTVglut2 and SFOVglut2 neurons were directly compared by normalizing the values to the total change, MnPO/OVLTVglut2 neurons showed a larger and sharper decrease in the activity in response to water bowl placement compared to SFOVglut2 neurons (Figure 2. 4f). Activity of

MnPO/OVLT Vglut2 neurons reached <50% of baseline within the first 1 min of drinking, whereas

SFO neurons showed a gradual decrease that is continued for over 5 min. Interestingly, the SON- projecting MnPO/OVLTvgat neurons showed a completely different pattern of response (Figure 2.

4h-k). Placement of the water bowl caused an abrupt increase in the activity of these neurons, which rapidly declined and returned to the baseline within 1 min. No drinking-related, long-term response was observed. Considering that these neurons are providing inhibitory inputs to AVP neurons, we hypothesize that SON-projecting MnPO/OVLTvgat neurons mainly signal the anticipation of water and are responsible for a rapid drop in AVP neuron activity immediately upon water bowl placement. These data indicate that the MnPO/OVLT is a key relay center for water-related presystemic information.

2.2.5 The LT does not mediate food-related presystemic regulation of vasopressin neurons

We found that the LT is involved in water-related presystemic regulation of SONAVP neurons with the MnPO/OVLT playing a key role. We next tested whether the same region was

37

involved in food-related presystemic regulation. Using the same approach, we recorded in vivo

calcium dynamics of SONAVP neurons while inhibiting the MnPO/OVLT using hM4Di (Figure 2.

2a). Mice were chronically food-restricted and were trained to take food pellets from a food bowl

placed in the cage randomly during the experimental session. In the control group, SONAVP neurons showed an increase in activity that began immediately upon feeding onset (Figure 2. 5a).

Unlike the response to a water bowl, placement of the food bowl had no effect on SONAVP

neurons activity. MnPO/OVLT inhibition did not affect the baseline activity (not shown) or the

feeding-induced activation of SONAVP neurons (Figure 2. 5a-c). These results demonstrate that

food-related presystemic regulation is mediated by a distinct circuit that does not involve the

MnPO/OVLT.

We then explored the activity of LT neurons during food intake. GCaMP6s was

expressed in SON-projecting MnPO/OVLTvglut2, SFOvglut2, and MnPO/OVLTvgat using HSV-

LSL-GCaMP6s as described above (Figure 2. 4a, g). In line with the result from MnPO/OVLT inhibition, none of the LT neurons we recorded displayed a response that can explain feeding- induced activation observed in SONAVP neurons (Figure 2. 5d). When traces were aligned to the feeding onset, SON-projecting MnPO/OVLT Vglut2 and MnPO/OVLT Vgat neurons showed a

decrease in activity (Figure 2. 5e). Simultaneous decrease in excitatory and inhibitory tone from

the MnPO/OVLT is unlikely to underlie a rapid activation of SONAVP neurons by feeding.

A previous study demonstrated rapid activation of excitatory SFO neurons by food intake and claimed this response was presystemic and occurred before any change in systemic osmolality (Zimmerman et al., 2016). To test whether a larger, or different excitatory SFO population was involved in food-related presystemic regulation of AVP neurons, we performed

38

Figure 2.5. The MnPO/OVLT is not involved in food-related presystemic regulation of SONAVP neurons.

AVP a, Heatmap showing single-trial timecourses of SON population activity in response to food bowl placement in saline and CNO trials. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks). n=5 mice.

AVP b, Average population activity of SON neurons in response to feeding onset in saline and CNO trials. n=5 mice. c, Average SONAVP population activity binned across feeding periods. n=5 mice.

39

Figure 2.5. (Continued) d, Single-trial timecourses of SON-projecting MnPO/OVLTVglut2, SFOVglut2, and Vgat MnPO/OVLT population activity in response to food bowl placement. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks). n= 7 (SFOVglut2), and 5 (MnPO/OVLTVglut2), and 9 (MnPO/OVLTVgat) mice.

Vglut2 Vgat e, Average population activity of SON-projecting MnPO/OVLT , and MnPO/OVLT neurons aligned to feeding onset. n= 5 (MnPO/OVLTVglut2), and 9 (MnPO/OVLTVgat) mice.

AVP Vglut2 Vglut2 f, Average population activity of SON , SFO , and SON-projecting SFO neurons aligned to feeding onset. n= 5 (SONAVP), 3 (SFOVglut2), and 7 (SON-projecting SFOVglut2) mice. Values are means ± SEMs.

fiber photometry from the entire SFOVglut2 population. We observed an increase in SFOVglut2

neuron activity as previously reported (Figure 2. 5f). However, when the response was directly

compared to the food-related presystemic response of SON AVP neurons, the response of

SFOVglut2 neurons was delayed and lacked an early feeding-induced activation occurring within 1 min of feeding onset. In a previous study, we have shown that an increase in blood osmolality occurs within 2 min of feeding (Figure 1.4) (Mandelblat-Cerf et al., 2017). This strongly suggests that feeding-induced activity in the entire SFOVglut2 population reflects systemic

osmolality changes rather than presystemic signals. However, it is difficult to predict the precise

onset of the feeding-induced increase in systemic osmolality as continuous real-time monitoring

of blood osmolality is not feasible in mice. All together, these results further support the

hypothesis that food- and water-related presystemic regulation of SONAVP neurons is mediated

by distinct neural circuits.

2.2.6 Vasopressin neurons receive excitatory inputs from A1/C1 neurons in the VLM

40

Figure 2.6. SONAVP neurons receive direct glutamatergic inputs from A1/C1 neurons of the VLM. a, b, Representative images showing expression of ChR2-mCherry in A1/C1 neurons of the VLM and A2, NTSVglut2, and entire NTS populations (a), and their efferent projections in the PVH and SON (b). c, d, Number of PVHAVP and SONAVP neurons (c), and non-GFP PVH and SON neurons (d) receiving direct synaptic inputs from A1/C1 neurons of the VLM and A2, NTSVglut2, and entire NTS populations as identified by CRACM. Mice used include AVP-GFP;Th-IRES-Cre (A1/C1, and A2 neurons), AVP-GFP;Vglut2-IRES-Cre (NTSVglut2), and AVP-GFP (entire NTS population). e, Representative trace showing light-evoked postsytnaptic currents from A1/C1 neurons to SONAVP neurons (left). The response was blocked by DNQX (middle) but was not affected by TTX/4-AP (right) demonstrating monosynaptic glutamatergic connection. Scale bar, 500µm.

41

A study using a traditional retrograde tracer in the SON showed that neurons in the SON

receive projections from the nucleus of the solitary tract (NTS) and A1/C1 neurons in the

ventrolateral medulla (VLM). We investigated whether any of these areas provide direct inputs

to AVP neurons. First, we explored the projections of different subtypes of NTS neurons and

A1/C1 neurons. We targeted three different NTS neuron subtypes using specific cre lines: A2

neurons with TH-IRES-Cre, excitatory neurons with Vglut2-IRES-Cre, and GLP-1 neurons with

Gcg-Cre (Figure 2. 6a). We also studied the projection of the entire NTS population using cre-

independent ChR2 (AAV-ChR2-mCherry). A1/C1 neurons in the VLM were targeted using TH-

IRES-Cre. We found that A1/C1 neurons send dense projections to the PVH and SON where

AVP neurons reside (Figure 2. 6b) as previously known (Card et al., 2006) . Projections from the

entire NTS population and individual NTS neuron subtypes were extremely sparse in the PVH

and were nearly absent in the SON. We subsequently performed CRACM to test the synaptic

connectivity of these brainstem projections to AVP neurons (Figure 2. 6c). AVP-GFP mice were

crossed to appropriate cre lines to achieve specific expression of ChR2 in each subtype of NTS

neurons and A1/C1 neurons. In line with the findings from projection mapping, only A1/C1

neurons provided robust direct synaptic input to PVHAVP and SONAVP neurons. This input was glutamatergic as it was blocked by a glutamate receptor blocker, DNQX (Figure 2. 6d). It is worth noting that the negative CRACM result from the NTS does not exclude the possibility that the NTS is providing neuromodulatory input to PVHAVP and SONAVP neurons. However, this

was not investigated further in our study because we concluded that slow postsynaptic effect of

neuromodulatory input cannot account for rapid presystemic responses in SONAVP neurons.

42

2.2.7 A1/C1 neurons in the VLM mediate hypotension-induced, but not feeding-induced,

activation of vasopressin neurons

To test the involvement of A1/C1 neurons in food-related presystemic regulation of AVP

neurons, we inhibited A1/C1 neurons with hM4Di while monitoring in vivo calcium dynamics of

AVP neurons. In order to achieve specific expression of hM4Di in A1/C1 neurons, we used

AVP-IRES-Cre;DBH-Flp mice in combination with a custom-made flp-dependent hM4Di virus

(AAV-fDIO-hM4Di-mCherry) (Figure 2. 7a). First, to test the effectiveness of hM4Di in silencing A1/C1 neurons, we studied the effect of CNO on hypotension-induced activation of

AVP neurons. A1/C1 neurons are required for response of SONAVP neuron to hypotension. Non-

specific lesion of the VLM causes a significant attenuation of this response (Day and Sibbald,

1990; Head et al., 1987). However, no recent studies using the latest neuroscience tools have

demonstrated a specific involvement of A1/C1 neurons in this process. Hypotension was induced

by an injection of Hydralazine (HDZ), a vasodilating drug. A1/C1 inhibition significantly

attenuated the activation of SONAVP neurons in response to HDZ-induced hypotension (Figure 2.

7b, c). Not only that this result validated our experimental system, we were also able to

reevaluate the previous finding with a specific chemogenetic inhibition of A1/C1 neurons.

SONAVP neurons are activated by a wide range of physiological and physical stressors (Antoni,

2017). We tested whether A1/C1 neurons are involved in mediating other stress-induced

responses using fiber photometry. It is known that intraperitoneal (i.p.) injection of

chloride (LiCl), a nausea-inducing agent, causes a large increase in AVP release (O'Connor et al.,

1987). We also observed that SONAVP neurons are strongly activated by aversive stimuli such as physical restraint by the experimenter or pain from acute i.p. injection regardless of the injected

43

Figure 2.7. A1/C1 neurons mediate hypotension-induced activation of SONAVP neurons, but are not required for presystemic regulation. a, Schematic of SONAVP photometry experiment with hM4Di-mediated inhibition of A1/C1 neurons. b, Average SONAVP population activity to hypotension induced by vasodilating drug HDZ in saline and CNO trials. c, Average SONAVP population response to HDZ injection in saline and CNO trials binned every 5 min. ***p < 0.001; RM 2-way ANOVA, n= 9 mice.

AVP d, Heatmap showing single-trial timecourses of SON population activity in response to food bowl placement in saline and CNO trials. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks).

AVP e, Average SON population activity in response to feeding onset in saline and CNO trials. f, Average SONAVP population activity binned across feeding periods. ns, p > 0.05; RM 2-way ANOVA, n= 8 mice. Scale bar, 500µm. Values are means ± SEMs.

44

substances. This is demonstrated by a transient, robust activation of SONAVP neurons

immediately following i.p. injection of HDZ or LiCl (Figure 2. 7b, Supplementary Figure 2.

3a).Therefore, we examined the effect of A1/C1 neuron inhibition on LiCl-induced

(Supplementary Figure 2. 3a and b) and i.p. injection-induced (Supplementary Figure 2. 3c)

activation of SONAVP neurons. Inhibition of A1/C1 neurons had no effect on any of these responses, demonstrating the specificity of the A1/C1 neuron pathway in mediating blood pressure information.

We then used the same approach to examine the involvement of A1/C1 neurons in food- related presystemic regulation. In contrast to the robust blockade of hypotension-induced

response, A1/C1 neuron inhibition had no effect on feeding-induced activation of AVP neurons

(Figure 2. 7d-f). Together, these data indicate that blood pressure information is mediated by a

circuit involving A1/C1 neurons and this circuit is distinct from the food-related presystemic circuit.

2.2.8 Perinuclear zone GABAergic neurons do not mediate food-related presystemic regulation of vasopressin neurons

Next, we explored putative upstream sites identified by rabies mapping conducted at the beginning of our study (Figure 2. 1a-f). We found two additional candidate regions that contained appreciable number of rabies-infected neurons. The first region is the perinuclear zone

(PNZ), which is an area surrounding the SON (Figure 2. 8a). There are number of lesion studies demonstrating involvement of the PNZ in relaying cardiovascular information to SONAVP

neurons (Cunningham et al., 2002; Grindstaff and Cunningham, 2001a, b). Specifically, lesion of

45

Figure 2.8. PNZGABA neurons show presystemic responses to feeding but are not required for food-related presystemic regulation of SONAVP neurons. a, Representative image showing rabies-labeled neurons in the PNZ that are monosynaptically connected to magnocellular SONAVP neurons. b, Number of SONAVP neurons (left) and non-GFP SON neurons (right) receiving direct synaptic inputs from PNZGABA neurons as identified by CRACM in AVP-GFP;Vgat-IRES-Cre. c, Schematic of PNZGABA photometry experiment.

GABA d, Heatmap showing single-trial timecourses of PNZ population activity in response to food bowl placement. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks). n= 8 mice.

GABA e, Average PNZ population activity in response to water/food bowl placement. n= 3 mice. f, Average PNZGABA population activity binned across drinking/feeding periods. *p < 0.05; 2- way ANOVA, n= 3 mice.

46

Figure 2.8. (Continued) g, Schematic of SONAVP photometry experiment with hM4Di-mediated inhibition of PNZGABA neurons.

AVP h, Heatmap showing single-trial timecourses of SON population activity in response to food bowl placement in saline and CNO trials. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks). n= 6 mice.

AVP i, Average of SON population activity in response to feeding onset in saline and CNO trials. n= 6 mice. j, Average SONAVP population activity binned across feeding periods in saline and CNO trials. ns, *p < 0.05; 2-way ANOVA, n= 6 mice. Scale bar, 200µm. Values are means ± SEMs.

47

this area abolishes hypertension-induced inhibition of SONAVP neurons, suggesting an

involvement of GABAergic PNZ (PNZGABA) inputs. We therefore studied the synaptic

connectivity and neural dynamics of PNZGABA neurons using CRACM and fiber photometry.

PNZGABA neurons provided direct inhibitory input to SONAVP neurons (Figure 2. 8b) and were

inhibited selectively by feeding but not drinking (Figure 2. 8c-f), suggesting that the feeding-

induced reduction in inhibitory tone from PNZGABA neurons might underlie feeding-induced

activation of SONAVP neurons. In order to test this idea, we selectively inhibited PNZGABA neurons while recording in vivo calcium dynamics of SONAVP neurons using AVP-IRES-

Cre;Vgat-Flp mice (Figure 2. 8g). Inhibiting PNZGABA neurons had no effect on the feeding- induced presystemic response of SONAVP neurons (Figure 2. 8h-j) demonstrating that PNZGABA inputs are not required for food-related presystemic regulation of SONAVP neurons.

2.2.9 AgRP and POMC neurons do not mediate food-related presystemic regulation of vasopressin neurons

Another candidate region is the arcuate nucleus (ARC) of hypothalamus (Figure 2. 9a). There are a number of anatomical studies suggesting the presence of ARC inputs to magnocellular AVP neurons but no functional studies have been performed to validate this (Douglas et al., 2002;

Leng et al., 1988; Ludwig and Leng, 2000; Pineda et al., 2016). The ARC is an extremely heterogeneous structure. Recent RNA-seq studies in the ARC reported a large number of genetically-distinct cell types, the majority of which are of unknown function (Campbell et al.,

2017). We first decided to focus on the most functionally-relevant neuronal populations in the

ARC: AgRP and POMC neurons. AgRP and POMC neurons have opposite roles in regulating

48

food intake and energy balance. AgRP and POMC neurons are themselves under presystemic

regulation and, notably, their presystemic response is highly tuned to food-predicting cues with

water-predicting cue or drinking having no effect (Zimmerman et al., 2016), which fits nicely

with our hypothesis that food- and water-related information is conveyed by distinct neural circuits. However, it is worth noting that lack of food cue response in SONAVP neurons argues

against the involvement of AgRP and POMC neurons. We still decided to proceed with our

investigation considering the possibility that SON-projecting AgRP and POMC neurons might

behave differently than the entire population, as what we observed in SFOVglut2 neurons. To investigate whether AgRP and POMC neurons are providing direct inputs to PVHAVP and

SONAVP neurons, we performed projection mapping and CRACM. AgRP and POMC projections were examined in the PVH and SON using immunostaining (Supplementary Figure 2. 4a). Dense

AgRP and POMC projections in the PVH were not surprising as the PVH is the main effector site of AgRP and POMC neurons. We also found noticeable amount of projections in the SON.

However, when we performed CRACM using AVP-GFP;AgRP-IRES-Cre mice, no synaptic connection was found between AgRP and PVHAVP and SONAVP neurons (Supplementary Figure

2. 4b). Synaptic connectivity from POMC neurons could not be tested because POMC neurons do not release fast neurotransmitter that can be examined with CRACM. However, it should be noted that lack of fast neurotransmitter release was only validated in the PVH and remains to be tested in other downstream projections of POMC neurons. In line with the lack of direct synaptic connections, inhibition of AgRP or POMC neurons had no effect on the food-related presystemic response of SONAVP neurons when investigated using fiber photometry in AVP-IRES-

Cre;AgRP/POMC-IRES-Cre mice (Figure 2. 9a-d). Therefore, AgRP and POMC neurons were excluded from candidate ARC inputs for food-related presystemic regulation.

49

Figure 2.9. non-AgRP/POMC neurons in the ARC mediates food-related presystemic regulation of SONAVP neurons. a, Representative image showing rabies-labeled neurons in the ARC that are monosynaptically connected to magnocellular SONAVP neurons. b, Schematic of SONAVP photometry experiment with hM4Di-mediated inhibition of AgRP or POMC neurons.

AVP c, Average SON population activity in response to feeding onset in saline and CNO trials of AVP-IRES-Cre;Agrp-IRES-Cre (top) and AVP-IRES-Cre;Pomc-IRES-Cre (bottom) mice.

AVP d, Average SON population activity binned across feeding periods in saline and CNO trials of AVP-IRES-Cre;Agrp-IRES-Cre (top) and AVP-IRES-Cre;Pomc-IRES-Cre (bottom) mice.

50

2.2.10 Food-related presystemic regulation of vasopressin neurons is mediated by unknown neurons in the ARC

Next, we decided to focus on the entire ARC population to investigate their involvement in food-related presystemic regulation of SONAVP neurons. In order to obtain an ARC-specific effect and overcome difficulty of restricting hM4Di expression in the ARC, we created three groups of SONAVP fiber photometry animals with differing hM4Di expression: 1) ARC + VMH

+ DMH (Figure 2. 9e), 2) VMH only (Figure 2. 9i), and 3) DMH only (Figure 2. 9l). This approach allowed full coverage of the ARC with hM4Di for effective silencing, which cannot be achieved with restricted injection into the ARC. All three groups of mice underwent the same experimental paradigm to study the effect of inhibition on feeding-induced activation of SONAVP

neurons. Mice were chronically food-restricted and were habituated to the experimental session

as described above. Inhibition of all three nuclei (ARC+VMH+DMH) caused a significant

attenuation of feeding-induced activation that persisted throughout the recording (Figure 2. 9f-h).

In contrast, inhibition of the VMH (Figure 2. 9j, k) or DMH (Figure 2. 9m, n) alone, fully

sparing the ARC, had no effect on the SONAVP neuron response. Injection of CNO in all three groups had no effect on the baseline activity of SONAVP neurons (not shown). Together, these

results indicate that the ARC contains a cell type, distinct from AgRP or POMC neurons, that

mediates food-related presystemic signals to SONAVP neurons.

51

52

Figure 2.9. (Continued) e, i, l, Schematic of SONAVP photometry experiment with hM4Di-mediated non-specific inhibition of the ARC+VMH+DMH (e), VMH only (i), and DMH only (l).

AVP f, Heatmap showing ingle-trial timecourses of SON population activity in response to food bowl placement in saline and CNO trials of ARC+VMH+DMH group. Trials are sorted according to latency from food bowl placement to feeding onset (black ticks). n= 5 mice.

AVP g, j, m, , Average SON population activity in response to feeding onset in saline and CNO trials of ARC+VMH+DMH (g), VMH only (j), and DMH only (m) groups. n= 5 (ARC+VMH+DMH), 4 (VMH only), 7 (DMH only) mice.

AVP h, k, n, Average SON population activity binned across feeding periods in saline and CNO trials of ARC+VMH+DMH (h), VMH only (k), and DMH only (n) groups. ns, p > 0.05; *p < 0.05; RM 2-way ANOVA, n= 5 (ARC+VMH+DMH), 4 (VMH only), 7 (DMH only) mice. Scale bar, 200µm. Values are means ± SEMs.

53

2.3 Discussion

Discovery of rapid, anticipatory modulation of key hunger and thirst neurons during the

last several years has completely overturned our view of these neurons as regulated exclusively

by homeostatic feedback signals (Augustine et al., 2018; Betley et al., 2015; Beutler et al., 2017;

Chen et al., 2015; Mandelblat-Cerf et al., 2015; Su et al., 2017; Zimmerman et al., 2016).

Surprisingly, a wide array of information is conveyed to these neurons, emphasizing the need to

understand the extended neural circuitry underlying presystemic regulation.

In this study, we described another homeostatic population that is under both presystemic

and systemic regulation: magnocellular AVP neurons. We used a combination of optical and

chemogenetic approaches to investigate how SONAVP neurons develop presystemic responses to different types of stimuli. We showed that presystemic responses of SONAVP neurons to water and food are mediated by two distinct neural circuits. Water-related presystemic information is relayed to SONAVP neurons by the LT. We identified three neural populations in the LT that are involved in different aspects of the water-related presystemic response: MnPO/OVLTVglut2,

SFOVglut2, and MnPO/OVLTVgat neurons. Food-related presystemic information, on the other hand, is relayed by a yet unidentified population of neurons in the ARC. We demonstrated that

key feeding-regulating neurons, i.e. AgRP and POMC neurons, are not involved in this process.

We also validated a circuit involving A1/C1 neurons in the VLM that mediates hypotension-

induced activation of SONAVP neurons. This circuit was also shown to be involved in

hypoglycemia-induced activation as will be discussed in Chapter 3. Collectively, these findings

reveal that neural circuits regulating SONAVP neurons activity, thus, AVP release, are highly

functionally-specialized.

54

2.4 Methods

2.4.1 Ethics

All animal care and experimental procedures were approved in advance by the National

Institute of Health and Beth Israel Deaconess Medical Center Institutional Animal Care and Use

Committee.

2.4.2 Mice

Animals were housed at 22°C–24°C on a 12:12 light/dark cycle (light cycle: 6:00 am to

6:00 pm) with standard mouse chow (Teklad F6 Rodent Diet 8664; Harlan Teklad) and water provided ad libitum, unless specified otherwise. Adult male and female mice (8-16 weeks old) were used for experiments. Mice were maintained on a mixed background. Transgenic mouse strains used: Avp-IRES-Cre (Pei et al., 2014), Th-Cre (Jackson Labs Stock 008601), Dbh-flp

(MMRCC), Gcg-Cre (Jackson Labs Stock 030663), Vglut2-IRES-cre (Jackson Labs Stock

028863), Avp-GFP (MMRCC), Vgat-flp (Jackson Labs Stock 029591), Agrp-IRES-Cre (Jackson

Labs Stock 012899), Pomc-IRES-Cre (Fenselau et al., 2017).

2.4.3 Recombinant adeno-associated viral (rAAV) vectors

The following viral vectors were used in this study: AAV8-FLEX-TVA-mCherry (UNC

Vector Core), AAV1-CA-FLEX-RG (UNC Vector Core), SADΔG–EGFP (EnvA) rabies (Salk

Gene Transfer Targeting and Therapeutics Core), AAV9-EF1α-DIO-ChR2(H134R)-mCherry

55

(Penn Vector Core), AAV8-DIO-synaptophysin-YFP (Virovek, Inc), AAV9-CAG-

ChR2(H134R)-mCherry (Penn Vector Core ), HSV-hEF1a-LS1L-GCaMP6s (Vector Core

Facility at the Massachusetts Institute of Technology), AAV1-hSyn-GCaMP6s (Penn Vector

Core), AAV1-hSyn-EYFP (UNC Vector Core), AAV8-hSyn-hM4Di-mCherry (UNC Vector

Core), AAV8-nEF-fDIO-hM4Di-mCherry (custom-made vector; Boston Children’s Hospital

Viral Core).

2.4.4 Stereotaxic surgery and viral injections

For viral injections, mice were anaesthetized with ketamine/xylazine (100 and 10 mg/kg,

respectively, i.p.) and then placed in a stereotaxic apparatus (David Kopf model 940). A pulled

glass micropipette (20-40 μm diameter tip) was used for stereotaxic injections of adeno-

associated virus (AAV). Virus was injected into the posterior pituitary (100 nl; AP: −3.0 mm;

ML: ±0 mm; DV: -6.0 mm from bregma), SON/PNZ (100 nl/side; AP: −0.65 mm; ML: ±1.25

mm; DV: -5.4 mm from bregma), MnPO/OVLT (10 nl/depth; AP: +0.25 mm; ML: 0 mm; DV: -

4.8, 4.5 mm from bregma), SFO (10 nl; AP: -0.6 mm; ML: 0 mm; DV: -2.6 mm from bregma),

PVH (50 nl/side; AP: -0.75 mm; ML: ±0.3 mm; DV: -4.85 mm from bregma), DMH (50 nl/side;

AP: -1.85 mm; ML: ±0.3 mm; DV: -5.2 mm from bregma), VMH (50 nl/side; AP: -1.6 mm; ML:

±0.4 mm; DV: -5.5 mm from bregma), or Arcuate (50 nl/side; AP: -1.5 mm; ML: ±0.25 mm; DV:

-5.8 mm from bregma) by an air pressure system using picoliter air puffs through a solenoid

valve (Clippard EV 24VDC) pulsed by a Grass S48 stimulator to control injection speed (40

nL/min). The pipette was removed 3 min post-injection followed by wound closure using tissue

adhesive (3M Vetbond). For viral injections into the NTS and VLM, mice were placed into a

56

stereotaxic apparatus with the head angled down at approximately 45°. An incision was made at

the level of the cisterna magna, then skin and muscle were retracted to expose the dura mater

covering the 4th ventricle. A 28-gauge needle was used to make an incision in the dura and allow

access to the NTS and VLM. Virus was then injected into the NTS (10nl/side; AP: -0.2mm; ML:

±0.2mm; DV: -0.2mm from obex) VLM (50nl*2/side; AP: -0.3 and -0.6mm; ML: ±1.3mm; DV:

-1.7mm from obex) as described above. The pipette was removed 3 min post-injection followed by wound closure using absorbable suture for muscle and silk suture for skin. For fibre photometry, an optic fibre (200 µm diameter, NA=0.39, metal ferrule, Thorlabs) was implanted in the MnPO/OVLT, SFO, or SON and secured to the skull with dental cement. Subcutaneous injection of sustained release Meloxicam (4 mg/kg) was provided as postoperative care. The mouse was kept in a warm environment and closely monitored until resuming normal activity

2.4.5 Brain slice electrophysiology and Channelrhodopsin-assisted circuit mapping

(CRACM)

To prepare brain slices for electrophysiological recordings, brains were removed from anesthetized mice and immediately placed in ice-cold cutting solution consisting of (in mM): 72 sucrose, 83 NaCl, 2.5 KCl, 1 NaH2PO4, 26 NaHCO3, 22 glucose, 5 MgCl2, 1 CaCl2, oxygenated with 95% O2 /5% CO2 , measured osmolarity 310-320 mOsm/l. Cutting solution was prepared and used within 72 hours. 250 μm-thick coronal sections containing the PVH and

SON were cut with a vibratome (7000smz2-Campden Instruments) and incubated in oxygenated cutting solution at 34 °C for 25 min. The SON was located by using the bifurcation of the anterior and middle cerebral arteries on the ventral surface of the brain as a landmark. Slices

57

were transferred to oxygenated aCSF (126 mM NaCl, 21.4 mM NaHCO3, 2.5 mM KCl, 1.2 mM

NaH2PO4, 1.2 mM MgCl2, 2.4 mM CaCl2, 10 mM glucose) and stored in the same solution at

room temperature (20-24 °C) for at least 60 min prior to recording. A single slice was placed in the recording chamber where it was continuously super-fused at a rate of 3–4 ml per min with

oxygenated aCSF. Neurons were visualized with an upright microscope equipped with infrared-

differential interference contrast and fluorescence optics. Borosilicate glass microelectrodes (5–

7 MΩ) were filled with internal solution. Whole-cell voltage clamp recordings were obtained using a Cs-based internal solution containing (in mM): 135 CsMeSO3, 10 HEPES, 1 EGTA, 4

MgCl2, 4 Na2-ATP, 0.4 Na2-GTP, 10 Na2-phosphocreatine (pH 7.3; 295 mOsm). To photostimulate ChR2-positive A1/C1 fibres, an LED light source (473 nm) was used. The blue light was focused on to the back aperture of the microscope objective, producing a wide-field exposure around the recorded cell of 1 mW. The light power at the specimen was measured using an optical power meter PM100D (ThorLabs). The light output is controlled by a programmable pulse stimulator, Master-8 (AMPI Co. Israel) and the pClamp 10.2 software

(AXON Instruments). All recordings were made using Multiclamp 700B amplifier, and data was filtered at 2 kHz and digitized at 10 kHz. The photostimulation-evoked EPSC detection protocol consisted of four blue light laser pulses administered 1 s apart during the first 4 s of an 8 s sweep, repeated for a total of 30 sweeps. We attempted to maximize our ability to detect light-evoked currents by biasing our recordings to cell bodies within the densest axon fields. In some experiments, TTX (1 mM) and 4-AP (100 mM) was added to the bath solution in order to confirm monosynaptic connectivity. All CRACM results presented are from 2-3 mice per group.

All analysis was conducted off-line in MATLAB.

58

2.4.6 Fiber photometry experiments and analysis of photometry data

All experiments were conducted in the home-cage in freely moving mice. Beginning

three weeks post-surgery (details above), animals prepared for in vivo fiber photometry

experiments (outlined above), were food restricted to 85-90% of initial body weight. Over this

one-week period, mice were acclimated to chow pellets (500mg, Bio-Serv Dustless Precision

Pellets), and to the food bowl and water bowl used in subsequent photometry experiments. Mice

were habituated to the paradigm for at least 5 days prior to the first recording day. In vivo fiber

photometry was conducted as previously described (Mandelblat-Cerf et al., 2017). A fiber optic

cable (1-m long, metal ferrule, 400 µm diameter; Doric Lenses) was attached to the implanted

optic cannula with zirconia sleeves (Doric Lenses). Laser light (473 nm) was focused on the

opposite end of the fiber optic cable to titrate the light intensity entering the brain to 0.1-0.2 mW.

Emitted light was passed through a dichroic mirror (Di02-R488-25x36, Semrock) and GFP

emission filter (FF03-525/50-25, Semrock), before being focused onto a sensitive photodetector

(Newport part #2151). The GCaMP6 signal was passed through a low-pass filter (50 Hz), and digitized at 1 KHz using a National Instruments data acquisition card and MATLAB software.

The recorded data was exported and then imported into MATLAB for analysis. Fluorescent traces were down-sampled to 1 Hz. We calculated the fractional change in GCaMP6s fluorescence according to the following equation: ΔF/F=(F – F0)/F0, where F is fluorescence measurement and F0 is the mean fluorescence in the 30 sec prior to food/water bowl presentation or feeding onset or 2 min prior to drug injection. In some experiments, to facilitate the comparison across different groups, we normalized the responses of each mouse. In Figure 2. 4f, normalization was performed by (ΔF/F)/FtotalΔF, where FtotalΔF was maximum average change

59

over the entire recording period. In Figure 2. 9e-n, normalization was performed by

(ΔF/F)/Fmax,control, where Fmax,control was maximum average response of control trials within 5 min

of feeding onset.

2.4.7 Brain tissue preparation

Animals were terminally anesthetized with 7% chloral hydrate diluted in saline (350

mg/kg) and transcardially perfused with phosphate- buffered saline (PBS) followed by 10%

neutral buffered formalin (PFA). Brains were removed, stored in the same fixative overnight,

transferred into 20% sucrose at 4 °C overnight, and cut into 40- μm sections on a freezing microtome (Leica) coronally into two equal series.

2.4.8 Immunohistochemistry

Brain sections were washed in PBS with Tween-20, pH 7.4 (PBST) and blocked in 3% normal donkey serum in PBST for 1 h at room temperature. Brain sections were then incubated overnight at room temperature in blocking solution containing primary antiserum (rat anti- mCherry, Life Technologies M11217, 1:1,000; rabbit anti-dsRed, Clontech 632496, 1:1,000; chicken anti-GFP, Life Technologies A10262, 1:1,000; rabbit anti-vasopressin, Sigma-Aldrich

AB1565, 1:1,000; rabbit anti-TH, Millipore AB152, 1:1,000; rabbit anti-POMC precursor,

Phoenix Pharmaceuticals H-029-30, 1:1,000; goat anti-AgRP, Neuromics GT15023, 1:1,000;

chicken anti-GFP, Life Technologies A10262, 1:1,000;). The next morning sections were

extensively washed in PBS and then incubated in Alexa-fluorophore secondary antibody

(1:1,000) for 1 h at room temperature. After several washes in PBS, sections were mounted on

60 gelatin-coated slides and fluorescence images were captured with Olympus VS120 slide scanner microscope.

2.4.9 Statistical analysis

Statistical analyses were performed using SigmaPlot software. Electrophysiological traces were analyzed on Clamp fit 10 (Molecular Devices) and Origin (Origin Lab) software. No statistical method was used to predetermine sample size. Blinding methods were not used. All data presented met the assumptions of the statistical test employed. Specific statistical tests are specified in the figure legends. Animals were excluded from analysis if histological validation revealed poor or inaccurate reporter expression or inaccurate fiber placement unless otherwise noted. n values reflect the final number of validated animals per group included in the analysis.

61

References

Abbott, S.B., Machado, N.L., Geerling, J.C., and Saper, C.B. (2016). Reciprocal Control of Drinking Behavior by Median Preoptic Neurons in Mice. The Journal of neuroscience : the official journal of the Society for Neuroscience 36, 8228-8237.

Antoni, F.A. (2017). Vasopressin as a Stress Hormone. Hbk Stress Ser 2, 97-108.

Augustine, V., Gokce, S.K., Lee, S., Wang, B., Davidson, T.J., Reimann, F., Gribble, F., Deisseroth, K., Lois, C., and Oka, Y. (2018). Hierarchical neural architecture underlying thirst regulation. Nature 555, 204-209.

Betley, J.N., Cao, Z.F., Ritola, K.D., and Sternson, S.M. (2013). Parallel, redundant circuit organization for homeostatic control of feeding behavior. Cell 155, 1337-1350.

Betley, J.N., Xu, S., Cao, Z.F., Gong, R., Magnus, C.J., Yu, Y., and Sternson, S.M. (2015). Neurons for hunger and thirst transmit a negative-valence teaching signal. Nature.

Beutler, L.R., Chen, Y.M., Ahn, J.S., Lin, Y.C., Essner, R.A., and Knight, Z.A. (2017). Dynamics of Gut-Brain Communication Underlying Hunger. Neuron 96, 461-+.

Bolignano, D., Cabassi, A., Fiaccadori, E., Ghigo, E., Pasquali, R., Peracino, A., Peri, A., Plebani, M., Santoro, A., Settanni, F., et al. (2014). Copeptin (CTproAVP), a new tool for understanding the role of vasopressin in pathophysiology. Clinical chemistry and laboratory medicine : CCLM / FESCC 52, 1447-1456.

Campbell, J.N., Macosko, E.Z., Fenselau, H., Pers, T.H., Lyubetskaya, A., Tenen, D., Goldman, M., Verstegen, A.M., Resch, J.M., McCarroll, S.A., et al. (2017). A molecular census of arcuate hypothalamus and median eminence cell types. Nature neuroscience 20, 484-496.

Card, J.P., Sved, J.C., Craig, B., Raizada, M., Vazquez, J., and Sved, A.F. (2006). Efferent projections of rat rostroventrolateral medulla C1 catecholamine neurons: Implications for the central control of cardiovascular regulation. The Journal of comparative neurology 499, 840-859.

Chen, Y., Lin, Y.C., Kuo, T.W., and Knight, Z.A. (2015). Sensory detection of food rapidly modulates arcuate feeding circuits. Cell 160, 829-841.

Cunningham, J.T., Bruno, S.B., Grindstaff, R.R., Grindstaff, R.J., Higgs, K.H., Mazzella, D., and Sullivan, M.J. (2002). Cardiovascular regulation of supraoptic vasopressin neurons. Progress in brain research 139, 257-273.

Day, T.A., and Sibbald, J.R. (1990). Involvement of the A1 cell group in baroreceptor inhibition of neurosecretory vasopressin cells. Neuroscience letters 113, 156-162.

62

Douglas, A.J., Bicknell, R.J., Leng, G., Russell, J.A., and Meddle, S.L. (2002). Beta-endorphin cells in the arcuate nucleus: projections to the supraoptic nucleus and changes in expression during pregnancy and parturition. Journal of neuroendocrinology 14, 768-777.

Ferguson, J.W., Therapondos, G., Newby, D.E., and Hayes, P.C. (2003). Therapeutic role of vasopressin receptor antagonism in patients with liver cirrhosis. Clinical science 105, 1-8.

Grindstaff, R.R., and Cunningham, J.T. (2001a). Cardiovascular regulation of vasopressin neurons in the supraoptic nucleus. Exp Neurol 171, 219-226.

Grindstaff, R.R., and Cunningham, J.T. (2001b). Lesion of the perinuclear zone attenuates cardiac sensitivity of vasopressinergic supraoptic neurons. American journal of physiology Regulatory, integrative and comparative physiology 280, R630-638.

Head, G.A., Quail, A.W., and Woods, R.L. (1987). Lesions of A1 noradrenergic cells affect AVP release and heart rate during hemorrhage. The American journal of physiology 253, H1012-1017.

Kohl, J., Babayan, B.M., Rubinstein, N.D., Autry, A.E., Marin-Rodriguez, B., Kapoor, V., Miyamishi, K., Zweifel, L.S., Luo, L., Uchida, N., et al. (2018). Functional circuit architecture underlying parental behaviour. Nature 556, 326-331.

Leng, G., Yamashita, H., Dyball, R.E., and Bunting, R. (1988). Electrophysiological evidence for a projection from the arcuate nucleus to the supraoptic nucleus. Neuroscience letters 89, 146- 151.

Livneh, Y., Ramesh, R.N., Burgess, C.R., Levandowski, K.M., Madara, J.C., Fenselau, H., Goldey, G.J., Diaz, V.E., Jikomes, N., Resch, J.M., et al. (2017). Homeostatic circuits selectively gate food cue responses in insular cortex. Nature 546, 611-616.

Livneh, Y., Sugden, A.U., Madara, J.C., Essner, R.A., Flores, V.I., Sugden, L.A., Resch, J.M., Lowell, B.B., and Andermann, M.L. (2020). Estimation of Current and Future Physiological States in Insular Cortex. Neuron.

Ludwig, M., and Leng, G. (2000). GABAergic projection from the arcuate nucleus to the supraoptic nucleus in the rat. Neuroscience letters 281, 195-197.

Mandelblat-Cerf, Y., Kim, A., Burgess, C.R., Subramanian, S., Tannous, B.A., Lowell, B.B., and Andermann, M.L. (2017). Bidirectional Anticipation of Future Osmotic Challenges by Vasopressin Neurons. Neuron 93, 57-65.

Mandelblat-Cerf, Y., Ramesh, R.N., Burgess, C.R., Patella, P., Yang, Z., Lowell, B.B., and Andermann, M.L. (2015). Arcuate hypothalamic AgRP and putative POMC neurons show opposite changes in spiking across multiple timescales. eLife 4.

McKinley, M.J., Mathai, M.L., McAllen, R.M., McClear, R.C., Miselis, R.R., Pennington, G.L., Vivas, L., Wade, J.D., and Oldfield, B.J. (2004). Vasopressin secretion: osmotic and hormonal regulation by the lamina terminalis. Journal of neuroendocrinology 16, 340-347.

63

Nielsen, S., Chou, C.L., Marples, D., Christensen, E.I., Kishore, B.K., and Knepper, M.A. (1995). Vasopressin increases water permeability of kidney collecting duct by inducing translocation of aquaporin-CD water channels to plasma membrane. Proceedings of the National Academy of Sciences of the United States of America 92, 1013-1017.

O'Connor, E.F., Cheng, S.W., and North, W.G. (1987). Effects of intraperitoneal injection of lithium chloride on neurohypophyseal activity: implications for behavioral studies. Physiology & behavior 40, 91-95.

Oka, Y., Ye, M., and Zuker, C.S. (2015). Thirst driving and suppressing signals encoded by distinct neural populations in the brain. Nature.

Pineda, R., Sabatier, N., Ludwig, M., Millar, R.P., and Leng, G. (2016). A Direct Neurokinin B Projection from the Arcuate Nucleus Regulates Magnocellular Vasopressin Cells of the Supraoptic Nucleus. Journal of neuroendocrinology 28.

Stricker, E.M., and Hoffmann, M.L. (2007). Presystemic signals in the control of thirst, salt appetite, and vasopressin secretion. Physiology & behavior 91, 404-412.

Stricker, E.M., and Stricker, M.L. (2011). Pre-systemic controls of fluid intake and vasopressin secretion. Physiology & behavior 103, 86-88.

Su, Z.W., Alhadeff, A.L., and Betley, J.N. (2017). Nutritive, Post-ingestive Signals Are the Primary Regulators of AgRP Neuron Activity. Cell reports 21, 2724-2736.

Zimmerman, C.A., Lin, Y.C., Leib, D.E., Guo, L., Huey, E.L., Daly, G.E., Chen, Y., and Knight, Z.A. (2016). Thirst neurons anticipate the homeostatic consequences of eating and drinking. Nature 537, 680-684.

64

Chapter 3

Regulation of Vasopressin Secretion by Hypoglycemia

65

3.1 Introduction

Glucose is a primary fuel source for neurons. The brain uses nearly 60% of glucose

consumed by the body and normal brain function critically depends on the supply of blood

glucose (Berg, 2002; Mergenthaler et al., 2013). Hypoglycemia leads to a decline in cognitive

function and, when severe, can cause coma and even death (Cryer, 2007). In order to prevent

deleterious declines in blood glucose, the body reacts to hypoglycaemia by triggering a multitude of counter-regulatory responses. The first line of defence against hypoglycemia is the release of

glucagon (Cryer, 1994). Glucagon, a hormone released by alpha-cells of the pancreatic islets,

stimulates hepatic glucose production to maintain euglycemia.

Tight regulation of glucagon secretion is crucial for glucose homeostasis. Hyperglycemia

in type 1 (T1D) and type 2 (T2D) diabetes mellitus is attributed to a combination of excess

glucagon and insufficient insulin (Unger and Cherrington, 2012). In addition, impaired counter-

regulatory glucagon secretion in diabetic patients is a major cause of insulin treatment-induced

hypoglycemic attacks (McCrimmon and Sherwin, 2010; Siafarikas et al., 2012; Sprague and

Arbelaez, 2011). Despite its central role in glucose homeostasis and the pathophysiology of

diabetes, controversies still remain over how glucagon secretion is regulated (Gylfe, 2013; Lai et

al., 2018). While studies in isolated pancreatic islets support intrinsic (Basco et al., 2018;

Rorsman et al., 2014; Yu et al., 2019) and paracrine (Briant et al., 2018; Vergari et al., 2019)

mechanisms that exist within alpha-cells and between multiple islet cell types, respectively, it is

becoming evident that central regulation plays an equally important role. Glucose-sensitive

neurons have been identified in multiple brain regions (Burdakov et al., 2005; Garfield et al.,

2014; Labouebe et al., 2016; Lamy et al., 2014; Meek et al., 2016; Wang et al., 2004).

66

Manipulation of some of these neurons has been shown to significantly alter pancreatic islet function and blood glucose. These effects are thought to be mediated by autonomic innervation of the pancreas (Lamy et al., 2014; Marty et al., 2005; Thorens, 2014).

Another line of evidence suggests that AVP, released by magnocellular AVP neurons of the hypothalamus, is involved in glucose homeostasis. Hypoglycemia is a potent stimulus for

AVP release (Baylis and Heath, 1977; Baylis and Robertson, 1980; Fisher et al., 1987), and

intraperitoneal (i.p) injection or intravenous (i.v.) infusion of AVP causes a transient increase in

blood glucose (Hems et al., 1975; Spruce et al., 1985; Taveau et al., 2017). Elevation of AVP by

hyperosmotic stimuli potentiates glucagon secretion to hypoglycaemia (Adler and Majzoub,

1987). Moreover, a recent RNA-seq study demonstrated that artificial induction of

hyperglucagonemia by a glucagon receptor antibody resulted in a significant upregulation of V1b

subtype of AVP receptor (V1bR) mRNA expression in pancreatic alpha-cells (Sloop et al.,

2004). Together, these data suggest that hypoglycemia-induced AVP release plays a role in

counter-regulatory glucagon release.

To test this hypothesis, we employed a wide range of in vitro and in vivo techniques to investigate the neural and molecular pathways mediating glucoregulatory effect of AVP.

67

3.2 Results

3.2.1 Glucagon secretion in response to an insulin tolerance test is driven by AVP

Hypoglycemia induced by an insulin tolerance test (ITT), a bolus i.p. injection of insulin, robustly stimulates glucagon secretion in vivo (Figure 3. 1a, b). However, reducing the glucose concentration from 8 to 4 mM (mimicking an ITT) does not stimulate glucagon secretion from isolated (ex vivo) islets (Figure 3. 1c). Therefore, mechanisms extrinsic to the islet must control glucagon secretion in vivo, at least during an ITT. We hypothesised that this extrinsic stimulus is

AVP. First, we investigated whether AVP neuron activity is increased in response to an ITT in vivo using fiber photometry (Figure 3. 1d-h). The increase in circulating insulin evoked an increase in SONAVP neuron activity (Figure 3. 1f-h). Saline vehicle treatment did not increase

SONAVP neuron population activity (Figure 3. 1f-h). Finally, we confirmed that plasma AVP is elevated following an ITT by measuring its surrogate marker, copeptin (Figure 3. 1i).

To investigate the kinetics and glucose-dependency of this response, we simultaneously recorded SONAVP neuron activity and plasma glucose (by continuous glucose monitoring; Figure

3. 2a, b). This revealed that AVP neuron activity initiates at 4.9 ±0.4 mM glucose (Figure 3. 2c) at a delay of ~10 mins (Figure 3. 2d). Once the threshold has been reached, SONAVP neuron displayed random bursts of activity while hypoglycemia is maintained. The amount of SONAVP

neuron activation did not correlate with the severity of hypoglycemia. Neither frequency nor

peak amplitude correlated with the blood glucose change. Insulin-induced hypoglycemia was

accompanied by a significant decrease in the body temperature (Figure 3. 2e), as previously

reported (Buchanan et al., 1991; Freinkel et al., 1972).

68

Figure 3.1. Insulin-induced hypoglycaemia enhances population activity of AVP neurons in the SON, driving glucagon secretion a. Insulin tolerance test (ITT; 0.75 U/kg) in n=5 wild-type mice. Plasma glucose falls from ~8 mM to ~4 mM in 30 mins. b. Plasma glucagon prior to, and 30 minutes following an ITT. n=5 wild-type mice. Paired t-test, p<0.01=**, c. Glucagon secretion from isolated (ex vivo) mouse islets in response to different concentrations of glucose. n=7 wild-type mice. One-way ANOVA with Tukey post-hoc. 4 mM vs 8 mM glucose, p=0.99.

69

Figure 3.1. (Continued) d. In vivo fibre photometry measurements of population GCaMP6s activity in pituitary- projecting AVP neurons in the supraoptic nucleus (SON). AAV-DIO-GCaMP6s was injected into the SON of Avpires-Cre+ mice. GCaMP6s was then imaged as indicated by the protocol in the horizontal bar. e. Expression of GCaMP6s in AVP neurons in the SON. The tip of the optic fibre can be seen (arrow). f. Raw population GCaMP6s activity (F) recording following an insulin tolerance test (ITT; 1.5 U/kg) and saline control. Baseline fluorescence (F0) was calculated from signal 3 minutes prior to ITT/saline. Recordings contained an artefact from the injection (arrow) which was removed from plotted time-series and analysis. g. Upper panel: heatmap of population activity (GCaMP6s) response to ITT for each mouse (n=6). Data represented as % of maximum signal. Lower panel: heatmap of population activity (GCaMP6s) response to saline vehicle for each mouse (n=6). h. GCaMP6s signal (normalized) in response to insulin (n=6) or vehicle (n=6). GCaMP6s data represented as (F-F0)/F0. i. Plasma copeptin at 30 minutes following either saline or insulin injection. Mann-Whitney U test, p=0.021. n=15 (saline) and n=18 (insulin).

To determine whether AVP contributes to glucagon secretion during an ITT, we injected

wild-type mice with the V1bR antagonist SSR149415 (nelivaptan; Serradeil-Le Gal et al. (2002))

prior to an ITT (Figure 3. 3a-b). This reduced insulin-induced glucagon secretion by 70%.

Similarly, in Avpr1b knockout mice (Avpr1b-/-; Lolait et al. (2007)) glucagon secretion was decreased by 65% compared to wild-type littermates (Avpr1b-/-; Figure 3. 3c-d). There was not a change in the depth of hypoglycemia induced following V1bR antagonism (Figure 3. 3b) or in

Avpr1b-/- mice (Figure 3. 3d), despite the drastic reduction in plasma glucagon. Thus, it appears that the supraphysiological concentrations of insulin likely dominate the hyperglycemic action of endogenous glucagon. This ‘glucagonostatic’ effect of Avpr1b antagonism or KO was also observed during deeper hypoglycemia, achieved with a higher dose of insulin (Figure 3. 3e-h).

We note that this glucose phenotype is in keeping with that originally reported phenotype

in the Avpr1b knockout mouse (see Lolait et al. (2007); cf. Fujiwara et al. (2007)). AVP can also

increase circulating corticosterone (Antoni, 1993) and adrenaline (Grazzini et al., 1996).

70

71

Figure 3.2. Simultaneous continuous glucose monitoring (CGM) and in vivo fibre photometry of AVP neurons a. Simultaneous continuous glucose monitoring (CGM) and in vivo fibre photometry of AVP neurons (in Avpires-Cre+ mice; as per Figure 3. 1d) in response to an ITT (1.5 U/kg; 3 trials; red) and saline (blue), in two mice. Dashed line indicates the time of injection (which is accompanied with an artefact in the GCaMP6 signal). b. Zoom of each insulin trial from a). Dashed line indicates time of injection. In one trial, the glucose concentration at which the GCaMP6 signal first exhibited a peak (solid line), and first exceeded 2 SD (dashed-dot) and 3 SD (dot) of the baseline signal, are indicated. The 2-3 min period of baseline recording prior to the insulin injection is also depicted. c. This analysis was conducted for each trial. Grouped analysis as depicted in b) of the glucose concentration at which the GCaMP6 signal crosses >2 SD from baseline, crosses > 3 SD from baseline and first exhibits a peak, following an insulin tolerance test. n=2 mice, 6 trials. One- way ANOVA, p<0.05=*. d. Grouped analysis as depicted in b) of the delay (from insulin injection) taken for the GCaMP6 signal to cross >2 SD from baseline, cross > 3 SD from baseline and first exhibits a peak. n=2 mice, 6 trials. One-way ANOVA, p<0.05=*. e. Body temperature following insulin injection (dashed line). Mean ± SEM of six trials in n=2 mice. Dotted line indicates (average) time GCaMP6 activity first exhibits a peak, as per d).

However, the reduction in glucagon following V1bR blockade was due to a direct

antagonism of the action of AVP on alpha-cells. This is suggested by the observation that when tested at concentrations measured in plasma following insulin-induced hypoglycemia (see Lolait et al. (2007) and Christensen (1974)), neither adrenaline (Supplementary Figure 3. 1a-c) nor corticosterone (Supplementary Figure 3. 1d and e) increased glucagon output or alpha-cell activity in isolated mouse or human islets.

We next investigated the role of AVP in controlling glucagon secretion evoked by 2- deoxy-D-glucose (2DG). 2DG is a non-metabolizable glucose molecule that evokes a state of perceived glucose deficit, mimicking hypoglycemia; the resultant glucopenia triggers a robust counter-regulatory response, including glucagon secretion (Marty et al., 2005). We postulated that AVP contributes to this response. We monitored SONAVP neuron activity in response to

72

73

Figure 3.3. The vasopressin 1b receptor mediates insulin-induced glucagon secretion. a) Plasma glucagon following an ITT (0.5 U/kg; injection at 30 min). 30 min prior to the ITT (at 0 min), either the V1bR antagonist SSR149415 (30 mg/kg) or vehicle was administered i.p. Two-way RM ANOVA with Sidak’s (between conditions) and Tukey’s (within condition) multiple comparisons test. Vehicle vs. SSR149415; p<0.01=†† (60 minutes) and p=0.08 (90 minutes). Within the condition, SSR149415 did not increase glucagon secretion (all p>0.6 for 30 vs 60 and 90 min), whereas glucagon was increased in the vehicle treatment group (all p<0.02 for 30 vs 60 and 90 min). Time, p<0.0001; Treatment, p=0.045; Interaction, p=0.016. n=9 wild-type mice. b) Blood glucose for cohort in (a). Two-way RM ANOVA. Time, p<0.0001; Treatment, p=0.40; Interaction, p=0.51. n=9 wild-type mice. c) Same as a) but for a larger dose of insulin (0.75 U/kg). d) Same as b) but for a larger dose of insulin (0.75 U/kg). e) Plasma glucagon following an ITT (injection at 0 min) in Avpr1b-/- mice and littermate controls (Avpr1b+/+). Two-way RM ANOVA with Sidak’s multiple comparisons test. Avpr1b-/- vs. Avpr1b+/+; p<0.001=††† (30 minutes). 0 vs. 30 minutes; p<0.001 (Avpr1b+/+); p=0.097 (Avpr1b-/-). Time, p<0.0001; Genotype, p=0.008; Interaction, p=0.009. n=8-9 mice. Note that even after removal of the two high plasma glucagon samples at 30 minutes in Avpr1b+/+ mice, the glucagon is still higher than in Avpr1b-/- mice (p=0.0028). f) Blood glucose for cohort in (c). Two-way RM ANOVA. Time, p<0.0001; Treatment, p=0.40; Interaction, p=0.51. n=8-9 mice. g) Same as e) but for a larger dose of insulin (0.75 U/kg). h) Same as f) but for a larger dose of insulin (0.75 U/kg).

2DG by in vivo fiber photometry (as per Figure 3. 1d), and correlated this to changes in plasma glucose during this metabolic challenge (Supplementary Figure 3. 2). Injection (i.p.) of 2DG increased blood glucose (Supplementary Figure 3. 2a) and triggered a concomitant elevation of

GCaMP6s signal in SONAVP neuron (Supplementary Figure 3. 2b-d). The hyperglycemic response to 2DG was suppressed by pre-treatment with either the V1bR antagonist SSR149415 or the glucagon receptor antagonist LY2409021 (Supplementary Figure 3. 2e; Kazda et al.

(2016)). Finally, the elevation in plasma glucagon by 2DG injection was reduced following pretreatment with the V1bR antagonist SSR149415 (Supplementary Figure 3. 2f). Therefore, we conclude that AVP contributes to the hyperglycemic response to 2DG by stimulating glucagon release.

74

3.2.2 AVP evokes hyperglycemia and hyperglucagonemia

Next, we investigated the mechanisms by which AVP stimulates glucagon during

hypoglycemia. We first investigated the metabolic effects of AVP in vivo (Figure 3. 4a-d). We

expressed the modified human M3 muscarinic receptor hM3Dq (see Alexander et al. (2009)) in

SONAVP neuron by bilaterally injecting a Cre-dependent virus containing hM3Dq (AAV-DIO-

hM3Dq-mCherry) into the SON of mice bearing Avp promoter-driven Cre recombinase (Avpires-

Cre+ mice; Figure 3. 4a). Expression of hM3Dq was limited to the SON (Supplementary Figure 3.

3), thus allowing targeted activation of magnocellular AVP neurons (which determine circulating

AVP). Patch-clamp recordings from brain slices of the SON prepared from these mice revealed

that bath application of clozapine-N-oxide (CNO; 5-10 µM) – a specific, pharmacologically inert agonist for hM3Dq - induced membrane depolarisation and increased the firing rate in hM3Dq-

expressing SONAVP neuron (Supplementary Figure 3. 3a and b). Injection of CNO (3 mg/kg i.p.)

in vivo resulted in an increase in blood glucose (Figure 3. 4b), as well as copeptin (Figure 3. 4c).

To understand the contribution of glucagon to this hyperglycemic response, we pre-treated mice

with the glucagon receptor antagonist LY2409021. This completely abolished the hyperglycemic

action of CNO (Figure 3. 4b). Similarly, to understand the contribution of vasopressin 1b

receptor (V1bR) signalling, we pre-treated mice with the V1bR antagonist SSR149415. This also

abolished the hyperglycemic effect of CNO (Figure 3. 4b). Measurements of plasma glucagon

during CNO treatment revealed it was elevated by ~30% (Figure 3. 4d). We note that this

response was quite small (40% increase in glucagon compared to saline injection), likely

reflecting the variable expression of hM3Dq. CNO did not change food intake (Supplementary

75

Figure 3.4. Metabolic response to elevating AVP in vivo a. AAV-DIO-hM3Dq-mCherry was injected bilaterally into the supraoptic nucleus (SON) of Avpires-Cre+ mice. Mice were fasted for 4 hours (beginning at 10:00 am), and then CNO (or saline vehicle) was injected (3 mg/kg i.p.). In the same cohort, during a different trial, the glucagon receptor antagonist LY2409021 (5 mg/kg) or V1bR antagonist SSR149415 (30 mg/kg) was injected i.p. 30 minutes prior to CNO. The lower panel gives an overview of the protocol. Plasma glucose and glucagon were then measured. See also Supplementary Figure 3. 3. b. Change in blood glucose from 0 min sample. Two-way RM ANOVA with Tukey’s multiple comparison test. CNO 0 mins vs. CNO at 15, 30 and 60 mins; p<0.05=*, p<0.01=**. 30 mins comparison of CNO vs. Saline, CNO + LY2409021 or CNO + SSR149415 at 30 mins; p<0.01=††. Time, p<0.0001; Treatment, p=0.0006; Interaction, p<0.0001. N=6 mice. c. As in (b), but plasma copeptin 30 minutes following saline (-) or CNO (+). N=12 mice. Mann Whitney t-test (p=0.0025). N=18 mice. Removal of < 1 pg/ml measurements (N=3 in saline treatment) yields p=0.007. d. As in (b), but plasma glucagon at 30 mins following saline (-) or CNO (+) injection. Represented as % of baseline glucagon (0 min). N=11 mice. Mann-Whitney U test, p=0.03=*. e. Islets from GcgCre+-GCaMP3 mice were injected into the anterior chamber of the eye (ACE) of recipient mice (n=5 islets in 5 mice). After >4 weeks, GCaMP3 was imaged in vivo in response to i.v. AVP (10 µg/kg) or saline administration. Saline did not change the GCaMP3 signal in all mice tested. AVP evoked an increase in calcium activity, typically starting with a large transient. f. Response of alpha-cell to i.v. AVP. Lower panel shows raster plot of response in different cells. g. Integrated F/F0 (area under curve) response for all alpha-cells in recorded islets (5 islets, N=3 mice). The area under the curve was calculated 30 secs before i.v. injection, 30 secs after and 120 secs after. One-way RM ANOVA with Tukey’s multiple comparison test; p<0.01=**. h. Image of islets (arrows) engrafted in the ACE. 76

Figure 3. 3f) and did not have an off-target effect on blood glucose in Avpires-Cre+ mice expressing

a passive protein (mCherry) under AAV transfection in the SON (Supplementary Figure 3. 3g).

Injection of (exogenous) AVP also caused an increase in glucose and glucagon relative to saline

injection in wild-type mice (Supplementary Figure 3. 3h-j). In summary, CNO stimulates AVP

release, which in turn elevates plasma glucose through stimulation of glucagon release.

To determine whether the hyperglycemic and hyperglucagonemic actions of AVP are due

to AVP directly stimulating glucagon secretion from alpha-cells, we assessed the response of islets to AVP. First, we assessed this response in vivo; we isolated islets from GcgCre+ mice

crossed with a Cre-dependent GCaMP3 reporter mouse (from hereon, Gcg-GCaMP3 mice) and

implanted them in the anterior chamber of the eye (Figure 3. 4e-h; see Salem et al. (2019)). This

enabled alpha-cell Ca2+ to be imaged in vivo, and doing so revealed that administration (i.v.) of

AVP results in an increase in alpha-cell activity (Figure 3. 4f-g).

To understand whether AVP stimulates glucagon secretion directly from islets, we made

a detailed assessment of the response of isolated (ex vivo) mouse islets and the in situ perfused

pancreas to AVP. First, we characterised the expression of vasopressin receptors in mouse islets.

V1bR mRNA (Avpr1b) was expressed in whole mouse islets, whereas vasopressin receptor

subtypes 1a and 2 mRNA were not; instead, they were detected in various extra-pancreatic

tissue, consistent with their distinct roles in the regulation of blood pressure and diuresis

(Bourque, 2008), respectively (Supplementary Figure 3. 4a). To determine whether Avpr1b

expression was enriched in alpha-cells, mice bearing a proglucagon promoter-driven Cre-

recombinase (GcgCre+ mice) were crossed with mice expressing a Cre-driven fluorescent reporter

(RFP). qPCR of the fluorescence-activated cell sorted RFP+ and RFP- fractions revealed

77

that Avpr1b is upregulated in mouse alpha-cells with ~43-fold enrichment (Supplementary

Figure 3. 4c and d).

AVP dose-dependently stimulated glucagon secretion from ex vivo mouse islets

(Supplementary Figure 3. 5). Similar stimulatory effects of AVP were observed in the in situ

perfused mouse pancreas (Supplementary Figure 3. 5d). In mouse islets, the AVP-induced increase in glucagon was prevented by SSR149415 (Supplementary Figure 3. 5c). AVP did not affect insulin secretion (Supplementary Figure 3. 5e-h). To understand the intracellular

mechanisms by which AVP stimulates glucagon secretion, we performed perforated patch-clamp recordings of electrical activity and Ca2+ imaging in alpha-cells in isolated islets from GcgCre+

mice crossed with a Cre-dependent GCaMP3 reporter mouse (from hereon, Gcg-GCaMP3 mice).

AVP increased action potential firing (Supplementary Figure 3. 5i, j) and Ca2+ oscillations

(Supplementary Figure 3. 5k, l) in a dose-dependent manner. The Ca2+ response was abolished

following application of the V1bR antagonist SSR149415 (Supplementary Figure 3. 5m, n).

2+ AVP-induced Ca activity was dependent on Gq-protein activation, because it was blocked with

the Gq-inhibitor YM254890 (Takasaki et al. (2004); Supplementary Figure 3. 6a, b) and it

increased intracellular diacylglycerol (Supplementary Figure 3. 6c, d). In line with Gq activation,

electrical activity evoked by AVP exhibited membrane potential oscillations (as revealed by

power-spectrum analysis; Supplementary Figure 3. 6e). Conducting mathematical modelling of the intracellular signalling pathway following Gq receptor activation demonstrated that the

2+ canonical Gq pathway can evoke Ca oscillations, further supporting the involvement of VlbR, a

Gq protein-coupled receptor, in stimulating glucagon secretion from alpha cells (Supplementary

Figure 3. 6f, g).

78

3.2.3 Hypoglycemia evokes AVP secretion via activation of A1/C1 neurons

Many physiological stressors activate hindbrain catecholamine neurons, which release

noradrenaline (A1) or adrenaline (C1) and reside in the ventrolateral portion of the medulla

oblongata (VLM). Activation of C1 neurons by targeted glucoprivation or chemogenetic

manipulation is known to elevate blood glucose (Li et al., 2018; Ritter et al., 2000; Zhao et al.,

2017) and plasma glucagon (Andrew et al., 2007). Furthermore, C1 cell lesions severely

attenuate the release of AVP in response to hydralazine-evoked hypotension (Madden et al.,

2006), indicating that this hindbrain site may be a key regulator of magnocellular AVP neuron

activity during physiological stress.

To characterize any functional connectivity between A1/C1 neurons and SONAVP neurons, we conducted channelrhodpsin-2-assisted circuit mapping (CRACM; Petreanu et al. (2007)) and

viral tracer studies (Figure 3. 5a-e). We injected a Cre-dependent viral vector containing the

light-gated ion channel channelrhodopsin-2 (AAV-DIO-ChR2-mCherry) into the VLM

(targeting A1/C1 neurons) of AvpGFP x ThCre+ mice (Figure 3. 5a). Projections from the A1/C1 neurons were present in the SON and PVH and co-localized with AVP-immunoreactive neurons

(Figure 3. 5b and Supplementary Figure 3. 7a). Brain slice electrophysiology revealed that in the

majority (89%) of GFP+ SONAVP neurons, opto-activation of A1/C1 neuron terminals results in excitatory post-synaptic currents (EPSCs; Figure 3. 5c, d). These EPSCs were glutamatergic, as they were abolished by the AMPA and kainate receptor antagonist DNQX (Figure 3. 5c).

Furthermore, these EPSCs could be blocked with TTX, but reinstated with addition of 4-AP

(Figure 3. 5e), indicating that A1/C1 neurons connect to SONAVP neurons in the SON monosynaptically.

79

Figure 3.5. Insulin-induced AVP secretion is mediated by A1/C1 neurons a. Upper panel: AAV-DIO-ChR2-mCherry was injected into the VLM of ThCre+ mice or AvpGFP x ThCre+ mice, targeting A1/C1 neurons. Brains were then prepared for viral projection mapping or channelrhodopsin2-assisted circuit mapping (CRACM), respectively. Lower panel: CRACM. Excitatory post-synaptic currents (EPSCs) were recorded in voltage-clamp mode in GFP+ (AVP) and GFP- neurons in the SON and PVH. The number (n) of neurons that

80

Figure 3.5. (Continued) responded to opto-activating A1/C1 terminals. Total of 36 neurons recorded. See also Supplementary Figure 3. 7a. b. Upper panel: Viral expression of ChR2-mCherry in A1/C1 neurons. Lower panel: A1/C1 neuron terminals co-localise with AVP-immunoreactive neurons. c. Left panel: EPSCs evoked by opto-activation of A1/C1 terminals with 473 nm light pulses (arrows). Right panel: Light-evoked EPSCs following application of DNQX (20 µM). d. EPSC waveforms in a single GFP+ (AVP) neuron in response to repeated opto-activation of A1/C1 neuron terminals. Mean and SD of all EPSCs shown in black line and shaded area. Timing of light pulse shown in blue bar. e. EPSCs evoked by opto-activating A1/C1 terminals at baseline (left) and following addition of TTX (1 µM; middle) and 4-AP (1 mM; right). f. AAV-DIO-hM3Dq was injected into ThCre+ mice, targeting A1/C1 neurons. CNO (1 mg/kg) was then injected (i.p.). Antagonists (or vehicle) for the V1bR (SSR149415, 30 mg/kg) or glucagon receptor (GCGR; LY2409021, 5 mg/kg) were injected 30 minutes prior to CNO. Plasma glucose and glucagon was then measured. See also Supplementary Figure 3. 7b,c. g. Plasma glucose in response to CNO and pre-treatment with antagonists. N=8 mice. Two-way RM ANOVA (Sidak’s multiple comparison’s test).. Time (p<0.0001), Treatment (p=0.03) and Interaction (p=0.0002). h. Same as (g), but 30 minute data only. i. Plasma glucagon at 30 minutes post CNO (or vehicle) injection. Two-way RM ANOVA with Tukey’s (within treatment) and Sidak’s (between treatments) multiple comparisons. p<0.05=*, ns=not significant. Within treatment, CNO increased glucagon at 30 min vs. 0 min (p=0.022). Saline did not (p=0.96). Between treatments, CNO increased glucagon at 30 min vs. saline (p=0.001). N=6 mice. All data are represented as mean ± SEM. j. In vivo fiber photometry measurements of population GCaMP6 activity in pituitary-projecting SON AVP neurons during A1/C1 neuron inhibition. AAV-DIO-GCaMP6s was injected into the SON and AAV-fDIO-hM4Di-mCherry into the VLM of Avpires-Cre+ x Dbhflp+ mice. GCaMP6s was then imaged in response to an insulin tolerance test (ITT), following inhibition of the A1/C1 neuron (with CNO at 1 mg/kg), as indicated by the protocol in the lower horizontal bar. See Supplementary Figure 3. 8b. k. Example population activity in one mouse in response to an ITT, following saline or CNO treatment (on different trials). Baseline fluorescence (F0) was calculated from signal 3 minutes prior to ITT, and subtracted from the fluorescence signal (F). l. Left panel: Average GCaMP6 signal (normalized) in response to insulin with saline pre- treatment (N=9 mice). Right panel: Average GCaMP6 signal in response to insulin with CNO pre-treatment (N=9). GCaMP6 signal is normalized to baseline ((F-F0)/F0). m. Binned population activity (normalized GCaMP6 signal) following an ITT with pretreatment of CNO or saline. One-way RM ANOVA; CNO vs. Saline, p<0.001=†††. N=9 mice in each treatment. Bin width = 3 minutes. n. Plasma glucagon in response to an ITT, following CNO (or saline) injection (to inhibit A1/C1 neurons). Two-way RM ANOVA by both factors with Bonferroni’s multiple comparisons test. Saline vs. CNO; p<0.05=* (30 mins); p>0.99 (0 mins). 0 vs. 30 mins; p=0.005 (Vehicle); p=0.0293 (SSR149415). Time, p=0.028; Treatment, p=0.032; Interaction, p=0.027. N=6 wild- type mice.

81

To explore the consequences of A1/C1 activation in vivo, we injected AAV-DIO-

hM3Dq-mCherry bilaterally into the VLM of ThCre+ mice (Figure 3. 5f, Supplementary Figure 3.

7b, c). Activation of A1/C1 neurons with CNO evoked a ~4.5 mM increase in plasma glucose

(Figure 3. 5g, h). Pre-treatment with the glucagon receptor antagonist LY2409021 inhibited this response (Figure 3. 5g, h). In line with this, plasma glucagon was increased following CNO application (Figure 3. 5i). The hyperglycemic response was also dependent on functional V1bRs, because it was abolished following pre-treatment with the V1bR antagonist SSR149415 (Figure

3. 5g, h). CNO had no effect on blood glucose in ThCre+ mice expressing mCherry in A1/C1

neurons (Supplementary Figure 3. 7c). Together, these data indicate that A1/C1 neuron activation evokes AVP and glucagon secretion. We therefore hypothesised that hypoglycemia- induced AVP release is due to projections from A1/C1 neurons.

In line with this hypothesis, c-Fos expression (a marker of neuronal activity) was

increased in A1/C1 neurons following an ITT (Supplementary Figure 3. 8a and b). To determine

the contribution of A1/C1 neurons SONAVP neuron activity during an ITT, we sought to inhibit

A1/C1 neurons whilst monitoring SONAVP neuron activity. To this end, we expressed an

inhibitory receptor (the modified human muscarinic M4 receptor hM4Di; Armbruster et al.

(2007)) in A1/C1 neurons by injecting AAV-fDIO-hM4Di-mCherry into the VLM, whilst co-

injecting AAV-DIO-GCaMP6s into the SON of Dbhflp+ x Avpires-Cre+ mice (Figure 3. 5j). We then measured AVP neuron population Ca2+ activity (with in vivo fiber photometry) and plasma

glucagon following inhibition of A1/C1 neurons with CNO (Figure 3. 5j and Supplementary

Figure 3. 8c). SONAVP neuron population activity during an ITT was reduced by A1/C1 silencing

compared to vehicle injection (Figure 3. 5k-m). Furthermore, glucagon secretion was reduced

82 following CNO silencing of A1/C1 neurons (Figure 3. 5n). Together, these data suggest that

AVP-dependent glucagon secretion during an ITT is mediated by A1/C1 neurons.

3.2.4 AVP-induced glucagon secretion maintains glucose homeostasis during dehydration

Collectively, these data point to AVP being an important regulator of glucagon secretion, but do not explain why a receptor for a hormone commonly known for its anti-diuretic action is expressed in alpha-cells. We speculated that glucagon may play an important homeostatic role during dehydration - a physiological state where circulating AVP is elevated.

Mice were water restricted for 24 hours, but given unrestricted access to food. This resulted in a reduction in food intake (Figure 3. 6a) – a critically important behavioral response to water restriction, known as dehydration anorexia (Watts and Boyle, 2010). Despite the 30% reduction in food intake, plasma glucose was maintained (Figure 3. 6b). We hypothesized that this was due to AVP-induced glucagon secretion. In support of this hypothesis, glucagon was indeed elevated in these animals following 24 hour water restriction (Figure 3. 6c). In contrast, when the same mice were given ad lib access to water (which would not elevate plasma AVP) but had their food consumption restricted to that consumed in the dehydration trial, blood glucose fell and plasma glucagon was not elevated (Figure 3. 6b, c). The elevation in plasma glucagon was due to AVP, because pre-treatment with the V1bR antagonist SSR149415 (or vehicle) during water deprivation, blunted this increase (Figure 3. 6d, e). Therefore, in addition to serving an antidiuretic function, AVP stimulates glucagon secretion to aid the maintenance of plasma glucose in the dehydrated state (in spite of the reduction in feeding).

83

Figure 3.6. AVP-induced glucagon secretion during dehydration a. AVP-induced glucagon secretion during dehydration. Wild-type mice were first subjected to a 24 hour water restriction trial, wherein food consumption during was monitored. During the second trial (food restriction trial), the same mice were given unrestricted access to water, but had the same quantity of food available consumed in the first (water restriction) trial. One- way RM ANOVA; Dunnett’s post-hoc, p<0.05=*. N=8 wild-type mice. b. Blood glucose in the water and food restriction trial. Two way RM ANOVA by both factors; p<0.05=* and ns=0.71 (within trial); p<0.01=†† (between trials). Tukey’s multiple comparison tests. N=8 wild-type mice. c. Left panel: Plasma glucagon in the water and food restriction trial. Two way RM ANOVA by both factors; p<0.05=** and ns=0.49 (within trial). Tukey’s multiple comparison tests. Right panel: Same data, represented as plasma glucagon at 24 hour minus 0 hours. Mann Whitney- U paired t-test (p=0.023). N=8 wild-type mice. d. Left panel: Plasma glucagon before and after 24 hour water restriction, where either vehicle or the V1b receptor antagonist SSR149415 (30 mg/kg) was injected 90 minutes prior to termination of the restriction. Two way RM ANOVA by both factors; p<0.01=**, ns=not significant. Tukey’s multiple comparison tests. Right panel: Same data, represented as plasma glucagon at 24 hour minus 0 hours. Mann Whitney-U paired t-test (p<0.01=**). N=11-12 wild-type mice.

84

3.2.5 Insulin-induced AVP secretion may underlie counter-regulatory glucagon secretion in

human, and is diminished in subjects with Type 1 Diabetes

To understand whether this physiological pathway contributes to counter-regulatory glucagon secretion in human, samples were analyzed from an ongoing clinical trial

(NCT03954873). In two “saline arms” of the trial, healthy volunteers were given a hypoglycemic clamp during one visit, and a euglycemic clamp during another visit in a randomized order

(Figure 3. 7a-d; see also Supplementary Figure 3. 9a-b). In response to hypoglycemia (blood glucose of 2.8±0.1 mM, Figure 3. 7a), plasma glucagon rose (Figure 3. 7b) and so did plasma copeptin (Figure 3. 7c). In contrast, during euglycemia (blood glucose of 5.1±0.1 mM), plasma glucagon (Figure 3. 7a) and copeptin (Figure 3. 7b) fell. There was a correlation between glucagon and copeptin in the samples taken from both clamps (Figure 3. 7d).

To understand whether circulating AVP could be contributing to counter-regulatory

glucagon secretion in human subjects, islets isolated from human donors were studied. In human

islets from 8 donors, AVPR1B was the most abundant of the vasopressin receptor family (Figure

3. 7e). These data are supported both by a recent meta-analysis of single-cell RNA-seq data from

human donors (Mawla and Huising, 2019), and bulk sequencing of human islet fractions (Nica et

al., 2013). In human islets from 9 donors, AVP resulted in an increase in glucagon secretion

(Figure 3. 7f). Finally, AVP increased Ca2+ (Fluo-4) activity in islets isolated from 5 human

donors (Figure 3. 7g-h).

In T1D, deficiency of the secretory response of glucagon to hypoglycemia is an early

acquired abnormality of counter-regulation (Cryer, 2002). We therefore measured copeptin, AVP and glucagon during hypoglycemic clamps in subjects with T1D and age- and BMI-matched healthy controls from two previously published studies conducted on male participants

85

Figure 3.7. Insulin-induced hypoglycaemia evokes copeptin and glucagon secretion in human subjects a. Blood glucose was clamped at euglycaemia (Eug) and followed during insulin-induced hypoglycaemia (Hypo). N=10 healthy human subjects (see Supplementary Table 1). The data is from an ongoing clinical trial (NCT03954873). The clamp was initiated at time 0 min and terminated at 60 min. b. Plasma copeptin measurement during and following clamping period. Represented as change in copeptin from baseline (time = 0 min). Two-way RM ANOVA by both factors. Indicated time point vs. 0 min; p<0.05=*, p<0.01=**. Between treatments; p<0.05=†. Raw data is shown in Supplementary Figure 9a. c. Plasma glucagon measurement during insulin-induced hypoglycaemia. Two-way RM ANOVA by both factors. Time point vs. 0 min; p<0.05=*. Between treatments; p<0.05=†. Represented as change in glucagon from 0 min time point. Raw data is shown in Supplementary Figure 9b. d. Correlation of change in copeptin and change in glucagon, with a linear regression (dashed line). e. mRNA expression of AVPR1A, AVPR1B, AVPR2 and OXTR in human islets. Human islet samples N=7-8 donors. Calculated with the Pfaffl method, using Actb as the reference gene. f. Glucagon secretion from islets isolated from human donors, in response to AVP. One-way ANOVA, p<0.05=*. N=5-9 human donors. g. Fluo4 signal from a putative alpha-cell in a human islet in response to AVP (10, 100 and 1000 pM). Recording in 3 mM glucose. h. Area under curve (AUC, normalized to duration) for Fluo4 signal in each human islet, for each experimental condition. 26 islets, N=4 human donors. One-way ANOVA, p<0.05=*; p<0.001=***.

86

(Christensen et al., 2015; Christensen et al., 2011). At 60 minutes, the achieved glucose clamp

was similar between the subjects (Figure 3. 8a; see also Supplementary Figure 3. 9c-d). However, copeptin was significantly elevated in controls but not subjects with T1D (Figure 3. 8b). Insulin- induced hypoglycemia failed to increase circulating glucagon in subjects with T1D (Figure 3. 8d).

Finally, a correlation of copeptin and glucagon demonstrated that control subjects either had a greater change in copeptin in response to hypoglycemia, or a similar change in copeptin but this resulted in a larger increase in plasma glucagon (Figure 3. 8d).

Figure 3.8. Insulin-induced copeptin and glucagon secretion is diminished in type 1 diabetic subjects a. Hypoglycaemia was induced by an insulin infusion in patients with T1D (N=10; Christensen et al. (2015)) and non-diabetic individuals (Controls; N=10; Christensen et al. (2011)). The insulin infusion was initiated at 0 min and terminated at 90 min. See Supplementary Table 2 for details of these two cohorts. b. Plasma copeptin measurement during and following clamping period. Represented as change in copeptin from baseline (time = 0 min). Two-way RM ANOVA by both factors. Time point vs. 0 min; p<0.05=*, p<0.01=**. Between groups; p<0.05=†. Raw data is shown in Supplementary Figure 9c. c. Plasma glucagon measurement during the clamping period. Two-way RM ANOVA by both factors. Time vs. 0 min; p<0.05=*. Between groups; p<0.05=†. Represented as change in glucagon from 0 min time point. Raw data is shown in Supplementary Figure 9d. d. Correlation of change in copeptin and change in glucagon following hypoglycaemic clamping for control subjects (circle) and T1D (square).

87

3.3 Discussion

Here, we adopted a wide range of in vitro and in vivo techniques in mice and humans to provide a comprehensive analysis of central and peripheral pathways mediating AVP’s glucoregulatory function. We performed the very first in vivo simultaneous real-time monitoring of blood glucose and magnocellular AVP neuron activity in mice to demonstrate robust activation of magnocellular AVP neurons by hypoglycemia. We further demonstrated that

A1/C1 neurons in the brainstem mediate this activation. Activation of SONAVP neurons caused a significant increase in blood glucose, which was mediated by V1bR. Experiments using isolated mouse and human islets indicated that this effect was due to a direct action of AVP on glucagon- secreting alpha-cells. V1bR blockade during dehydration and either insulin- or 2DG-induced hypoglycemia, robustly reduced glucagon release, suggesting that AVP is a key mediator of glucagon secretion.

Despite substantial evidence suggesting a glucoregulatory function of AVP, mechanisms underlying its role in glucose homeostasis have remained enigmatic. Here, we provide a

complete neural and molecular description of the AVP glucoregulatory pathway: A1/C1 neurons

of the VLM  magnocellular AVP neurons  V1bR in pancreatic alpha-cells  glucagon

release  hepatic glucose production  increase in blood glucose. Moreover, with

combinatorial use of the latest neuroscience and metabolic analysis techniques, our study

provides the first direct evidence for glucoregulatory function of endogenous AVP in vivo.

We demonstrate a second-to-second relationship between blood glucose and SONAVP

neuron activity using continuous glucose monitoring and fiber photometry. This is a substantial improvement from previous studies measuring blood glucose and AVP levels, which have the temporal resolution of several minutes. Also, such studies based on blood levels cannot be done

88

in mice due to the large blood volume necessary to measure AVP. It is worth noting that combined use of continuous glucose monitoring and fiber photometry is unprecedented and this approach will open up countless opportunities to explore central regulation of glucose homeostasis because it allows one to precisely, in real time, correlate blood glucose levels with neuron activity.

In addition, we investigated the effect of elevated circulating AVP on blood glucose and glucagon using a chemogenetic approach. Stimulation of SONAVP neurons with hM3Dq resulted in an increased release of endogenous AVP into the circulation, whereas previous studies mainly depended on i.p injection of i.v. infusion of exogenous AVP or synthetic AVP analogs (Hems et al., 1975; Spruce et al., 1985; Taveau et al., 2017). This is likely a reason why the hyperglycemic effect observed in our study lasted much longer than previously reported (1h vs. 10 min). AVP has a biological half-life of 5-12 min (Fenske et al., 2018; Janaky et al., 1982). Rapid degradation of AVP might account for the transient increase in blood glucose observed after acute injection or infusion of AVP. In contrast, chemogenetic activation typically lasts several hours, allowing better simulation of SONAVP neuron activity in vivo in responses to hypoglycemia.

In sum, our study provides biological and methodological insights into understanding the mechanism of counter-regulatory glucagon release and pathophysiology of T1D.

89

3.4 Methods

3.4.1 Ethics

All animal experiments were conducted in strict accordance to regulations enforced by the research institution. Experiments conducted in the UK were done so in accordance with the

UK Animals Scientific Procedures Act (1986) and University of Oxford and Imperial College

London ethical guidelines, and were approved by the local Ethical Committee. All animal care and experimental procedures conducted in the U.S.A. were approved by the Beth Israel

Deaconess Medical Center Institutional Animal Care and Use Committee. Animal experiments conducted in Goteborg University were approved by a local Ethics Committee.

Human pancreatic islets were isolated, with ethical approval and clinical consent, at the

Diabetes Research and Wellness Foundation Human Islet Isolation Facility (OCDEM, Oxford,

UK) or Alberta Diabetes Institute IsletCore (University of Alberta, AB, Canada). Islets from a total of 19 human donors were used in this study. Donor details were as follows; age = 42 ± 4 years; BMI = 27.1 ± 3; Sex = 10/9 (M/F).

3.4.2 Animals

All animals were kept in a specific pathogen-free (SPF) facility under a 12:12 hour light:dark cycle at 22 °C, with unrestricted access to standard rodent chow and water. C57BL/6J mice used in this study are referred to as wild-type mice. To generate alpha-cell specific expression of the genetically-encoded Ca2+ sensor GCaMP3, mice carrying Cre recombinase under the control of the proglucagon promoter (GcgCre+ mice) were crossed with mice with a floxed green calmodulin (GCaMP3) Ca2+ indicator in the ROSA26 locus (The Jackson

Laboratory). These mice are referred to as Gcg-GCaMP3 mice. To generate mice expressing

90

RFP in alpha-cells, GcgCre+ were crossed with mice containing a floxed tandem-dimer red

fluorescent protein (tdFRP) in the ROSA26 locus (Gcg-RFP mice). Both of these mouse models

have previously been described (Dadi et al., 2015) and were kept on a C57BL/6 background.

Other transgenic mouse strains used – namely, Avpires-Cre+ (Pei et al., 2014), ThCre+ (The Jackson

Laboratory), Dbhflp+ (MMRCC) and AvpGFP (MMRCC) - were heterozygous for the transgene and maintained on a mixed background. Avpr1b-/- and littermate controls (Avpr1b+/+) were bred and maintained as previously described (Wersinger et al., 2002). Leptin receptor-deficient

(db/db) mice and controls (wt/wt) were purchased from Jackson Laboratories (stock number

000642 BKS.Cg-Dock7m +/+ Leprdb/J (Wildtype for Dock7m, Homozygous for Leprdb), 000662

C57BLKS/J).

3.4.3 Isolation of mouse islets

Mice of both sex and 11-16 weeks of age were killed by cervical dislocation (UK

Schedule 1 procedure). Pancreatic islets were isolated by liberase digestion followed by manual

picking. Islets were used acutely and were, pending the experiments, maintained in tissue culture

for <36 hour in RPMI 1640 (11879-020, Gibco, Thermo Fisher Scientific) containing 1%

pennicillin/streptomycin (1214-122, Gibco, Thermo Fisher Scientific), 10%FBS (F7524-500G,

Sigma-Aldrich) and 11 mM glucose, prior to experiments.

3.4.4 Patch-clamp electrophysiology in islets

Mouse islets were used for patch-clamp electrophysiological recordings. These

recordings (in intact islets) were performed at 33-34 C using an EPC-10 patch-clamp amplifier

and PatchMaster software (HEKA Electronics, Lambrecht/Pfalz,⁰ Germany). Unless otherwise

91

stated, recordings were made in 3 mM glucose, to mimic hypoglycaemic conditions in mice.

Currents were filtered at 2.9 kHz and digitized at > 10 kHz. A new islet was used for each

recording. Membrane potential ( ) recordings were conducted using the perforated patch-

푀 clamp technique, as previously described푉 (Briant et al., 2018). The pipette solution contained (in mM) 76 K2SO4, 10 NaCl, 10 KCl, 1 MgCl2·6H20 and 5 Hepes (pH 7.35 with KOH). For these experiments, the bath solution contained (mM) 140 NaCl, 3.6 KCl, 10 Hepes, 0.5 MgCl2·6H20,

0.5 Na2H2PO4, 5 NaHCO3 and 1.5 CaCl2 (pH 7.4 with NaOH). Amphotericin B (final concentration of 25 mg/mL, Sigma-Aldrich) was added to the pipette solution to give electrical access to the cells (series resistance of <100 MΩ). Alpha-cells in Gcg-GCaMP3 islets were confirmed by the presence of GCaMP3.

The frequency of action potential firing was calculated in MATLAB v. 6.1 (2000; The

MathWorks, Natick, MA). In brief, a peak-find algorithm was used to detect action potentials.

This was then used to calculate firing frequency in different experimental conditions (AVP concentrations). Power-spectrum analysis of was conducted in Spike2 (CED, Cambridge,

푀 UK). was moving-average filtered (interval푉 of 200 ms) and the mean subtracted. A

푀 푀 power푉-spectrum was then produced (Hanning window with 0.15 Hz resolution)푉 during 3 mM

glucose alone, and with 10 pM AVP.

3.4.5 GCaMP3 imaging in mouse islets

2+ 2+ Time-lapse imaging of the intracellular Ca concentration ([Ca ]i) in Gcg-GCaMP3

mouse islets was performed on an inverted Zeiss AxioVert 200 microscope, equipped with the

Zeiss LSM 510-META laser confocal scanning system, using a 403/1.3 NA objective. Mouse

islets were transferred into a custom-built recording chamber. Islets were then continuously

92

perfused with bath solution at a rate of 200 µL/min. The bath solution contained (in mM): 140

NaCl, 5 KCl, 1.2 MgCl2, 2.6 CaCl2, 1 NaH2PO4, 10 Hepes, 17 mannitol and 3 glucose. GCaMP3

was excited at 430 nm and recorded at 300-530 nm. The pinhole diameter was kept constant, and

frames of 256 x 256 pixels were taken every 800 ms. Unless otherwise stated, recordings were

made in 3 mM glucose, to mimic hypoglycaemic conditions in mice. Raw GCaMP3 data was

processed as follows; regions of interest (ROIs) were manually drawn around each GCaMP3+

cell in ImageJ and the time-series of the GCaMP3 signal for each cell was exported. These data

were first imported into Spike2 7.04 (CED, Cambridge, UK), wherein the data was median

filtered to remove baseline drift. The size of the filter was optimised for each individual cell to

remove drift/artefacts but preserve Ca2+ transients. Ca2+ transients were then automatically detected using the built in peak-find algorithm; the amplitude of peaks to be detected was dependent on the SNR but was typically > 20% of the maximal signal intensity. Following this, frequency of Ca2+ transients could be determined. For plotting Ca2+ data, the data was imported

into MATLAB.

3.4.6 DAG measurements in mouse islets

The effects of AVP on the intracellular diacylglycerol concentration (DAG) in pancreatic islet cells was studied using a recombinant circularly permutated probe, Upward DAG (Montana

Molecular). Islets isolated from wild-type mice were gentle dispersed (using Trypsin ES) into clusters and platted on rectangular coverslips. Cell clusters were then transfected with Upward

DAG, delivered via a BacMam infection (according to the manufacturer’s guidelines).

Coverslips were then were placed in a custom built chamber. Imaging experiments were performed 36-48 hours after infection using a Zeiss AxioZoom.V16 zoom microscope equipped

93

with a ×2.3/0.57 objective (Carl Zeiss). The fluorescence was excited at 480 nm, and the emitted

light was collected at 515 nm. The cells were kept at 33-35 °C and perfused continuously

throughout the experiment with KRB solution supplemented with 3 mM glucose. The images

were acquired using Zen Blue software (Carl Zeiss). The mean intensity for each cell was

determined by manually drawing ROIs in ImageJ. Data analysis and representation was

performed with MATLAB. All data was processed using a moving average filter function

(smooth) with a span of 50 mins, minimum subtracted and then normalised to maximum signal

intensity in the time-series. AUC was calculated using the trapz function and then divided by the

length of the condition.

3.4.7 Ca2+ imaging in human islets

2+ Time-lapse imaging of [Ca ]i in human islets was performed on the inverted Zeiss

SteREO Discovery V20 Microscope, using a PlanApo S 3.5x mono objective. Human islets were loaded with 5 µg/µL of the Ca2+-sensitive dye Fluo-4 (1923626, Invitrogen, Thermo Fisher

Scientific) for 60 min before being transferred to a recording chamber. Islets were then

continuously perfused with DMEM (11885-084, Gico, Thermo Fisher Scientific) with 10% FBS,

100 units/mL penicillin and 100 mg/mL streptomycin at a rate of 200 µL/min. Fluo-4 was

excited at 488 nm and fluorescence emission collected at 530 nm. The pinhole diameter was kept

constant, and frames of 1388x1040 pixels were taken every 3 sec. The mean intensity for each

islet was determined by manually drawing an ROI around the islet in ImageJ. Data analysis and

representation was performed with MATLAB. All data was processed using a moving average

filter function (smooth) with a span of 20 mins, minimum subtracted and then normalised to

94 maximum signal intensity in the time-series. AUC was calculated using the trapz function and then divided by the length of the condition.

3.4.8 Pancreatic islet isolation, transplantation and in vivo imaging of islets implanted into the anterior chamber of the eye (ACE).

Pancreatic islets from Gcg-GCaMP3 mice were isolated and cultured as described above.

For transplantation, 10-20 islets were aspirated with a 27-gauge blunt eye cannula

(BeaverVisitec, UK) connected to a 100 µl Hamilton syringe (Hamilton) via 0.4-mm polyethylene tubing (Portex Limited). Prior to surgery, mice (C57BL6/J) were anaesthetised with

2-4% isoflurane (Zoetis) and placed in a stereotactic frame. The cornea was incised near the junction with the sclera, then the blunt cannula (pre-loaded with islets) was inserted into the ACE and islets were expelled (average injection volume 20 μl for 10 islets). Carprofen (Bayer, UK) and eye ointment were administered post-surgery. A minimum of four weeks was allowed for full implantation before imaging. Imaging sessions were performed with the mouse held in a stereotactic frame and the eye gently retracted, with the animal maintained under 2-4% isoflurane anaesthesia. All imaging experiments were conducted using a spinning disk confocal microscope (Nikon Eclipse Ti, Crest spinning disk, 20x water dipping 1.0 NA objective). The signal from GCaMP3 (ex. 488 nm, em. 525±25 nm) was monitored at 3 Hz for up to 20 min.

After a baseline recording, mice received a bolus of AVP (10 µg/kg) i.v. (tail vein). Data were imported into ImageJ for initial movement correction (conducted with the StackReg and

TurboReg plugins) and ROI selection. Analysis was then conducted in MATLAB.

95

3.4.9 Hormone secretion measurements from mouse and human islets

Islets, from human donors or isolated from wild-type mice, were incubated for 1 h in

RPMI or DMEM supplemented with 7.5 mM glucose in a cell culture incubator. Size-matched batches of 15-20 islets were pre-incubated in 0.2 ml KRB with 2 mg/ml BSA (S6003, Sigma-

Aldrich) and 3 mM glucose for 1 hour in a water-bath at 37 C. Following this islets were statically subjected to 0.2 ml KRB with 2 mg/ml BSA with the⁰ condition (e.g. 10 pM AVP) for 1 hour. After each incubation, the supernatant was removed and kept, and 0.1 ml of acid:etoh

(1:15) was added to the islets. Both of these were then stored at -80 C. Each condition was repeated in at least triplicates. ⁰

Glucagon and insulin measurements in supernatants and content measurements were performed using a dual mouse insulin/glucagon assay system (Meso Scale Discovery, MD,

U.S.A.) according to the protocol provided.

3.4.10 Hormone secretion measurements in the perfused mouse pancreas

Dynamic measurements of glucagon were performed using the in situ perfused mouse pancreas. Briefly, the aorta was cannulated by ligating above the coeliac artery and below the superior mesenteric artery, and the pancreas was perfused with KRB at a rate of ~0.45ml/min using an Ismatec Reglo Digital MS2/12 peristaltic pump. The KRB solution was maintained at

37 C with a Warner Instruments temperature control unit TC-32 4B in conjunction with a tube heater⁰ (Warner Instruments P/N 64-0102) and a Harvard Apparatus heated rodent operating table. The effluent was collected by cannulating the portal vein and using a Teledyne ISCO Foxy

R1 fraction collector. The pancreas was first perfused for 10 min with 3 mM glucose before

96

commencing the experiment to establish the basal rate of secretion. Glucagon measurements in

collected effluent were performed using RIA.

3.4.11 Flow cytometry of islet cells (FACS), RNA extraction, cDNA synthesis and

quantitative PCR

The expression of the AVPR gene family was analysed in tissues from 12-week old

C57BL6/J mice (3 mice) and pancreatic islets from human donors (2 samples, each comprised of

pooled islet cDNA from 7 and 8 donors, respectively). Total RNA was isolated using a

combination of TRIzol and PureLink RNA Mini Kit (Ambion, Thermofisher Scientific) with

incorporated DNase treatment.

Pancreatic islets from Gcg-RFP mice were isolated and then dissociated into single cells by trypsin digestion and mechanical dissociation. Single cells were passed through a MoFlo

Legacy (Beckman Coulter). Cells were purified by combining several narrow gates. Forward and side scatter were used to isolate small cells and to exclude cell debris. Cells were then gated on pulse width to exclude doublets or triplets. RFP+ cells were excited with a 488 nm laser and the fluorescent signal was detected through a 580/30 bandpass filter (i.e. in the range 565-595 nm).

RFP-negative cells were collected in parallel. The levels of gene expression in the RFP+ and in

the RFP- FAC-sorted fractions were determined using real-time quantitative PCR (qPCR). RNA

from FACS-sorted islet cells was isolated using RNeasy Micro Kit (Qiagen). cDNA was

synthesized using the High Capacity RNA-to-cDNA kit (Applied Biosystems, Thermofisher

Scientific). Real time qPCR was performed using SYBR Green detection and gene specific

QuantiTect Primer Assays (Qiagen) on a 7900HT Applied Biosystems analyser. All reactions

97

were run in triplicates. Relative expression was calculated using ΔCt method Actb as a reference

gene.

3.4.12 Stereotaxic surgery and viral injections

For viral injections into the SON, mice were anaesthetized with ketamine/xylazine (100

and 10 mg/kg, respectively, i.p.) and then placed in a stereotaxic apparatus (David Kopf model

940). A pulled glass micropipette (20-40 μm diameter tip) was used for stereotaxic injections of adeno-associated virus (AAV). Virus was injected into the SON (200 nl/side; AP: −0.65 mm;

ML: ±1.25 mm; DV: -5.4 mm from bregma) by an air pressure system using picoliter air puffs through a solenoid valve (Clippard EV 24VDC) pulsed by a Grass S48 stimulator to control injection speed (40 nL/min). The pipette was removed 3 min post-injection followed by wound closure using tissue adhesive (3M Vetbond). For viral injections into the VLM, mice were placed into a stereotaxic apparatus with the head angled down at approximately 45°. An incision was made at the level of the cisterna magna, then skin and muscle were retracted to expose the dura mater covering the 4th ventricle. A 28-gauge needle was used to make an incision in the dura and allow access to the VLM. Virus was then injected into the VLM (50nl*2/side; AP: -0.3 and -

0.6mm; ML: ±1.3mm; DV: -1.7mm from obex) as described above. The pipette was removed 3 min post-injection followed by wound closure using absorbable suture for muscle and silk suture for skin. For fibre photometry, an optic fibre (200 µm diameter, NA=0.39, metal ferrule,

Thorlabs) was implanted in the SON and secured to the skull with dental cement. Subcutaneous injection of sustained release Meloxicam (4 mg/kg) was provided as postoperative care. The mouse was kept in a warm environment and closely monitored until resuming normal activity.

Chemogenetic experiments utilized AAV8-hSyn-DIO-hM3Dq-mCherry (Addgene:44361) and

98

AAV8-nEF-fDIO-hM4Di-mCherry (custom-made vector) produced from Boston Children’s

Hospital Viral Core and AAV5-EF1α-DIO-mCherry purchased from the UNC Vector Core.

Fibre photometry experiments were conducted using AAV1-hSyn-FLEX-GCaMP6s purchased

from the University of Pennsylvania (School of Medicine Vector Core). Projection mapping and

ChR2-assisted circuit mapping were done using AAV9-EF1α-DIO-ChR2(H134R)-mCherry

purchased from the University of Pennsylvania (School of Medicine Vector Core).

3.4.13 Fiber photometry experiments and analysis of photometry data

In vivo fiber photometry was conducted as previously described (Mandelblat-Cerf et al.

(2017)). A fiber optic cable (1-m long, metal ferrule, 400 µm diameter; Doric Lenses) was

attached to the implanted optic cannula with zirconia sleeves (Doric Lenses). Laser light (473

nm) was focused on the opposite end of the fiber optic cable to titrate the light intensity entering

the brain to 0.1-0.2 mW. Emitted light was passed through a dichroic mirror (Di02-R488-25x36,

Semrock) and GFP emission filter (FF03-525/50-25, Semrock), before being focused onto a

sensitive photodetector (Newport part #2151). The GCaMP6 signal was passed through a low-

pass filter (50 Hz), and digitized at 1 KHz using a National Instruments data acquisition card and

MATLAB software.

All experiments were conducted in the home-cage in freely moving mice. Animals

prepared for in vivo fiber photometry experiments (outlined above), were subjected to an ITT or

2DG injection after overnight fasting. Prior to insulin or 2DG injection, a period of GCaMP6s

activity was recorded (3 min) to establish baseline activity. Insulin (i.p. 2 U/kg), 2DG (i.p.

500mg/kg), or saline vehicle was then administered, and GCaMP6 activity recorded for a further

40 min. In some experiments, mice were pre-treated i.p. with CNO (1mg/kg) or saline, 30

99

minutes prior to insulin or 2DG. The recorded data was exported and then imported into

MATLAB for analysis. Fluorescent traces were down-sampled to 1 Hz and the signal was

normalised to the baseline (F0 mean activity during baseline activity), with 100% signal being

defined as the maximum signal in the entire trace (excluding the injection artefact). Following

the ITT, the signal was binned (1 min) and a mean for each bin calculated. These binned signals

were compared to baseline signal using a one-way RM ANOVA.

3.4.14 Surgery for continuous glucose monitoring

Animals that have undergone fibre photometry surgeries (3 weeks prior) were anesthetized and maintained with isoflurane. Once mice were fully anesthetized, the ventral abdomen and underside of the neck were shaved and disinfected. Animals were placed on their backs on a heated surgical surface. For transmitter implantation, a ventral midline abdomen incision was made and the abdominal wall was incised. The transmitter was placed in the abdominal cavity with the lead exiting cranially and the sensor and connector board exteriorized.

The incision was sutured incorporating the suture rib into the closure. For glucose probe implantation, a midline neck incision was performed and the left common carotid artery was isolated. The vessel was then perforated and the sensor of the glucose probe (HD-XG, Data

Sciences International) was advanced into the artery towards the heart, within a final placement in the aortic arch. Once in place, the catheter was secured by tying the suture around the catheter and vessel, and overlying opening in tissue was closed. Mice were kept warm on a heating pad and monitored closely until fully recovered from anaesthesia.

100

3.4.15 Simultaneous AVP fiber photometry and continuous glucose monitoring

All experiments were conducted in the home-cage in freely moving mice. Animals prepared for in vivo fibre photometry and continuous glucose monitoring (outlined above), were subjected to an ITT after overnight fasting. After establishing > 3 min of baseline activity, insulin (i.p. 1 or 1.5 U/kg) or saline vehicle was administered. GCaMP6s activity, blood glucose, and body temperature were recorded throughout 2 h of experiment. Each recording was separated by at least 48h. GCaMP6s recording was performed as described above. Blood glucose and body temperature were acquired using Dataquest A.R.T. 4.36 system and analysed using

MATLAB. Calibration of HD-XG device was performed as per manufacturer’s manual.

3.4.16 In vivo measurements of plasma glucose, glucagon and copeptin

Samples for blood glucose and plasma glucagon measurements were taken from mice in response to different metabolic challenges (described in detail below). Both sexes were used for these experiments. Blood glucose was measured with an Accu-Chek Aviva (Roche Diagnostic,

UK) and OneTouch Ultra (LifeScan, UK). Plasma copeptin was measured using an ELISA

(MyBioSource, USA and Neo Scientific, USA).

3.4.17 Insulin tolerance test

Mice were restrained and a tail vein sample of blood was used to measure fed plasma glucose. A further sample was extracted into EDTA coated tubes for glucagon measurements.

Aprotinin (1:5, 4 TIU/ml; Sigma-Aldrich, UK) was added to all blood samples. These blood samples were kept on ice until the end of the experiment. Mice were first administered with any necessary pre-treatment and then individually caged. Pre-treatments included SSR149415 (30

101 mg/kg in PBS with 5% DMSO and 5% Cremophor EL), LY2409021 (5 mg/kg in PBS with 5%

DMSO), CNO (1-3 mg/kg in PBS with 5% DMSO) or the appropriate vehicle. After a 30 minute period, mice were restrained again, and blood was taken via a tail vein or submandibular bleed.

This was used for blood glucose measurements, and also for glucagon. Insulin (0.75, 1 or 1.5

U/kg) was then administered i.p., and the mice were re-caged. At regular time intervals after the insulin injection, mice were restrained and a blood sample extracted. Blood glucose was measured, and blood was taken for glucagon measurements. At the end of the experiment, blood samples were centrifuged at 2700 rpm for 10 min at 4 °C to obtain plasma. The plasma was then removed and stored at -80 °C. Plasma glucagon measurements were conducted using the 10-µl glucagon assay system (Mercodia, Upsala, Sweden), according to the manufacturer’s protocol.

3.4.18 Fasting experiments

Wild-type mice were used for fasting experiments. Mice (8-9 weeks of age) were restrained, and blood was taken for blood glucose and glucagon measurements (as above). The mice were then home caged for the period of the fast. Food was removed at 08:30, and samples were taken at 14:30 and 16:00 (7.5 hour fast). SSR149415 (30 mg/kg) or vehicle was administered i.p. at 14:30 (90 minutes prior to termination of the fast). Blood glucagon was measured as indicated above. All animals had free access to water during the fast.

3.4.19 Water restriction (dehydration) experiments

Wild-type mice (8-9 weeks of age) were used for water restriction experiments. Mice were single housed one week prior to experimental manipulation. For trial 1 (water restriction), blood was taken at 16:00 on day 1 and 2 for blood glucose and glucagon measurements (as

102

above). Mice were water restricted for 24h between day 1 and 2, but had ad lib access to food.

Amount of food consumed during this period was measured at the end of the trial and used for

trial 2. For trial 2 (food restriction), blood was taken at 16:00 on day 1 and 2 for blood glucose

and glucagon measurements. Mice had ad lib access to water between day 1 and 2, but were

given the same amount of food consumed in trial 1. The two trials were separated by 2 days without any manipulation.

3.4.20 Glucoprivic response to 2-Deoxy-D-glucose

Wild-type mice were used for 2-Deoxy-D-glucose (2DG) experiments. The mice were single housed one week prior to experimental manipulation. On the experimental day, food was removed 4 hours prior to the experiment. 2DG (500 mg/kg) or saline vehicle was then administered i.p., and blood samples taken at regular intervals for blood glucose and plasma glucagon measurements. In some cohorts, the V1bR antagonist SSR149415 (30 mg/kg in PBS with 5% DMSO and 5% Tween 80), glucagon receptor antagonist LY240901 (5 mg/kg in PBS with 5% DMSO and 5% Tween 80) or appropriate vehicle was administered i.p. 30 minutes prior to administration of 2DG. Plasma glucagon was measured as described above.

3.4.21 Brain slice electrophysiology

To prepare brain slices for electrophysiological recordings, brains were removed from anesthetized mice (4–8 weeks old) and immediately placed in ice-cold cutting solution

consisting of (in mM): 72 sucrose, 83 NaCl, 2.5 KCl, 1 NaH2PO4, 26 NaHCO3, 22 glucose, 5

MgCl2, 1 CaCl2, oxygenated with 95% O2 /5% CO2 , measured osmolarity 310-320 mOsm/l.

Cutting solution was prepared and used within 72 hours. 250 μm-thick coronal sections

103

containing the PVH and SON were cut with a vibratome (7000smz2-Campden Instruments) and

incubated in oxygenated cutting solution at 34 °C for 25 min. Slices were transferred to

oxygenated aCSF (126 mM NaCl, 21.4 mM NaHCO3, 2.5 mM KCl, 1.2 mM NaH2PO4, 1.2 mM

MgCl2, 2.4 mM CaCl2, 10 mM glucose) and stored in the same solution at room temperature (20-

24 °C) for at least 60 min prior to recording. A single slice was placed in the recording chamber

where it was continuously super-fused at a rate of 3–4 ml per min with oxygenated aCSF.

Neurons were visualized with an upright microscope equipped with infrared-differential

interference contrast and fluorescence optics. Borosilicate glass microelectrodes (5–7 MΩ) were

filled with internal solution. All recordings were made using Multiclamp 700B amplifier, and

data was filtered at 2 kHz and digitized at 10 kHz. All analysis was conducted off-line in

MATLAB.

3.4.22 Channelrhodopsin-2 assisted circuit mapping (CRACM) of connections from A1/C1

neurons to the SON

A Cre-dependent viral vector containing the light-gated ion channel channelrhodopsin-2

(AAV-DIO-ChR2-mCherry) was injected into the VLM (targeting A1/C1 neurons) of AvpGFP x

ThCre+ mice (see Figure 3. 6a). Brain slices were prepared (as above) from these mice. The SON was located by using the bifurcation of the anterior and middle cerebral arteries on the ventral surface of the brain as a landmark. Sections of 250 µm thickness of the SON were then cut with a Leica VT1000S or Campden Instrument 7000smz-2 vibratome, and incubated in oxygenated aCSF (126 mM NaCl, 21.4 mM NaHCO3, 2.5 mM KCl, 1.2 mM NaH2PO4, 1.2 mM MgCl2, 2.4

mM CaCl2, 10 mM glucose) at 34 °C for 25 min. Slices recovered for 1 hr at room temperature

(20–24°C) prior to recording. Whole-cell voltage clamp recordings were obtained using a Cs-

104

based internal solution containing (in mM): 135 CsMeSO3, 10 HEPES, 1 EGTA, 4 MgCl2, 4

Na2-ATP, 0.4 Na2-GTP, 10 Na2-phosphocreatine (pH 7.3; 295 mOsm). To photostimulate ChR2- positive A1/C1 fibres, an LED light source (473 nm) was used. The blue light was focused on to the back aperture of the microscope objective, producing a wide-field exposure around the recorded cell of 1 mW. The light power at the specimen was measured using an optical power meter PM100D (ThorLabs). The light output is controlled by a programmable pulse stimulator,

Master-8 (AMPI Co. Israel) and the pClamp 10.2 software (AXON Instruments).

3.4.23 Activation of hM3Dq with CNO in AVP neurons

The modified human M3 muscarinic receptor hM3Dq (Alexander et al., 2009) was

expressed in AVP neurons by injecting a Cre-dependent virus containing hM3Dq (AAV-DIO-

hM3Dq-mCherry) into the SON of mice bearing Avp promoter-driven Cre recombinase (Avpires-

Cre+ mice; see Figure 3. 1a). The intracellular solution for current clamp recordings contained the

following (in mM): 128 K gluconate, 10 KCl, 10 HEPES, 1 EGTA, 1 MgCl2, 0.3 CaCl2, 5

Na2ATP, 0.3 NaGTP, adjusted to pH 7.3 with KOH.

3.4.24 Brain immunohistochemistry

Mice were terminally anesthetized with chloral hydrate (Sigma-Aldrich) and trans- cardially perfused with phosphate-buffered saline (PBS) followed by 10% neutral buffered formalin (Fisher Scientific). Brains were extracted, cryoprotected in 20% sucrose, sectioned coronally on a freezing sliding microtome (Leica Biosystems) at 40 μm thickness, and collected in two equal series. Brain sections were washed in 0.1 M PBS with Tween-20, pH 7.4 (PBST),

105

blocked in 3% normal donkey serum/0.25% Triton X-100 in PBS for 1 h at room temperature

and then incubated overnight at room temperature in blocking solution containing primary

antiserum (rat anti-mCherry, Invitrogen M11217, 1:1,000; chicken anti-GFP, Life Technologies

A10262, 1:1,000; rabbit anti-vasopressin, Sigma-Aldrich AB1565, 1:1,000; rabbit anti-TH,

Millipore AB152, 1:1,000). The next morning, sections were extensively washed in PBS and then incubated in Alexa-fluor secondary antibody (1:1000) for 2 h at room temperature. After several washes in PBS, sections were incubated in DAPI solution (1 µg/ml in PBS) for 30 min.

Then, sections were mounted on gelatin-coated slides and fluorescence images were captured using an Olympus VS120 slide scanner.

3.4.25 Reagents

AVP, hydrocortisone and adrenaline were all from Sigma (Sigma-Aldrich, UK). The Gq

inhibitor YM-254890 (Takasaki et al., 2004) was from Wako-Chem (Wake Pure Chemical

Corp). The selective V1bR antagonist SSR149415 (nelivaptan; Serradeil-Le Gal et al. (2002)) was from Tocris (Bio-Techne Ltd, UK). The glucagon receptor antagonist LY2409021

(adomeglivant; Kazda et al. (2016)) was from MedKoo Biosciences (USA).

3.4.26 Clamping studies in human participants

Clamping studies were conducted at Gentofte Hospital, University of Copenhagen. The

studies were approved by the Scientific-Ethical Committee of the Capital Region of Denmark

(registration no. H-D-2009-0078) and was conducted according to the principles of the

Declaration of Helsinki (fifth revision, Edinburgh, 2000).

106

3.4.27 Comparison of copeptin in healthy subjects undergoing a hypoglycemic and

euglycemic clamp

Samples from the “saline arm” from 10 male subjects enrolled in an ongoing,

unpublished clinical trial (https://clinicaltrials.gov/ct2/show/NCT03954873) were used to

compare copeptin secretion during euglycaemia and hypoglycaemia.

For the study, two cannulae were inserted bilaterally into the cubital veins for infusions

and blood sampling, respectively. For the euglycaemic clamp, 20% dextrose (w/v) was

administered to keep plasma glucose at pre-clamp levels (~5 mM). For the hypoglycaemic

clamp, an intravenous insulin (Actrapid; Novo Nordisk, Bagsværd, Denmark) infusion was

initiated at time 0 min to lower plasma glucose. Plasma glucose was measured bedside every 5

min and kept >2 mM. Arterial blood was drawn at regular time intervals prior to and during

insulin infusion.

3.4.28 Comparison of copeptin in subjects with T1DM and healthy controls

Samples from 20 male (N=10 control and N=10 T1DM patients) from the “saline arms” of two previously published studies (Christensen et al., 2015; Christensen et al., 2011) were used

to compare copeptin secretion during a hypoglycaemic clamp between T1DM and control

subjects. The samples from healthy individuals (Controls) were from Christensen et al. (2011).

2 These 10 healthy male subjects were of age 23 ± 1 years, BMI 23 ± 0.5 kg/m and HbA1c 5.5 ±

0.1%. The T1DM patient samples were from (Christensen et al., 2015). These patients were; C-

2 peptide negative, age 26 ± 1 years, BMI 24 ± 0.5 kg/m , HbA1c 7.3 ± 0.2%, positive islet cell

and/or GAD-65 antibodies, treated with multiple doses of insulin (N = 9) or insulin pump (N =

1), without late diabetes complications, without hypoglycemia unawareness, and without residual

107

β-cell function (i.e., C-peptide negative after a 5-g arginine stimulation test). For the study, a hypoglycaemic clamp was conducted as outlined above.

3.4.29 Measurement of copeptin, glucagon and AVP in human plasma

Copeptin in human plasma was analysed using the KRYPTOR compact PLUS (Brahms

Instruments, Thermo Fisher, DE). Glucagon was measured using human glucagon ELISA from

Mercodia. Plasma AVP was measured using a human AVP ELISA kit (Cusabio, China).

3.4.30 Statistical tests of data

All data are reported as mean ± SEM. Unless otherwise stated, N refers to the number of mice. Statistical significance was defined as p < 0.05. All statistical tests were conducted in

Prism8 (GraphPad Software, San Diego, CA, USA). For two groupings, a t test was conducted with the appropriate post hoc test. For more than two groupings, a one-way ANOVA was conducted (repeated measures, if appropriate). If data were separated by two treatments/factors, then a two-way ANOVA was conducted. A repeated measures (RM) two-way ANOVA was used

(if appropriate), and a mixed-models ANOVA was used in the event of a repeated measures with missing data.

2+ 3.4.31 Mathematical model of Gq-mediated Ca oscillations

The mathematical model of the processes governing Ca2+ signalling in a single cell

following activation of a metabotropic receptor (i.e. V1bR) is based on the formalism by Li and

Rinzel (1994) and Lemon et al. (2003). The simulations were conducted in the open source

package software XPPAUT (http://www.math.pitt.edu/~bard/xpp/xpp.html).

108

3.5 References

Adler, G.K., and Majzoub, J.A. (1987). Influence of infused hypertonic saline on the response to insulin-induced hypoglycemia in man. The Journal of clinical endocrinology and metabolism 65, 116-121.

Alexander, G.M., Rogan, S.C., Abbas, A.I., Armbruster, B.N., Pei, Y., Allen, J.A., Nonneman, R.J., Hartmann, J., Moy, S.S., Nicolelis, M.A., et al. (2009). Remote control of neuronal activity in transgenic mice expressing evolved G protein-coupled receptors. Neuron 63, 27-39.

Andrew, S.F., Dinh, T.T., and Ritter, S. (2007). Localized glucoprivation of hindbrain sites elicits corticosterone and glucagon secretion. Am J Physiol Regul Integr Comp Physiol 292, R1792-1798.

Antoni, F.A. (1993). Vasopressinergic control of pituitary adrenocorticotropin secretion comes of age. Front Neuroendocrinol 14, 76-122.

Armbruster, B.N., Li, X., Pausch, M.H., Herlitze, S., and Roth, B.L. (2007). Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc Natl Acad Sci U S A 104, 5163-5168.

Basco, D., Zhang, Q., Salehi, A., Tarasov, A., Dolci, W., Herrera, P., Spiliotis, I., Berney, X., Tarussio, D., Rorsman, P., et al. (2018). alpha-cell glucokinase suppresses glucose-regulated glucagon secretion. Nat Commun 9, 546.

Baylis, P.H., and Heath, D.A. (1977). Plasma-arginine-vasopressin response to insulin-induced hypoglycaemia. Lancet 2, 428-430.

Baylis, P.H., and Robertson, G.L. (1980). Rat vasopressin response to insulin-induced hypoglycemia. Endocrinology 107, 1975-1979.

Berg, J.M.T., J. L.; Stryer, L. (2002). Each Organ Has a Unique Metabolic Profile. In Biochemistry. 5th edition. (New York: W H Freeman).

Bourque, C.W. (2008). Central mechanisms of osmosensation and systemic osmoregulation. Nat Rev Neurosci 9, 519-531.

Briant, L.J.B., Reinbothe, T.M., Spiliotis, I., Miranda, C., Rodriguez, B., and Rorsman, P. (2018). delta-cells and beta-cells are electrically coupled and regulate alpha-cell activity via somatostatin. J Physiol 596, 197-215.

Buchanan, T.A., Cane, P., Eng, C.C., Sipos, G.F., and Lee, C. (1991). Hypothermia Is Critical for Survival during Prolonged Insulin-Induced Hypoglycemia in Rats. Metabolism-Clinical and Experimental 40, 330-334.

109

Burdakov, D., Gerasimenko, O., and Verkhratsky, A. (2005). Physiological changes in glucose differentially modulate the excitability of hypothalamic melanin-concentrating hormone and orexin neurons in situ. The Journal of neuroscience : the official journal of the Society for Neuroscience 25, 2429-2433.

Christensen, M., Calanna, S., Sparre-Ulrich, A.H., Kristensen, P.L., Rosenkilde, M.M., Faber, J., Purrello, F., van Hall, G., Holst, J.J., Vilsboll, T., et al. (2015). Glucose-dependent insulinotropic polypeptide augments glucagon responses to hypoglycemia in type 1 diabetes. Diabetes 64, 72- 78.

Christensen, M., Vedtofte, L., Holst, J.J., Vilsboll, T., and Knop, F.K. (2011). Glucose- dependent insulinotropic polypeptide: a bifunctional glucose-dependent regulator of glucagon and insulin secretion in humans. Diabetes 60, 3103-3109.

Christensen, N.J. (1974). Plasma norepinephrine and epinephrine in untreated diabetics, during fasting and after insulin administration. Diabetes 23, 1-8.

Cryer, P.E. (1994). Banting Lecture. Hypoglycemia: the limiting factor in the management of IDDM. Diabetes 43, 1378-1389.

Cryer, P.E. (2002). Hypoglycaemia: the limiting factor in the glycaemic management of Type I and Type II diabetes. Diabetologia 45, 937-948.

Cryer, P.E. (2007). Hypoglycemia, functional brain failure, and brain death. J Clin Invest 117, 868-870.

Dadi, P.K., Luo, B., Vierra, N.C., and Jacobson, D.A. (2015). TASK-1 Potassium Channels Limit Pancreatic alpha-Cell Calcium Influx and Glucagon Secretion. Mol Endocrinol 29, 777- 787.

Fenske, W.K., Schnyder, I., Koch, G., Walti, C., Pfister, M., Kopp, P., Fassnacht, M., Strauss, K., and Christ-Crain, M. (2018). Release and Decay Kinetics of Copeptin vs AVP in Response to Osmotic Alterations in Healthy Volunteers. The Journal of clinical endocrinology and metabolism 103, 505-513.

Fisher, B.M., Baylis, P.H., and Frier, B.M. (1987). Plasma oxytocin, arginine vasopressin and atrial natriuretic peptide responses to insulin-induced hypoglycaemia in man. Clinical endocrinology 26, 179-185.

Freinkel, N., Metzger, B.E., Harris, E., Robinson, S., and Mager, M. (1972). The hypothermia of hypoglycemia. Studies with 2-deoxy-D-glucose in normal human subjects and mice. N Engl J Med 287, 841-845.

Fujiwara, Y., Hiroyama, M., Sanbe, A., Aoyagi, T., Birumachi, J., Yamauchi, J., Tsujimoto, G., and Tanoue, A. (2007). Insulin hypersensitivity in mice lacking the V1b vasopressin receptor. J Physiol 584, 235-244.

110

Garfield, A.S., Shah, B.P., Madara, J.C., Burke, L.K., Patterson, C.M., Flak, J., Neve, R.L., Evans, M.L., Lowell, B.B., Myers, M.G., Jr., et al. (2014). A parabrachial-hypothalamic cholecystokinin neurocircuit controls counterregulatory responses to hypoglycemia. Cell Metab 20, 1030-1037.

Grazzini, E., Lodboerer, A.M., Perez-Martin, A., Joubert, D., and Guillon, G. (1996). Molecular and functional characterization of V1b vasopressin receptor in rat adrenal medulla. Endocrinology 137, 3906-3914.

Gylfe, E. (2013). Glucose Control of Glucagon Secretion: There Is More to It Than K-ATP Channels. Diabetes 62, 1391-1393.

Hems, D.A., Whitton, P.D., and Ma, G.Y. (1975). Metabolic actions of vasopressin, glucagon and adrenalin in the intact rat. Biochimica et biophysica acta 411, 155-164.

Janaky, T., Laszlo, F.A., Sirokman, F., and Morgat, J.L. (1982). Biological half-life and organ distribution of [3H]8-arginine-vasopressin in the rat. The Journal of endocrinology 93, 295-303.

Kazda, C.M., Ding, Y., Kelly, R.P., Garhyan, P., Shi, C.X., Lim, C.N., Fu, H.D., Watson, D.E., Lewin, A.J., Landschulz, W.H., et al. (2016). Evaluation of Efficacy and Safety of the Glucagon Receptor Antagonist LY2409021 in Patients With Type 2 Diabetes: 12-and 24-Week Phase 2 Studies. Diabetes Care 2016;39:1241-1249. Diabetes Care 39, E199-E200.

Labouebe, G., Boutrel, B., Tarussio, D., and Thorens, B. (2016). Glucose-responsive neurons of the paraventricular thalamus control sucrose-seeking behavior. Nature neuroscience 19, 999- 1002.

Lai, B.K., Chae, H., Gomez-Ruiz, A., Cheng, P.P., Gallo, P., Antoine, N., Beauloye, C., Jonas, J.C., Seghers, V., Seino, S., et al. (2018). Somatostatin Is Only Partly Required for the Glucagonostatic Effect of Glucose but Is Necessary for the Glucagonostatic Effect of K-ATP Channel Blockers. Diabetes 67, 2239-2253.

Lamy, C.M., Sanno, H., Labouebe, G., Picard, A., Magnan, C., Chatton, J.Y., and Thorens, B. (2014). Hypoglycemia-activated GLUT2 neurons of the nucleus tractus solitarius stimulate vagal activity and glucagon secretion. Cell Metab 19, 527-538.

Lemon, G., Gibson, W.G., and Bennett, M.R. (2003). Metabotropic receptor activation, desensitization and sequestration-II: modelling the dynamics of the pleckstrin homology domain. J Theor Biol 223, 113-129.

Li, A.J., Wang, Q., and Ritter, S. (2018). Selective Pharmacogenetic Activation of Catecholamine Subgroups in the Ventrolateral Medulla Elicits Key Glucoregulatory Responses. Endocrinology 159, 341-355.

Li, Y.X., and Rinzel, J. (1994). Equations for InsP3 receptor-mediated [Ca2+]i oscillations derived from a detailed kinetic model: a Hodgkin-Huxley like formalism. J Theor Biol 166, 461- 473.

111

Lolait, S.J., Stewart, L.Q., Jessop, D.S., Young, W.S., 3rd, and O'Carroll, A.M. (2007). The hypothalamic-pituitary-adrenal axis response to stress in mice lacking functional vasopressin V1b receptors. Endocrinology 148, 849-856.

Madden, C.J., Stocker, S.D., and Sved, A.F. (2006). Attenuation of homeostatic responses to hypotension and glucoprivation after destruction of catecholaminergic rostral ventrolateral medulla neurons. Am J Physiol Regul Integr Comp Physiol 291, R751-759.

Mandelblat-Cerf, Y., Kim, A., Burgess, C.R., Subramanian, S., Tannous, B.A., Lowell, B.B., and Andermann, M.L. (2017). Bidirectional Anticipation of Future Osmotic Challenges by Vasopressin Neurons. Neuron 93, 57-65.

Marty, N., Dallaporta, M., Foretz, M., Emery, M., Tarussio, D., Bady, I., Binnert, C., Beermann, F., and Thorens, B. (2005). Regulation of glucagon secretion by glucose transporter type 2 (glut2) and astrocyte-dependent glucose sensors. J Clin Invest 115, 3545-3553.

Mawla, A.M., and Huising, M.O. (2019). Navigating the Depths and Avoiding the Shallows of Pancreatic Islet Cell Transcriptomes. Diabetes 68, 1380-1393.

McCrimmon, R.J., and Sherwin, R.S. (2010). Hypoglycemia in type 1 diabetes. Diabetes 59, 2333-2339.

Meek, T.H., Nelson, J.T., Matsen, M.E., Dorfman, M.D., Guyenet, S.J., Damian, V., Allison, M.B., Scarlett, J.M., Nguyen, H.T., Thaler, J.P., et al. (2016). Functional identification of a neurocircuit regulating blood glucose. Proc Natl Acad Sci U S A 113, E2073-2082.

Mergenthaler, P., Lindauer, U., Dienel, G.A., and Meisel, A. (2013). Sugar for the brain: the role of glucose in physiological and pathological brain function. Trends in Neurosciences 36, 587- 597.

Nica, A.C., Ongen, H., Irminger, J.C., Bosco, D., Berney, T., Antonarakis, S.E., Halban, P.A., and Dermitzakis, E.T. (2013). Cell-type, allelic, and genetic signatures in the human pancreatic beta cell transcriptome. Genome Res 23, 1554-1562.

Pei, H., Sutton, A.K., Burnett, K.H., Fuller, P.M., and Olson, D.P. (2014). AVP neurons in the paraventricular nucleus of the hypothalamus regulate feeding. Mol Metab 3, 209-215.

Petreanu, L., Huber, D., Sobczyk, A., and Svoboda, K. (2007). Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat Neurosci 10, 663-668.

Ritter, S., Dinh, T.T., and Zhang, Y. (2000). Localization of hindbrain glucoreceptive sites controlling food intake and blood glucose. Brain Res 856, 37-47.

Rorsman, P., Ramracheya, R., Rorsman, N.J., and Zhang, Q. (2014). ATP-regulated potassium channels and voltage-gated calcium channels in pancreatic alpha and beta cells: similar functions but reciprocal effects on secretion. Diabetologia 57, 1749-1761.

112

Salem, V., Silva, L.D., Suba, K., Georgiadou, E., Gharavy, S.N.M., Akhtar, N., Martin-Alonso, A., Gaboriau, D.C.A., Rothery, S.M., Stylianides, T., et al. (2019). Leader beta-cells coordinate Ca2+ dynamics across pancreatic islets in vivo. Nat Metab 1, 615-629.

Serradeil-Le Gal, C., Wagnon, J., Simiand, J., Griebel, G., Lacour, C., Guillon, G., Barberis, C., Brossard, G., Soubrie, P., Nisato, D., et al. (2002). Characterization of (2S,4R)-1-[5-chloro-1- [(2,4-dimethoxyphenyl)sulfonyl]-3-(2-methoxy-phenyl)-2-oxo- 2,3-dihydro-1H-indol-3-yl]-4- hydroxy-N,N-dimethyl-2-pyrrolidine carboxamide (SSR149415), a selective and orally active vasopressin V1b receptor antagonist. J Pharmacol Exp Ther 300, 1122-1130.

Siafarikas, A., Johnston, R.J., Bulsara, M.K., O'Leary, P., Jones, T.W., and Davis, E.A. (2012). Early loss of the glucagon response to hypoglycemia in adolescents with type 1 diabetes. Diabetes Care 35, 1757-1762.

Sloop, K.W., Cao, J.X., Siesky, A.M., Zhang, H.Y., Bodenmiller, D.M., Cox, A.L., Jacobs, S.J., Moyers, J.S., Owens, R.A., Showalter, A.D., et al. (2004). Hepatic and glucagon-like peptide-1- mediated reversal of diabetes by glucagon receptor antisense oligonucleotide inhibitors. J Clin Invest 113, 1571-1581.

Sprague, J.E., and Arbelaez, A.M. (2011). Glucose counterregulatory responses to hypoglycemia. Pediatr Endocrinol Rev 9, 463-473; quiz 474-465.

Spruce, B.A., McCulloch, A.J., Burd, J., Orskov, H., Heaton, A., Baylis, P.H., and Alberti, K.G. (1985). The effect of vasopressin infusion on glucose metabolism in man. Clinical endocrinology 22, 463-468.

Takasaki, J., Saito, T., Taniguchi, M., Kawasaki, T., Moritani, Y., Hayashi, K., and Kobori, M. (2004). A novel Galphaq/11-selective inhibitor. J Biol Chem 279, 47438-47445.

Taveau, C., Chollet, C., Bichet, D.G., Velho, G., Guillon, G., Corbani, M., Roussel, R., Bankir, L., Melander, O., and Bouby, N. (2017). Acute and chronic hyperglycemic effects of vasopressin in normal rats: involvement of V1A receptors. American journal of physiology. Endocrinology and metabolism 312, E127-E135.

Thorens, B. (2014). Neural regulation of pancreatic islet cell mass and function. Diabetes Obes Metab 16 Suppl 1, 87-95.

Unger, R.H., and Cherrington, A.D. (2012). Glucagonocentric restructuring of diabetes: a pathophysiologic and therapeutic makeover. J Clin Invest 122, 4-12.

Vergari, E., Knudsen, J.G., Ramracheya, R., Salehi, A., Zhang, Q., Adam, J., Asterholm, I.W., Benrick, A., Briant, L.J.B., Chibalina, M.V., et al. (2019). Insulin inhibits glucagon release by SGLT2-induced stimulation of somatostatin secretion. Nature Communications 10, 139.

Wang, R., Liu, X., Hentges, S.T., Dunn-Meynell, A.A., Levin, B.E., Wang, W., and Routh, V.H. (2004). The regulation of glucose-excited neurons in the hypothalamic arcuate nucleus by glucose and feeding-relevant peptides. Diabetes 53, 1959-1965.

113

Watts, A.G., and Boyle, C.N. (2010). The functional architecture of dehydration-anorexia. Physiol Behav 100, 472-477.

Wersinger, S.R., Ginns, E.I., O'Carroll, A.M., Lolait, S.J., and Young, W.S., 3rd (2002). Vasopressin V1b receptor knockout reduces aggressive behavior in male mice. Mol Psychiatry 7, 975-984.

Yu, Q., Shuai, H., Ahooghalandari, P., Gylfe, E., and Tengholm, A. (2019). Glucose controls glucagon secretion by directly modulating cAMP in alpha cells. Diabetologia.

Zhao, Z., Wang, L., Gao, W., Hu, F., Zhang, J., Ren, Y., Lin, R., Feng, Q., Cheng, M., Ju, D., et al. (2017). A Central Catecholaminergic Circuit Controls Blood Glucose Levels during Stress. Neuron 95, 138-152 e135.

114

Chapter 4

Conclusion

115

4.1 Summary of findings

This dissertation presents two independent studies investigating neural mechanism for regulation of AVP release. The first study focused on identifying and characterizing the neural circuit underlying anticipatory presystemic regulation of AVP release by water and food. The second study focused on understanding the neural and molecular pathways mediating hypoglycemia-induced AVP release.

Findings presented in this dissertation are summarized below (Figure 4.1):

1) Bidirectional anticipatory presystemic regulation of magnocellular AVP neurons by

water and food are mediated by two distinct circuits.

2) Water-related presystemic regulation is mediated by structures in the LT, including the

SFO, MnPO, and OVLT. The MnPO/OVLT functions as the main gateway by which

presystemic information enters the LT.

3) Food-related presystemic regulation is mediated by a yet unidentified population of

neurons in the ARC. AgRP and POMC neurons are not involved in this regulation.

4) A brainstem circuit involving A1/C1 neurons of the VLM mediates hypotension- and

hypoglycemia-induced activation of magnocellular AVP neurons. This pathway is not

involved in mediating activation of magnocellular AVP neurons by other stress-related

stimuli, including nausea and acute pain.

5) The hormone, AVP, serves a glucoregulatory role. Elevation of AVP causes an increase

in blood glucose levels.

116

Figure 4.1. Regulation of AVP release

117

6) The glucoregulatory role of AVP is mediated by V1bR, which is specifically expressed in

the glucagon-secreting alpha-cells of the pancreas. Binding of AVP to V1bR stimulates

glucagon release, which in turn stimulates hepatic glucose production to restore blood

glucose to normal level.

4.2 Function of presystemic regulation in Vasopressin neurons

Since the discovery of rapid, anticipatory responses in AgRP and thirst neurons (Betley et al., 2015; Chen et al., 2015; Mandelblat-Cerf; Zimmerman et al., 2016), multiple theories have been put forward to describe the function of presystemic regulation. One view states that presystemic regulation provides a negative-valence teaching signal to aid in food-cue association learning. This is based on the findings that both AgRP and thirst neurons have negative motivational valence and their transient silencing is rewarding (Allen et al., 2017; Betley et al.,

2015; Leib et al., 2017). Another view suggests that presystemic regulation allows transition from appetitive to consummatory phase. This is based on the fact that stimulation of AgRP neurons causes an extreme foraging behavior, which indicates that silencing of AgRP neurons might be required for animals to initiate feeding (Dietrich et al., 2015; Krashes et al., 2011).

Finally, third view focuses on homeostatic function of the circuit. It suggests that presystemic regulation prevents over/undershooting of homeostatic targets by terminating consummatory behavior preemeptively, taking into account the delay between consumption and its effect on systemic processes (Andermann and Lowell, 2017; Gizowski and Bourque, 2018; Lowell, 2019;

Stricker and Hoffmann, 2007).

118

Magnocellular AVP neurons are a homogeneous neuroendocrine population whose sole

function is to release AVP into the peripheral circulation. Unlike AgRP and thirst neurons, magnocellular AVP neurons have no known role in directly regulating behavior. There are several studies in the late 1900s suggesting behavioral effects of exogenous or artificially- induced AVP in promoting aversive learning and memory retention (Ettenberg et al., 1983; Koob et al., 1985; Le Moal et al., 1984). However, no recent studies have been performed to validate that this behavioral effect can be replicated with physiological level of endogenous peripheral

AVP. Another difference between magnocellular AVP neurons and AgRP and thirst neurons comes from their downstream neural circuitry. Magnocellular AVP neurons project directly to the posterior pituitary to release AVP into the systemic circulation. This is in stark contrast to

AgRP and thirst neurons, which send extensive projections to multiple downstream targets to regulate diverse behavioral and physiological responses to perturbations in energy and fluid balance. These major differences make it easier to interpret the role of presystemic regulation of

AVP neurons. It presumably solely functions to provide feedforward regulation of fluid homeostasis.

Activity of magnocellular AVP neurons directly determines the amount of AVP released into the circulation (Brown et al., 2013). The bidirectional nature of magnocellular AVP neurons’ presystemic response to food and water, which has not yet been observed in AgRP and thirst neurons, demonstrates how fluid balance can be tightly regulated under recurring hyper-and hypo-osmotic stresses. Rapid inhibition and activation of AVP neurons in response to water- predicting cues and feeding, respectively, adjusts the level of AVP release according to upcoming systemic osmolality changes. Delay between the onset of water and food consumption and their effect on systemic fluid balance is 15 min and 5 min, respectively (Figure 1.4)

119

(Mandelblat-Cerf et al., 2017). Failure to adjust AVP release presystemically during this period could lead to extreme hypo- or hyper-osmolality that can cause adverse side effects.

4.3 Significance of functionally distinct circuits in regulation of Vasopressin release

VP is a multifaceted hormone with diverse physiological functions (Koshimizu et al.,

2012). AVP, also known as an antidiuretic hormone, maintains fluid balance by regulating renal water excretion. AVP, when released in response to hypotension or hypovolemia, causes vasoconstriction to recover blood pressure. Our recent study also demonstrates that AVP is released in response to hypoglycemia and serves a glucoregulatory role by stimulating glucagon release (Chapter 3). Importance of precise regulation of AVP release is exemplified in two extreme cases of AVP diseases where deficiency (diabetes insipidus) (Christ-Crain, 2020) and excess (syndrome of inappropriate antidiuretic hormone secretion) (Miller, 2001) of AVP release results in deleterious conditions.

We demonstrated that SONAVP neurons are under the control of at least three

functionally-distinct neural circuits: 1) the LT conveys information about water-predicting cues,

water intake and systemic osmolality, 2) the arcuate nucleus conveys information about ongoing

food intake, and 3) A1/C1 neurons convey information about blood pressure and blood glucose.

These pathways are completely non-overlapping as blocking one pathway has no effect on the

SONAVP neuron response to cues, which use other circuits. With detailed cell type- and

projection-specific analysis of LT subpopulations, we demonstrated that each of the three LT

afferents contribute to different aspects of SONAVP neuron response to water: SON-projecting

MnPO/OVLTVglut2 provide water cue and drinking information, SON-projecting SFOVglut2

120

neurons provide systemic osmolality information, and SON-projecting MnPO/OVLTVgat neurons provide water cue information. Here, the MnPO/OVLT functions as a gateway, which integrates systemic osmolality information from the SFO and OVLT and presystemic information from extra LT structures. We identified multiple regions that provide direct synaptic inputs to the

SON-projecting MnPO/OVLTVglut2 neurons, including PVH, DMH, VMH, and LPBN, whose

involvement in presystemic regulation of magnocellular AVP neurons remains to be tested. We

also demonstrated that efferent projections of the LT are extensively collateralized, suggesting

that multiple downstream circuits are under simultaneous presystemic regulation. It should be

noted that the mechanism we describe here is yet far from complete. We showed that activation

of SONAVP neurons by nausea-inducing agent, LiCl, and acute injection-related pain cannot be accounted for by the circuits we have described.

Extensive and intricate circuit organization is a foundation for functional versatility of

AVP. Functionally-dedicated circuits ensure precise transmission of individual afferent signals to

magnocellular AVP neurons. Furthermore, the presence of functionally-relevant excitatory and

inhibitory neurons in a single pathway represents a powerful means for a wide range of

downstream modulation. We demonstrated that effects of water-predicting cue are mediated by

MnPO/OVLTVglut2 and MnPO/OVLTVgat neurons. Simultaneous decrease and increase in excitatory and inhibitory tone by water cues guarantees an effective silencing of SONAVP

neurons. The reverse is also true when hyperosmolality causes simultaneous activation of

MnPO/OVLTVglut2 and SFOVglut2 neurons, doubling the excitatory tone to SONAVP neurons to

achieve AVP release. Together, our findings demonstrate that AVP system is designed to be

highly effective and reliable such that every behavioral and physiological state of the animal is

precisely reflected in the final AVP output.

121

4.4 Neural mechanism for stress-induced Vasopressin release

Stress is a potent stimulus for AVP release. Stimulation of AVP release by aversive or

painful stimuli has been well-documented in multiple species including human. Nausea, caused

by motion sickness, migraine, or nausea-inducing drugs, results in a significant elevation of blood AVP level (Hampton et al., 1991; Koch et al., 1990; Verbalis et al., 1987). Pain related to surgical procedures, foot shock, heat, or subcutaneous formalin injection was also shown to cause a robust increase in blood AVP level (Bonjour and Malvin, 1970; Hamamura et al., 1984;

Kendler et al., 1978; Mirsky et al., 1954; Ueta et al., 2011).

Here, we performed the first real-time monitoring of magnocellular AVP neuron activity in response to injection of nausea-producing agent, lithium chloride (LiCl), and to acute pain induced by i.p. injection. Our data revealed a remarkable difference between the temporal dynamics of SONAVP neuron response to LiCl-induced nausea and acute pain, which have not been observed in previous studies due to lack of temporal resolution in blood AVP measurement.

The effect of acute pain on SONAVP neuron activity was extremely transient, lasting < 1 min. In contrast, effect of LiCl was slow-onset and long-lasting, suggesting progressive development of

nausea symptoms after LiCl injection. These results indicate that although SONAVP neurons are

responsive to a variety of aversive and painful stimuli, the resulting changes in SONAVP neuron

activity are highly variable depending on the nature of stress.

A neural mechanism for stress-induced activation of magnocellular AVP neurons is yet to be identified. We demonstrated that inhibition of A1/C1 neurons, which are involved in hypotension- and hypoglycemia-induced activation of magnocellular AVP neurons, has no effect

122

on LiCl- and acute pain-induced responses, implying that another dedicated circuit exists. The parabrachial nucleus (PBN) is a key brainstem structure for pain and aversion. The PBN receives

sensory information from the nucleus of the solitary tract and from the trigeminal and spinal

dorsal horns via spinobrachial tract, and regulates a wide range of behavioral and physiological responses to noxious stimuli via its downstream targets throughout the brain (Chiang et al., 2019;

Palmiter, 2018). Efferent projections of the PBN have been observed in the MnPO and PNZ

(Hermanson et al., 1998; Jhamandas et al., 1991), which are among the sites that provide direct

inputs to magnocellular AVP neurons. Direct innervation of magnocellular AVP neurons by the

PBN is unlikely as no PBN projection has been identified in to the SON. MnPO-projecting neurons of the PBN, characterized by expression of dynorphin, were shown to be activated by noxious stimuli (Hermanson et al., 1998), suggesting that pain-related signals mediated by PBN-

MnPO circuit might underlie stress-induced activation of magnocellular AVP neurons. PBN is

also known to contain neurons that transmit signals related to nausea and visceral malaise.

Injection of LiCl causes a marked increase in c-fos expression in the PBN, specifically in the

population that expresses calcitonin-gene-related peptide (CGRP) (Chen et al., 2018; Palmiter,

2018; St Andre et al., 2007). A number of evidences suggest that CGRP neurons might affect

magnocellular AVP neuron activity via indirect connection through the bed nuclei of the stria

terminalis (BNST). CGRP projections have been identified in the BNST (Palmiter, 2018). The

BNST sends projections to the PVH and SON, and pharmacological activation of the BNST causes a release of AVP (Crestani et al., 2013). Further studies examining the effect of PBN,

MnPO, BNST and PNZ inhibition on stress-induced activation of magnocellular AVP neurons

will provide clues to the neural mechanism underlying this response.

123

4.5 Lack of food-predicting cue response in magnocellular Vasopressin neurons

Presystemic regulation of magnocellular AVP neurons is characterized by bidirectional and asymmetric responses to water and food. Water-predicting cue and drinking cause a drop in magnocellular AVP neuron activity whereas feeding, but not food-predicting cue, causes an increase (Figure 4.1). Lack of response to food-predicting cue is striking, considering the rapid systemic effect of feeding (<5 min) compared to drinking (10-15 min) (Figure 1.4). An explanation for this unusual asymmetry can be found in the nature of osmotic stress provided by water and food and kinetics of AVP release and clearance.

Water and food have opposite effects on blood osmolality and AVP release. Water, a hypoosmotic stimulus, leads to decrease in AVP release. Food, a hyperosmotic stimulus, stimulates AVP release. Release of AVP happens immediately without any delay as it is highly dependent on the activity of magnocellular AVP neurons. AVP-containing secretory granules are positioned at the axon terminals in the posterior pituitary to ensure timely release of AVP upon activation of magnocellular AVP neurons (Brown et al., 2013). In contrast, degradation or clearance of AVP is not an instant process. It involves multiple enzymatic and molecular actions.

AVP clearance occurs in the kidney and liver. AVP is also subject to enzymatic degradation by vasopressinases that is present in the circulation (Sperling, 2008; Weitzman and Kleeman, 1979).

Although these processes initiate rapidly upon release of AVP, resulting in AVP’s rather short half-life of 5-12 min (Fenske et al., 2018; Janaky et al., 1982), process of AVP clearance still introduces a significant delay between the silencing of magnocellular AVP neurons and the effective reduction in circulating AVP.

124

The half-life of 5-12 min implies that complete silencing of magnocellular AVP neurons and cessation of AVP release are required to achieve 50% reduction from the original AVP level in 5-12 min. Considering that a significant decrease in systemic osmolality is observed 10-15

min after drinking, initiating the presystemic inhibition of magnocellular AVP neurons at the

onset of drinking is unlikely to ensure sufficient drop in AVP level before hypoosmolality

develops. For this reason, water-predicting cue represents an optimal time point for presystemic

inhibition to come into play. This is the earliest time point that availability of water and

consequent drinking-induced hypoosmolality can be precisely predicted. In contrast, early cue-

induced presystemic activation is not essential for feeding. Despite rapid systemic effect of food,

immediate release of AVP upon feeding-induced presystemic activation of magnocellular AVP

neurons guarantees that sufficiently high level of AVP is achieved before hyperosmolality

develops.

4.6 Technical consideration: rabies retrograde tracing

Development of monosynaptic retrograde rabies tracing technique has significantly

advanced our understanding of the neural circuitry. Monosynaptic rabies tracing is currently the

only method that allows identification monosynaptic inputs to molecularly-defined population of

neurons. However, rabies tracing possesses many caveats that should not be overlooked when

interpreting the result (Luo et al., 2018; Saleeba et al., 2019; Sun et al., 2019). A major concern

in our study was inefficiency and tropism of retrograde trans-synaptic labeling. To overcome this

problem, we implemented multiple validation steps, including antero- and retro-grade tracing,

ChR2-assisted circuit mapping (CRACM), and functional studies such as fiber photometry and

125

chemogenetic manipulation. During this process, we have encountered several unexpected

findings.

First, we found dense rabies labeling in the PVH when synaptic inputs to SONAVP

neurons were investigated using rabies tracing. The reverse was also true, suggesting strong

reciprocal connections between the PVH and SON. However, the SON is a homogeneous

structure that is comprised exclusively of AVP- and oxytocin-expressing magnocellular neurons.

Magnocellular neurons are known to send their axons only to the posterior pituitary, so it was unlikely that any neuron in the SON was providing direct input to the PVH. In order to evaluate

the synaptic connection between the PVH and SON, we performed a series of anterograde tracing, retrograde tracing with Cholera toxin subunit B (CTb), and CRACM experiments in the

PVH and SON (data not shown). We found no evidence indicating the direct connection between

the PVH and SON. This result lead us to conclude that unexpected rabies labeling we observed

in the PVH and SON in rabies tracing from magnocellular AVP neurons was due to non-trans-

synaptic spread of the rabies virus, mechanism of which is yet to be identified. Interestingly,

similar results have been obtained from the rabies tracing from magnocellular oxytocin neurons

(data not shown). This suggests that an unusual property of magnocellular neurons makes them

susceptible to unconventional non-trans-synaptic spread of the rabies virus.

Unlike typical neurons that release neurotransmitter into the confined synaptic space,

magnocellular neurons are capable of releasing massive amount of AVP and oxytocin directly

into the systemic circulation (Brown, 2016). An axon from single magnocellular neuron is

packed with 10 million neurosecretory vesicles containing hormones at near maximum capacity

(Leng and Ludwig, 2008). Each axon gives rise to ~2000 nerve terminals and swellings in the

126

posterior pituitary and releases up to ~10,000 vesicles per second upon stimulation. It is possible

that distinguishing structural feature of magnocellular neurons allows rabies to travel

anterogradely to the axon terminals at the posterior pituitary and to be retrogradely taken up by

other axon terminals nearby, which originate exclusively from magnocellular neurons in the

PVH and SON.

Another striking finding came from the same set of experiments looking at the direct inputs to magnocellular AVP neurons. A1/C1 neurons in the VLM are a well-known regulator of

AVP release. A1/C1 neurons send dense projections to the PVH and SON and provide noradrenergic/adrenergic input to magnocellular AVP neurons, respectively (Guyenet et al.,

2013). Using CRACM, we also demonstrated that A1/C1 neurons provide robust glutamatergic input to magnocellular AVP neurons. Despite strong evidences demonstrating direct synaptic connection between A1/C1 neurons and magnocellular AVP neurons, rabies tracing from

PVHAVP and SONAVP neurons did not identify any input in the VLM.

Although cellular mechanism of rabies trans-synaptic spread is not completely understood, it is believed that efficiency of rabies spread depends on the proximity of synapses to the starter neuron (Luo et al., 2018). It has been demonstrated that neuromodulatory synapses are especially inefficient in retrograde uptake of the rabies virus, presumably owing to the lack of tight junctions between the synapses (Wall et al., 2013). However, this is unlikely the case in

A1/C1 neurons because they are capable of releasing glutamate. It remains to be determined what properties of A1/C1 synapses make them significantly less susceptible to retrograde rabies infection than other glutamatergic synapses.

127

Together, these findings reemphasize the need to interpret rabies tracing data with caution. Without clear understanding of the tropism and trans-synaptic mechanism of the rabies virus, quantitative and qualitative interpretations of the rabies tracing data are unlikely to bear any biological significance. It is important that the rabies tracing is always complemented with additional tracing and electrophysiological approaches to avoid technical bias in the finding.

4.7 ARC population mediating food-related presystemic regulation

We demonstrated that non-specific inhibition of the ARC resulted in significant attenuation of food-related presystemic responses in AVP neurons. However, due to technical limitations, we were not able to identify a specific ARC cell type involved in this process. The most widely-studied, feeding-regulating populations of the ARC, AgRP and POMC neurons, were ruled out, as inhibition of these neurons had no effect on food-related presystemic response of SONAVP neurons. The ARC is comprised of multitude of molecularly-distinct cell types, of

which, only handful have been investigated.

Among functionally-defined cell types in the ARC, the most likely candidate for food-

related presystemic regulation of SONAVP neurons is (Oxtr)-expressing glutamatergic population (Fenselau et al., 2017). These neurons are orexigenic and are capable of stimulating food intake rapidly when activated. This is in stark contrast to POMC neurons, whose orexigenic effect is evident only after a long-term activation. Lack of delay in the behavioral effect, owing to their glutamatergic nature, makes the Oxtr-expressing ARC population a highly attractive candidate for mediating food-related presystemic regulation

SONAVP neurons. Further studies are needed to investigate their synaptic connections to SONAVP

128

neurons, their neural response to feeding, and involvement in food-related presystemic regulation

of SONAVP neurons.

4.8 Neural mechanism for hypoglycemia-induced Vasopressin release

Catecholaminergic neurons in the VLM are a key component of the central counter-

regulatory circuit. The ability of these neurons to evoke hyperglycemia (Li et al., 2018; Ritter et

al., 2000; Zhao et al., 2017) and hyperglucagonemia (Andrew et al., 2007) is well-established.

Recent studies have clearly demonstrated that spinally-projecting C1 neurons evoke hyperglycemia by stimulating the adrenal medulla (Li et al., 2018; Zhao et al., 2017). Here, we show that activation of A1/C1 neurons evokes hyperglycemia, but by promoting glucagon release through projections to the PVH and SON. A1/C1 neurons in the VLM have a well- documented ability to increase plasma AVP (Madden et al., 2006; Ross et al., 1984). Our study also demonstrated that A1/C1 neurons provide robust glutamatergic input to PVHAVP and

SONAVP neurons and are capable of activating SONAVP neurons in response to hypotension

(Chapter 2). Together, these data strongly support that the hyperglycemic and hyperglucagonemic effect of activating A1/C1 neurons is, at least in part, mediated by stimulation of AVP release. However, we recognize that A1/C1 neuron inhibition only partially prevented hypotension-induced activation of SONAVP neurons. It is not clear at this point

whether incomplete blockade of SONAVP neuron activation was due to technical reasons, such as inefficiency of hM4Di, or a presence of additional circuits mediating hypoglycemia information to SONAVP neurons.

129

Glucose-responsive neurons have been identified in multiple regions of the brain,

including the ARC, VMH, lateral hypothalamus, paraventricular thalamus, and brainstem

(Burdakov et al., 2005; Garfield et al., 2014; Labouebe et al., 2016; Lamy et al., 2014; Meek et

al., 2016; Wang et al., 2004). Several studies have shown that stimulating glucose-responsive

neurons causes an increase in blood glucose and glucagon, similar to the effect of SONAVP

neuron stimulation. It is worth investigating whether any of these glucose-responsive populations provides direct synaptic input to PVHAVP and SONAVP neurons and are involved in

hypoglycemia-induced AVP release.

Mechanism of hypoglycemia detection is still a controversial topic. A widely accepted view is that the activation of A1/C1 neurons and consequently magnocellular AVP neurons during hypoglycemia depends on peripheral glucose sensing at multiple sites, including the hepatic portal system (Marty et al., 2007; Verberne et al., 2014). The exact location of the glucose sensing is still controversial. Here, we show that activity of SONAVP neurons is increased at ~4.9 mM glucose. While it is unknown whether this threshold is sufficient to directly activate

A1/C1 neurons, vagal afferents in the hepatic portal vein – which convey vagal sensory information to the A1/C1 neurons via projections to the nucleus of the solitary tract (Verberne et al., 2014) - are responsive to alterations in glucose at this threshold (Niijima, 1982). Studies in isolated pancreatic islet demonstrated that blood glucose level of <4mM is required to trigger glucagon release via intrinsic mechanism (Walker et al., 2011). We therefore suggest that the activation of A1/C1 neurons and the subsequent increase in AVP release is critical for counter- regulatory glucagon release during the early stages of hypoglycemia (4-5 mM glucose), when intrinsic islet mechanisms are yet to be recruited.

130

There are studies suggesting direct neuro-stimulatory effect of insulin (Kleinridders et al.,

2014; Plum et al., 2005; Song et al., 2014). This is thought to be mediated by insulin receptors,

which are shown to be expressed in multiple areas of the brain, including the SON and brainstem

(Song et al., 2014; Unger et al., 1991). However, our study clearly demonstrates that the signal driving SONAVP neuron activation is unlikely a direct action of insulin on SONAVP or A1/C1 neurons because we observed a similar degree of SONAVP neuron activation regardless of whether hypoglycemia was induced by insulin or by 2-DG.

4.9 Relevance of findings to diabetes and diabetes complications

It is well-established that T1D patient show an early acquired abnormality of counter- regulation (McCrimmon and Sherwin, 2010). We also demonstrated that hypoglycemia fails to stimulate glucagon secretion in all T1D patients. More importantly, we found that insulin- induced copeptin secretion was compromised in T1D patients, with some subjects exhibiting no elevation at all in copeptin. This suggests that part of defective counter-regulatory glucagon secretion in T1D patients may be attributed to an insufficient hypoglycemia-induced copeptin response, strongly supporting our hypothesis.

Hypoglycemic attacks are the most common complication of insulin therapy in diabetic patients. T1D patients are especially vulnerable to exogenous insulin injection and develop more frequent and severe hypoglycemic attacks due to early-acquired defective counter-regulatory glucagon release. Frequent hypoglycemia attacks lead to further deterioration of counter- regulatory responses in T1D patients, reinforcing the vicious cycle. The mechanism is yet unknown but several hypotheses suggest that alteration in central glucose sensing or neural

131

circuits might underlie this phenomenon (Martin-Timon and Del Canizo-Gomez, 2015; Reno et al., 2013).

A wide range of central nervous system complications have been reported in the diabetic patients (Wrighten et al., 2009). We propose that impaired hypoglycemia-induced AVP release in T1D patients is caused by diabetes-induced changes in the brain. We postulate that recurrent hypoglycemia in patients with T1D might result in changes in glucose sensitivity of A1/C1 neurons or of other glucose-responsive neurons upstream of magnocellular AVP neurons. It is also possible that prolonged excessive stimulation of magnocellular AVP neurons during recurrent hypoglycemia causes a homeostatic plasticity, decreasing the sensitivity of magnocellular AVP neurons to drop in blood glucose. Future studies testing the effect of recurrent hypoglycemia on hypoglycemia-induced SONAVP neuron activation will provide

deeper mechanistic understanding of impaired counter-regulatory glucagon release in T1D.

132

References

Allen, W.E., DeNardo, L.A., Chen, M.Z., Liu, C.D., Loh, K.M., Fenno, L.E., Ramakrishnan, C., Deisseroth, K., and Luo, L. (2017). Thirst-associated preoptic neurons encode an aversive motivational drive. Science 357, 1149-1155.

Andermann, M.L., and Lowell, B.B. (2017). Toward a Wiring Diagram Understanding of Appetite Control. Neuron 95, 757-778.

Andrew, S.F., Dinh, T.T., and Ritter, S. (2007). Localized glucoprivation of hindbrain sites elicits corticosterone and glucagon secretion. Am J Physiol Regul Integr Comp Physiol 292, R1792-1798.

Betley, J.N., Xu, S., Cao, Z.F., Gong, R., Magnus, C.J., Yu, Y., and Sternson, S.M. (2015). Neurons for hunger and thirst transmit a negative-valence teaching signal. Nature.

Bonjour, J.P., and Malvin, R.L. (1970). Plasma concentrations of ADH in conscious and anesthetized dogs. The American journal of physiology 218, 1128-1132.

Brown, C.H. (2016). Magnocellular Neurons and Posterior Pituitary Function. Compr Physiol 6, 1701-1741.

Brown, C.H., Bains, J.S., Ludwig, M., and Stern, J.E. (2013). Physiological regulation of magnocellular neurosecretory cell activity: integration of intrinsic, local and afferent mechanisms. Journal of neuroendocrinology 25, 678-710.

Burdakov, D., Gerasimenko, O., and Verkhratsky, A. (2005). Physiological changes in glucose differentially modulate the excitability of hypothalamic melanin-concentrating hormone and orexin neurons in situ. The Journal of neuroscience : the official journal of the Society for Neuroscience 25, 2429-2433.

Chen, J.Y., Campos, C.A., Jarvie, B.C., and Palmiter, R.D. (2018). Parabrachial CGRP Neurons Establish and Sustain Aversive Taste Memories. Neuron 100, 891-899 e895.

Chen, Y., Lin, Y.C., Kuo, T.W., and Knight, Z.A. (2015). Sensory detection of food rapidly modulates arcuate feeding circuits. Cell 160, 829-841.

Chiang, M.C., Bowen, A., Schier, L.A., Tupone, D., Uddin, O., and Heinricher, M.M. (2019). Parabrachial Complex: A Hub for Pain and Aversion. The Journal of neuroscience : the official journal of the Society for Neuroscience 39, 8225-8230.

Christ-Crain, M. (2020). Diabetes insipidus - new concepts for diagnosis. Neuroendocrinology.

Crestani, C.C., Alves, F.H., Gomes, F.V., Resstel, L.B., Correa, F.M., and Herman, J.P. (2013). Mechanisms in the bed nucleus of the stria terminalis involved in control of autonomic and neuroendocrine functions: a review. Current neuropharmacology 11, 141-159.

133

Dietrich, M.O., Zimmer, M.R., Bober, J., and Horvath, T.L. (2015). Hypothalamic Agrp neurons drive stereotypic behaviors beyond feeding. Cell 160, 1222-1232.

Ettenberg, A., van der Kooy, D., Le Moal, M., Koob, G.F., and Bloom, F.E. (1983). Can aversive properties of (peripherally-injected) vasopressin account for its putative role in memory? Behavioural brain research 7, 331-350.

Fenselau, H., Campbell, J.N., Verstegen, A.M., Madara, J.C., Xu, J., Shah, B.P., Resch, J.M., Yang, Z., Mandelblat-Cerf, Y., Livneh, Y., et al. (2017). A rapidly acting glutamatergic ARC-- >PVH satiety circuit postsynaptically regulated by alpha-MSH. Nature neuroscience 20, 42-51.

Fenske, W.K., Schnyder, I., Koch, G., Walti, C., Pfister, M., Kopp, P., Fassnacht, M., Strauss, K., and Christ-Crain, M. (2018). Release and Decay Kinetics of Copeptin vs AVP in Response to Osmotic Alterations in Healthy Volunteers. The Journal of clinical endocrinology and metabolism 103, 505-513.

Garfield, A.S., Shah, B.P., Madara, J.C., Burke, L.K., Patterson, C.M., Flak, J., Neve, R.L., Evans, M.L., Lowell, B.B., Myers, M.G., Jr., et al. (2014). A parabrachial-hypothalamic cholecystokinin neurocircuit controls counterregulatory responses to hypoglycemia. Cell metabolism 20, 1030-1037.

Gizowski, C., and Bourque, C.W. (2018). The neural basis of homeostatic and anticipatory thirst. Nat Rev Nephrol 14, 11-25.

Guyenet, P.G., Stornetta, R.L., Bochorishvili, G., Depuy, S.D., Burke, P.G., and Abbott, S.B. (2013). C1 neurons: the body's EMTs. American journal of physiology Regulatory, integrative and comparative physiology 305, R187-204.

Hamamura, M., Shibuki, K., and Yagi, K. (1984). Noxious inputs to supraoptic neurosecretory cells in the rat. Neuroscience research 2, 49-61.

Hampton, K.K., Esack, A., Peatfield, R.C., and Grant, P.J. (1991). Elevation of Plasma Vasopressin in Spontaneous Migraine. Cephalalgia 11, 249-250.

Hermanson, O., Telkov, M., Geijer, T., Hallbeck, M., and Blomqvist, A. (1998). Preprodynorphin mRNA-expressing neurones in the rat parabrachial nucleus: subnuclear localization, hypothalamic projections and colocalization with noxious-evoked fos-like immunoreactivity. The European journal of neuroscience 10, 358-367.

Janaky, T., Laszlo, F.A., Sirokman, F., and Morgat, J.L. (1982). Biological half-life and organ distribution of [3H]8-arginine-vasopressin in the rat. The Journal of endocrinology 93, 295-303.

Jhamandas, J.H., Harris, K.H., and Krukoff, T.L. (1991). Parabrachial nucleus projection towards the hypothalamic supraoptic nucleus: electrophysiological and anatomical observations in the rat. The Journal of comparative neurology 308, 42-50.

134

Kendler, K.S., Weitzman, R.E., and Fisher, D.A. (1978). The effect of pain on plasma arginine vasopressin concentrations in man. Clinical endocrinology 8, 89-94.

Kleinridders, A., Ferris, H.A., Cai, W., and Kahn, C.R. (2014). Insulin action in brain regulates systemic metabolism and brain function. Diabetes 63, 2232-2243.

Koch, K.L., Summylong, J., Bingaman, S., Sperry, N., and Stern, R.M. (1990). Vasopressin and Oxytocin Responses to Illusory Self-Motion and Nausea in Man. J Clin Endocr Metab 71, 1269- 1275.

Koob, G.F., Dantzer, R., Rodriguez, F., Bloom, F.E., and Le Moal, M. (1985). Osmotic stress mimics effects of vasopressin on learned behaviour. Nature 315, 750-752.

Koshimizu, T.A., Nakamura, K., Egashira, N., Hiroyama, M., Nonoguchi, H., and Tanoue, A. (2012). Vasopressin V1a and V1b receptors: from molecules to physiological systems. Physiological reviews 92, 1813-1864.

Krashes, M.J., Koda, S., Ye, C., Rogan, S.C., Adams, A.C., Cusher, D.S., Maratos-Flier, E., Roth, B.L., and Lowell, B.B. (2011). Rapid, reversible activation of AgRP neurons drives feeding behavior in mice. The Journal of clinical investigation 121, 1424-1428.

Labouebe, G., Boutrel, B., Tarussio, D., and Thorens, B. (2016). Glucose-responsive neurons of the paraventricular thalamus control sucrose-seeking behavior. Nature neuroscience 19, 999- 1002.

Lamy, C.M., Sanno, H., Labouebe, G., Picard, A., Magnan, C., Chatton, J.Y., and Thorens, B. (2014). Hypoglycemia-activated GLUT2 neurons of the nucleus tractus solitarius stimulate vagal activity and glucagon secretion. Cell Metab 19, 527-538.

Le Moal, M., Dantzer, R., Mormede, P., Baduel, A., Lebrun, C., Ettenberg, A., van der Kooy, D., Wenger, J., Deyo, S., Koob, G.F., et al. (1984). Behavioral effects of peripheral administration of arginine vasopressin: a review of our search for a mode of action and a hypothesis. Psychoneuroendocrinology 9, 319-341.

Leib, D.E., Zimmerman, C.A., Poormoghaddam, A., Huey, E.L., Ahn, J.S., Lin, Y.C., Tan, C.L., Chen, Y.M., and Knight, Z.A. (2017). The Forebrain Thirst Circuit Drives Drinking through Negative Reinforcement. Neuron 96, 1272-+.

Leng, G., and Ludwig, M. (2008). Neurotransmitters and peptides: whispered secrets and public announcements. The Journal of physiology 586, 5625-5632.

Li, A.J., Wang, Q., and Ritter, S. (2018). Selective Pharmacogenetic Activation of Catecholamine Subgroups in the Ventrolateral Medulla Elicits Key Glucoregulatory Responses. Endocrinology 159, 341-355.

Lowell, B.B. (2019). New Neuroscience of Homeostasis and Drives for Food, Water, and Salt. The New England journal of medicine 380, 459-471.

135

Luo, L., Callaway, E.M., and Svoboda, K. (2018). Genetic Dissection of Neural Circuits: A Decade of Progress. Neuron 98, 865.

Madden, C.J., Stocker, S.D., and Sved, A.F. (2006). Attenuation of homeostatic responses to hypotension and glucoprivation after destruction of catecholaminergic rostral ventrolateral medulla neurons. Am J Physiol Regul Integr Comp Physiol 291, R751-759.

Mandelblat-Cerf, Y., Kim, A., Burgess, C.R., Subramanian, S., Tannous, B.A., Lowell, B.B., and Andermann, M.L. (2017). Bidirectional Anticipation of Future Osmotic Challenges by Vasopressin Neurons. Neuron 93, 57-65.

Mandelblat-Cerf, Y., Ramesh, R.N., Burgess, C.R., Patella, P, Lowell, B.B., Andermann, M.L. Arcuate hypothalamic AgRP and putative POMC neurons show opposite changes in spiking across multiple timescales. In Review, 19-02-2015-RA-eLife-07122.

Martin-Timon, I., and Del Canizo-Gomez, F.J. (2015). Mechanisms of hypoglycemia unawareness and implications in diabetic patients. World J Diabetes 6, 912-926.

Marty, N., Dallaporta, M., and Thorens, B. (2007). Brain glucose sensing, counterregulation, and energy homeostasis. Physiology (Bethesda) 22, 241-251.

McCrimmon, R.J., and Sherwin, R.S. (2010). Hypoglycemia in type 1 diabetes. Diabetes 59, 2333-2339.

Meek, T.H., Nelson, J.T., Matsen, M.E., Dorfman, M.D., Guyenet, S.J., Damian, V., Allison, M.B., Scarlett, J.M., Nguyen, H.T., Thaler, J.P., et al. (2016). Functional identification of a neurocircuit regulating blood glucose. Proc Natl Acad Sci U S A 113, E2073-2082.

Miller, M. (2001). Syndromes of excess antidiuretic hormone release. Crit Care Clin 17, 11-23, v.

Mirsky, I.A., Stein, M., and Paulisch, G. (1954). The Secretion of an Antidiuretic Substance into the Circulation of Adrenalectomized and Hypophysectomized Rats Exposed to Noxious Stimuli. Endocrinology 55, 28-39.

Niijima, A. (1982). Glucose-sensitive afferent nerve fibres in the hepatic branch of the vagus nerve in the guinea-pig. J Physiol 332, 315-323.

Palmiter, R.D. (2018). The Parabrachial Nucleus: CGRP Neurons Function as a General Alarm. Trends in neurosciences 41, 280-293.

Plum, L., Schubert, M., and Bruning, J.C. (2005). The role of insulin receptor signaling in the brain. Trends in endocrinology and metabolism: TEM 16, 59-65.

Reno, C.M., Litvin, M., Clark, A.L., and Fisher, S.J. (2013). Defective counterregulation and hypoglycemia unawareness in diabetes: mechanisms and emerging treatments. Endocrinol Metab Clin North Am 42, 15-38.

136

Ritter, S., Dinh, T.T., and Zhang, Y. (2000). Localization of hindbrain glucoreceptive sites controlling food intake and blood glucose. Brain Res 856, 37-47.

Ross, C.A., Ruggiero, D.A., Park, D.H., Joh, T.H., Sved, A.F., Fernandez-Pardal, J., Saavedra, J.M., and Reis, D.J. (1984). Tonic vasomotor control by the rostral ventrolateral medulla: effect of electrical or chemical stimulation of the area containing C1 adrenaline neurons on arterial pressure, heart rate, and plasma catecholamines and vasopressin. J Neurosci 4, 474-494.

Saleeba, C., Dempsey, B., Le, S., Goodchild, A., and McMullan, S. (2019). A Student's Guide to Neural Circuit Tracing. Frontiers in neuroscience 13, 897.

Song, Z.L., Levin, B.E., Stevens, W., and Sladek, C.D. (2014). Supraoptic oxytocin and vasopressin neurons function as glucose and metabolic sensors. Am J Physiol-Reg I 306, R447- R456.

Sperling, M. (2008). Pediatric endocrinology, 3rd edn (Philadelphia, PA: Saunders/Elsevier).

St Andre, J., Albanos, K., and Reilly, S. (2007). C-fos expression in the rat brain following lithium chloride-induced illness. Brain research 1135, 122-128.

Stricker, E.M., and Hoffmann, M.L. (2007). Presystemic signals in the control of thirst, salt appetite, and vasopressin secretion. Physiology & behavior 91, 404-412.

Sun, L.Q., Tang, Y.J., Yan, K.J., Yu, J.S., Zou, Y.Y., Xu, W.Z., Xiao, K., Zhang, Z.H., Li, W.M., Wu, B.L., et al. (2019). Differences in neurotropism and neurotoxicity among retrograde viral tracers. Molecular neurodegeneration 14.

Ueta, Y., Dayanithi, G., and Fujihara, H. (2011). Hypothalamic vasopressin response to stress and various physiological stimuli: Visualization in transgenic animal models. Hormones and behavior 59, 221-226.

Unger, J.W., Moss, A.M., and Livingston, J.N. (1991). Immunohistochemical localization of insulin receptors and phosphotyrosine in the brainstem of the adult rat. Neuroscience 42, 853- 861.

Verbalis, J.G., Richardson, D.W., and Stricker, E.M. (1987). Vasopressin release in response to nausea-producing agents and cholecystokinin in monkeys. The American journal of physiology 252, R749-753.

Verberne, A.J.M., Sabetghadam, A., and Korim, W.S. (2014). Neural pathways that control the glucose counterregulatory response. Front Neurosci-Switz 8.

Walker, J.N., Ramracheya, R., Zhang, Q., Johnson, P.R., Braun, M., and Rorsman, P. (2011). Regulation of glucagon secretion by glucose: paracrine, intrinsic or both? Diabetes, obesity & metabolism 13 Suppl 1, 95-105.

137

Wall, N.R., De La Parra, M., Callaway, E.M., and Kreitzer, A.C. (2013). Differential innervation of direct- and indirect-pathway striatal projection neurons. Neuron 79, 347-360.

Wang, R., Liu, X., Hentges, S.T., Dunn-Meynell, A.A., Levin, B.E., Wang, W., and Routh, V.H. (2004). The regulation of glucose-excited neurons in the hypothalamic arcuate nucleus by glucose and feeding-relevant peptides. Diabetes 53, 1959-1965.

Weitzman, R.E., and Kleeman, C.R. (1979). The clinical physiology of water metabolism. Part I: The physiologic regulation of arginine vasopressin secretion and thirst. West J Med 131, 373- 400.

Wrighten, S.A., Piroli, G.G., Grillo, C.A., and Reagan, L.P. (2009). A look inside the diabetic brain: Contributors to diabetes-induced brain aging. Biochimica et biophysica acta 1792, 444- 453.

Zhao, Z., Wang, L., Gao, W., Hu, F., Zhang, J., Ren, Y., Lin, R., Feng, Q., Cheng, M., Ju, D., et al. (2017). A Central Catecholaminergic Circuit Controls Blood Glucose Levels during Stress. Neuron 95, 138-152 e135.

Zimmerman, C.A., Lin, Y.C., Leib, D.E., Guo, L., Huey, E.L., Daly, G.E., Chen, Y., and Knight, Z.A. (2016). Thirst neurons anticipate the homeostatic consequences of eating and drinking. Nature 537, 680-684.

138