Skip to main content
Endocrinology logoLink to Endocrinology
. 2021 Apr 23;162(8):bqab082. doi: 10.1210/endocr/bqab082

Cardiovascular Neuroendocrinology: Emerging Role for Neurohypophyseal Hormones in Pathophysiology

Ato O Aikins 1,2, Dianna H Nguyen 1,2,2, Obed Paundralingga 1, George E Farmer 1, Caroline Gusson Shimoura 1, Courtney Brock 1, J Thomas Cunningham 1,
PMCID: PMC8234498  PMID: 33891015

Abstract

Arginine vasopressin (AVP) and oxytocin (OXY) are released by magnocellular neurosecretory cells that project to the posterior pituitary. While AVP and OXY currently receive more attention for their contributions to affiliative behavior, this mini-review discusses their roles in cardiovascular function broadly defined to include indirect effects that influence cardiovascular function. The traditional view is that neither AVP nor OXY contributes to basal cardiovascular function, although some recent studies suggest that this position might be re-evaluated. More evidence indicates that adaptations and neuroplasticity of AVP and OXY neurons contribute to cardiovascular pathophysiology.

Keywords: vasopressin, oxytocin, hypertension, heart failure


The hypothalamo-neurohypophyseal system (HNS) consists of magnocellular neurosecretory cells (MNCs) located in the supraoptic nucleus (SON) and paraventricular nucleus (PVN) of the hypothalamus (1). These neurons, along with accessory neurosecretory cells in the lateral hypothalamus, project to the posterior pituitary. From the posterior pituitary, the axon terminals of these neurons release the peptide hormones into the peripheral circulation (Fig. 1). Early studies of pituitary extracts demonstrated that they affected blood pressure (2, 3), urine excretion (4), parturition (5), and lactation (6). These classical endocrine effects were the focus of how the 2 hormones, arginine vasopressin (AVP) and oxytocin (OXY), were studied until much of the field shifted toward their behavior-associated actions as neuromodulators (7–9).

Figure 1.

Figure 1.

Magnocellular neurosecretory cells located in the supraoptic (SON) and paraventricular (PVN) nuclei of the hypothalamus project to the posterior pituitary where arginine vasopressin (AVP) and oxytocin (OXY) are released into the peripheral circulation. AVP acts on the kidneys to produce antidiuresis and on blood vessels to increase blood pressure. OXY produces uterine contractions that facilitate parturition and stimulates the milk ejection reflex. (Created with BioRender.com).

Among these classical functions, the cardiovascular effects of AVP and OXY have not been as widely investigated as their other actions. Share (3) suggested that there were several reasons why this is true for AVP. One of the foremost reasons is that hypothalamic diabetes insipidus is not normally associated with hypotension. Similarly, patients with the syndrome of inappropriate antidiuretic hormone secretion do not typically present with hypertension. Therefore, the physiological relevance of the pressor effects of vasopressin has been questioned (10, 11). Similarly, OXY appears to influence cardiovascular function more as a neurotransmitter in the central nervous system than as a hormone (12, 13). Recent studies suggest that, instead of normal physiology, AVP and OXY could be more involved in homeostasis and pathophysiology of the cardiovascular system.

AVP and Blood Pressure Regulation/Cardiovascular Function

AVP contributes to blood pressure homeostasis, and its release is stimulated by increased blood osmolality and decreased blood volume that occurs as a result of dehydration, hemorrhage, vasodilatory shock, and other physiological challenges (14–17). Changes in plasma osmolality and/or sodium are detected by osmoreceptors located in the circumventricular organs which activate MNCs located in the PVN and SON and by the MNCs themselves which are intrinsically osmosensitive (16). The effects of AVP are mediated by G protein–coupled receptors, with there being 2 major subtypes of AVP receptors: (1) V1 receptors, further subdivided into V1a and V1b, mediating vasoconstriction and corticotrophin secretion, respectively, and (2) V2 receptors, involved in fluid reabsorption by the collecting duct of the kidneys (18–21). V1a receptor–deficient (V1aR–/–) rats are hypotensive, have impaired arterial baroreceptor reflex activity, decreased blood volume, and decreased sympathetic nerve activity (22), and AVP-deficient rats have blunted blood pressure recovery following hemorrhage (23). Terlipressin (analog of AVP) and Selepressin (V1a receptor agonist) are being studied for their use in treating vasodilatory shock as AVP has become an important therapeutic target for this disorder (15, 24, 25).

Baroreceptors are primary sensory neurons that innervate the carotid sinus and the aortic arch and are activated by increases in blood pressure (26). Baroreflex stimulation of the vagus (X) and glossopharyngeal (IX) nerves activates the nucleus of the solitary tract (NTS), which distributes this information to the caudal ventrolateral medulla, nucleus ambiguus, and the hypothalamus (26, 27). In addition to autonomic regulation, early studies indicated baroreceptors regulate AVP release since baroreceptors denervation produced an acute elevation in circulating levels of AVP (28, 29). Furthermore, vagotomy, removal of aortic and cardiopulmonary baroreceptors, and carotid occlusion also increase circulating levels of AVP (3, 30, 31). Acute activation of baroreceptors inhibits the activity of putative AVP neurons, and these inhibitory effects are mediated by the diagonal band of Broca (DBB) (32–34) and the perinuclear zone (PNZ) (35, 36). Therefore, the stimulation of peripheral baroreceptors activates the DBB that projects to PNZ, a region adjacent to the SON, inhibiting the activity of AVP neurons. The inhibitory effects of baroreceptors on SON neurons seem to be mediated by the neurotransmitter gamma aminobutyric acid (GABA) (32). The blockade of GABA receptors at the SON or PVN prevents the inhibition of AVP neurons when the baroreceptors are stimulated (32, 37, 38). Furthermore, noradrenergic neurons in the locus coeruleus may also be involved in the baroreflex activation of the DBB (39). The A1 region of the ventrolateral medulla contributes to AVP release stimulated by baroreceptor unloading or the activation of low volume receptors (17).

Humoral signals can also stimulate AVP release, resulting in acute or chronic changes in blood pressure. Chronic osmotic stimulation associated with either deoxycorticosterone acetate (DOCA) salt hypertension or salt loading activates AVP neurons leading to increased blood pressure in male rats (40, 41). Increased activation of AVP neurons also contributes to hypertension in rats given angiotensin II (Ang II) and salt loading (42). Circulating Ang II acts at circumventricular organs (eg, the subfornical organ) and regulates blood pressure by multiple mechanisms, including AVP release (43). Recent studies using sophisticated genetic approaches have demonstrated that a brain angiotensin system may also contribute to AVP release and blood pressure regulation (44, 45). Together these studies demonstrate that neural and humoral signals related to cardiovascular function can alter AVP release in physiological and extraphysiological contexts.

In the central nervous system, AVP neurons from parvocellular PVN project to sympathetic premotor neurons in the rostral ventrolateral medulla and sympathetic preganglionic neurons in the spinal cord (26). Intrathecal injections of AVP increase blood pressure and sympathetic nerve activity (46–48). Studies using vasopressin-deficient Brattleboro rats indicate that, although AVP can contribute to sympathetic regulation, its absence can be compensated for by glutamate, and AVP may contribute to long-term regulation of blood pressure as a neurotransmitter (48). More recently, the dendritic release of AVP in the PVN has been shown to influence sympathetic nerve activity and blood pressure (49). In male rats, V1 receptors in the PVN contribute to hypertension associated with chronic stress (50). These observations are consistent with AVP contributing to long-term cardiovascular control and pathophysiology.

OXY and Blood Pressure Regulation/Cardiovascular Function

In addition to the brain, OXY is synthesized by the heart and large vessels such as the aorta and vena cava and mediates chronotropic effects on the heart and influences vascular tone (51). There is only 1 OXY receptor, which is coupled to Gq/11 protein and can be found in mammary and uterine smooth muscle, heart, kidney, brain, and vasculature (18–21). Despite that, OXY can also bind to AVP V1a receptors, causing vasoconstriction (52). Like AVP, OXY may not directly contribute to maintaining blood pressure in normal conditions. Some of the cardiovascular effects of OXY are indirect, mediated by its effects on atrial natriuretic peptide (ANP) release and nitric oxide (53, 54). ANP is a potent diuretic and natriuretic hormone isolated from the heart that affects blood pressure and electrolyte homeostasis (55). OXY increases plasma ANP and sodium excretion in rats (56, 57). Genetically modified animals that have increased plasma ANP are hypotensive, while ANP or OXY knockout mice have increased blood pressure (58–60). In rats, hypophysectomy completely inhibits ANP responses to blood volume expansion, indicating a link between HNS and ANP release (61). In the brain, OXY can act at the locus coeruleus, NTS, and dorsal motor nucleus of the vagal nerve, decreasing blood pressure and heart rate, possibly via α2-adrenoreceptors (12, 62, 63).

AVP and OXY in Control of the Blood Volume and Natriuresis

AVP and OXY are central to the regulation of blood volume and electrolyte homeostasis. A decrease in blood volume is a major stimulus for AVP release, and it is mediated via baroreceptor and volume receptor pathways (64–66). In the rat, OXY increases urinary sodium excretion at physiological concentrations (67). However, this effect is species specific (68). Plasma volume expansion stimulates the release of OXY, while AVP secretion is inhibited to allow restoration of plasma volume to homeostatic levels. Circulating OXY has a secondary effect on natriuresis by stimulating the release of ANP from the atrium, which also increases renal sodium excretion (69, 70). Increased secretion of AVP in some pathophysiological states can lead to an increase in water retention resulting in dilutional hyponatremia (71, 72), affecting cellular functions and the maintenance of homeostasis (73). Studies have shown that in salt-loaded rats there is a failure of the inhibitory mechanism for AVP secretion (74). This has been linked to an increased expression of brain-derived neurotrophic factor (BDNF) in the SON MNCs, leading to the reversal of the chloride gradient (74, 75). It is possible that a similar mechanism is at play in other states of increased AVP release, such as liver disease, heart failure, and even pregnancy.

Sex Differences in AVP and OXY and Cardiovascular Function

While the general biology of AVP and OXY as hormones is well understood, little is known about the existence or implications of any potential sex differences with respect to their biology. Sex differences in AVP and OXY synthesis and projections in the brain were reviewed in detail by Dumais and Veenema (76). They noted that while AVP synthesis is often sexually dimorphic, the functional significance of such dimorphism is unclear. They also noted that while sex differences in OXY immunoreactivity and mRNA expression have been reported in rodent hypothalamic nuclei with females showing higher levels than males, it remains unknown whether these differences translate into differences in OXY release (76). In most species, there are no sex differences reported for AVP or OXY expression in the PVN or SON, regions important in HNS communication with the cardiovascular system. Additionally, where sex differences exist, they appear to be species specific (76–78). Overall, limited sex difference research exists for both AVP and OXY release.

With expanding research suggesting AVP and OXY having a more substantial influence on pathophysiology, it is beneficial to consider the influence of sex in disease models. A few groups have reported sex differences in the regulation of AVP and OXY release that were either not mentioned in or published after Dumais and Veenema (76) which addressed this point. Results from early studies by Stone et al. suggested the influence of gonadal steroid hormones in adrenoreceptor-mediated control of AVP release where plasma AVP levels rose 4 times higher in the high estradiol phase (proestrus) of the estrous cycle when than low estradiol phases and males in response to intracerebroventricular (ICV) injection of norepinephrine (79). The same group later reported sex differences in central cholinergic and angiotensinergic mechanisms in the regulation of AVP release after finding plasma AVP concentration of male rats increased twice as much as female rats after ICV injection of the cholinergic agonist, carbachol; however, the increase in plasma AVP concentration of males was only half that of females after ICV injection of Ang II (80). OXY release has also been shown to be affected by gonadal steroid hormones. Scott et al. (81) revealed a tyrosine hydroxylase–expressing neuronal cluster in mice having monosynaptic inputs to OXY neurons in the PVN, which promoted OXY release in females but not in males. Other groups found sex differences in stress-induced OXY release in which short-term immobilization was used as a stressor resulting in increased plasma OXY levels in males, which was exacerbated after orchidectomy. Stress-induced increases in plasma OXY were also observed in females. However, that increase was unchanged when induced following gonadectomy in contrast to what was observed in the males (82–84). The acute immobilization used in these studies correlates to the human feelings of helplessness and is often used to model post-traumatic stress disorder (85). These studies could provide some insight into cardiovascular effects since sex differences in hypothalamic–pituitary–adrenal (HPA) axis response to stress have been linked to cardiovascular disease (86–88). Consistent with this theory, Dempster et al. report that children who experience adverse childhood experiences such as maltreatment and severe household dysfunction demonstrate altered HPA axis activity, which is associated with chronic changes in AVP and increased risk of chronic illness including cardiovascular disease (89). In regards to homeostatic stresses, a recent study from our group reported sex differences in the regulation of both AVP and OXY release in bile duct–ligated rats, a rodent model of hyponatremia associated with cirrhosis (90). Despite these reported findings, much remains to be uncovered regarding sex differences in the regulation of AVP and OXY release, particularly in the context of cardiovascular function.

AVP and OXY Receptors

To better understand the implications of changes to AVP and OXY regulation in cardiovascular function and disease, it is important to consider the expression of and main effects mediated by their receptors, which will be reviewed in the following sections based on location.

Cardiovascular System

In the heart, the V1a receptor is predominantly expressed. Located on the cardiac myocytes, V1a receptors contribute to the maintenance of physiological blood pressure by maintaining circulating blood volume and baroreflex sensitivity (91). While V1a receptors located in the heart play only a moderate role in the homeostatic maintenance of blood pressure, multiple animal models support the idea that cardiac V1a receptors play a significant role in the development and advancement of chronic heart failure (92, 93). In fact, cardiac V1a receptor expression more than doubles in patients with end-stage heart failure, making it a strong target candidate for the treatment of heart failure (94, 95).

V1a receptors are also expressed in vascular smooth muscle and blood vessels, affording them a significant role in regulating blood pressure (96). Here, the receptors mediate a vascular contractile effect in response to AVP (91). The contractile effect occurs via the phosphatidylinositol–bisphosphonate pathway, which leads to increases in intracellular calcium. Seemingly contradictory, both AVP and OXY can induce a vasodilatory effect in vascular endothelial cells through nitric oxide release (74), a result of OXY receptor activation (97, 98). OXY receptors have been identified in endothelial cells and are shown to have an equal affinity for both AVP and OXY.

Brain

The OXY receptor, a Gq-coupled receptor, has been identified in many different brain regions, including olfactory bulbs, anterior olfactory nucleus (AON), olfactory tubercle, nucleus accumbens, prelimbic cortex, ventral subiculum, central nucleus of the amygdala, ventromedial nucleus of the hypothalamus, cingulate cortex, dorsal motor nucleus of the vagus, nucleus tractus solitaries, among other areas. In the brain, OXY and V1 receptors (both V1a and V1b) are largely known to modulate social behaviors as opposed to normal cardiovascular function (76, 99–101). It is noteworthy though that brain V1 and OXY receptors have been implicated in various aspects of cardiovascular dysfunction or pathophysiology (eg, hypertension (102, 103), myocardial infarction (104, 105), cardiovascular response to stress (106, 107), and body fluid homeostasis (108)). Sex differences have been reported in OXY receptor and V1a receptor binding density with V1a receptors consistently higher in male rats than female rats and OXY receptor differences being region specific; however, the sex differences were less prominent than age differences (101). The OXY receptor gene contains a palindromic estrogen response element, allowing expression to be regulated by estrogens. Specifically, estrogen has been shown to increase OXY receptor expression in the male and female rat brains. According to that same study, AVP receptors in the brain are unaffected by gonadal steroid hormones (109).

Kidney

OXY receptors are present in the kidney, where they have been shown in rats to have a natriuretic response caused mainly by decreasing sodium reabsorption (67). Renal OXY receptor expression is thought to be, at least in part, influenced by estradiol and estrogen-induced water and solute reabsorption is mediated by renal OXY receptors (110, 111). The primary AVP receptor located in the kidney is the V2 receptor, and its mRNA expression has been shown to be significantly higher in female rats than in males (112). Consequently, female rats have potentially greater sensitivity for AVP-induced antidiuresis and susceptibility to blood volume disorders that can lead to other cardiovascular complications (112). A possible reason for the observed sex difference in V2 receptors is that the avpr2 gene is located in a region with a high chance of escaping inactivation on the X chromosome, which may lead to higher expression in females than males (113). The V2 receptor is important in maintaining normal fluid balance, and thus antagonistic drugs such as tolvaptan are the focus of treating hyponatremia associated with heart failure (114).

Plasticity

The activity of MNCs in the SON and PVN has been widely studied (115) and is influenced by both intrinsic and extrinsic mechanisms. Activation of these mechanisms can induce structural changes, not only in MNCs but in astrocyte morphology as well. Furthermore, synaptic inputs can induce synaptic plasticity and influence gene expression. Ultimately, these stimuli-induced changes in structure, synaptic plasticity, and gene expression in the MNCs regulate the release of AVP and OXY into the systemic circulation.

Structural Plasticity

The MNCs in the SON and PVN exhibit an intrinsic sensitivity to osmolality (116, 117). The osmosensitivity of the MNCs involves the activity of the mechanosensitive cation channel transient receptor potential vanilloid type 1 (TRPV1) (118, 119). In addition, MNCs express TRPV2 in the SON and PVN (120) and TRPV4 in the SON (120, 121), contributing to the inappropriate AVP release in a bile duct ligation model of hyponatremia. Furthermore, the expression of TRPV4 has been shown to be sensitive to ANG II (122). TRPV1 channels interact with structural proteins in the cells as part of their role in osmosensation. The osmosensitivity of MNCs involves the function of actin (123, 124) and microtubules (125). ANG II has been shown to influence actin density and enhance the response to osmotic challenges (126). Meanwhile, microtubules have been shown to interact with TRPV1 channels via β-tubulin (127, 128). Furthermore, the density of microtubules in the AVP cells is increased following salt loading (129).

Morphological plasticity influencing the function of the HNS is not limited to the MNCs and has also been studied in glial cells (130, 131). Retraction of astrocytic coverage observed in the SON and PVN during parturition, lactation, and dehydration can impact the synaptic function of MNCs in several ways (132, 133). This retraction can influence the clearance and diffusion of neurotransmitters as well as increase the availability of both glutamatergic and GABAergic synapses.

Synaptic Plasticity

Glutamatergic synapses in the SON are structurally changed following chronic salt loading (134, 135). Salt loading is associated with a decrease in NMDA receptor subunit expression (136) and altered Alpha-Amino-3-Hydroxy-5-Methyl-4-Isoxazole Propionic Acid (AMPA) subunit expression (137) that favors GluA1 and thus an increased calcium permeability. While this increased AMPA receptor–mediated calcium entry may induce excitotoxicity in other cell types, MNCs have been shown to have an increased resistance to neurotoxicity (138). Water deprivation is also associated with increases in phosphorylation of N-Methyl-D-Aspartic Acid (NMDA) receptor subunits (139).

Salt loading has been shown to alter GABA-mediated inhibition in the SON. Osmotic stress can change the structure of the GABA synapse (134, 140). Osmotic stress can alter chloride homeostasis, rendering GABA excitatory (40, 74, 141, 142). Increases in AVP release during salt loading are associated with an increase in BDNF in the SON (75), and BDNF appears to contribute to the changes in the valence of the GABA response (40, 74). Inappropriate AVP release associated with liver cirrhosis is also due in part to an upregulation of BDNF in male rats (143), while its effect on chloride transport remains to be determined. This effect of salt loading on AVP release contributes to a chronic increase in blood pressure (40). Similar changes in the function of AVP neurons have been reported in other models of hypertension, such as DOCA salt (41, 141), chronic ANG II (144), and Cyp1a1-Ren2 transgenic rats (145). The shift of GABA inhibition to excitation blunts the normal baroreceptor inhibition of AVP MNCs (145). Similar changes in GABA function have also been observed in presympathetic PVN neurons (146). The relevance of these mechanisms to human essential hypertension or heart disease has not been directly tested.

The epithelial sodium channel (ENaC) has been shown to regulate blood pressure through the reabsorption of Na+ at the distal portion of the nephron of the kidney (147). ENaCs have also been located in the AVP cells of the SON and PVN (148) and may play a role in salt-sensitive hypertension. ENaC in AVP neurons seems to contribute to a Na+ leak current at resting membrane potentials (149) but does not seem to influence the burst firing patterns characteristic of AVP cells (150). Interestingly, while high salt intake reduced the expression of ENaC in the kidney, high salt increases the expression of ENaC in the SON (150).

Repeated ICV injections of AVP produce sensitization of both its cardiovascular and behavioral effects (151, 152). Stress paradigms that cause central AVP release have also been shown to produce behavioral sensitization and cross-sensitization with corticotropin-releasing hormone (153). While this sensitization may not represent synaptic plasticity in a traditional sense, it is due to increased V1 receptor transduction (154). AVP and OXY were among the first peptides shown to be released from dendrites (155–157). Dendric release of AVP could sensitize V1 receptors in the SON and PVN, leading to potentiated cardiovascular responses.

Changes in Gene Expression

Using RNA sequencing methods, novel genes have been identified in the SON and PVN that are differentially regulated in response to sodium depletion (158), dehydration (159), and salt loading (159). This approach has identified a number of novel genes in the SON and PVN that are potentially related to body fluid homeostasis, such as creb3l1, a cAMP response element protein that has been shown to regulate avp transcription (160, 161).

A different approach has been used for another transcription factor that likely influences the function of MNCs. In the SON and PVN, 48 hours of water deprivation increased c-Fos and FosB expression. This increase could be reversed following 2 hours of rehydration (162, 163). Water deprivation has also been shown to increase the expression of a member of the AP-1 family of transcription factors, JunD, in the SON, with most of the JunD colocalizing with AVP neurons (164). Furthermore, c-Fos and FosB have been shown to be upregulated in the SON and PVN in response to volume expansion (165). ΔFosB was also increased in AVP and OXY cells in the SON following bile duct ligation, and blockade of the ΔFosB increase (using ΔJunD) attenuated the dilutional hyponatremia (166). The increased ΔFosB seen following bile duct ligation is not dependent on circulating ANG II acting at the subfornical organ (167). The genes regulated by ΔFosB have not yet been determined.

Both AVP and its receptors are regulated by epigenetic mechanisms. For example, increased methylation of CpG (cytosine-phosphate-guanine) sites is a common inhibitory epigenetic mechanism. DNA demethylation and de novo methylation of CpG residues in the proximal avp promoter occurs following dehydration but not salt loading in rat SON (168). Early-life stress in mice leads to epigenetic hypomethylation of a key regulatory region of the avp gene in the PVN. This modulation results in a persistent upregulation of avp expression in the PVN parvocellular division leading to a sustained hyperactivity of the HPA axis (169). Hypermethylation in the proximal promoter of the V1a receptor and protein kinase C isoform β (PKCβ) genes of the umbilical vein and placental vessels deactivates their transcription and ultimately leads to the attenuation of vasoconstriction responses to AVP in pre-eclamptic pregnancy (170, 171). This suggests that the epigenetic mechanisms that influence the expression of AVP and its receptors are determined by the physiological challenge applied to the system as well as the age of the organism.

Changes in Excitability

In animal models of heart failure and hypertension, increased activity of the MNCs that might be contributing to elevated circulating AVP has been demonstrated. In MNCs of rats with heart failure, the increased activity might be explained by reduced expression of the small conductance Ca2+-activated K+ (SK) channel and blunted coupling between SK channel and NMDAR (172, 173). The magnitude of medium spike after-hyperpolarization, a determinant of repetitive firing properties in AVP and OXY MNCs, depends on this SK channel. This channel is functionally coupled to NMDAR to form a negative feedback loop to NMDAR membrane depolarization.

In rats with renovascular hypertension, the activity of the MNCs is increased due to greater endogenous glutamate tone, which activates extrasynaptic NMDA receptors and tonically inhibits A-type K+ current (IA). (174). In the same model, NMDA receptor–mediated calcium signaling is also altered spatiotemporally to be more prolonged and larger than normal as the ER Ca2+ uptake mechanism is blunted, therefore contributing to the increased firing rate of MNCs (175). The activity of AVP and OXY MNCs can also be stimulated by prorenin through its prorenin receptor, an effect that also involves inhibition of A-type K+ current (IA) (176, 177). This prorenin effect will, in turn, elevate the circulating level of AVP in the DOCA salt hypertension model (178).

Summary

The participation of AVP and OXY in normal cardiovascular function remains controversial. This could be a function of the redundant mechanisms that regulate blood pressure. What is clearer is that activity-dependent changes in the function of AVP MNCs can negatively influence cardiovascular function and body fluid balance. Stress-induced plasticity has been demonstrated to facilitate adaptation in the HPA axis (179) and sensitize blood pressure regulatory systems (180, 181). Activity-dependent changes in cell morphology, synaptic strength, and gene expression in the SON and PVN come together to alter the excitability and firing patterns of AVP and OXY neurons to ultimately change the release of these neuropeptides into the systemic circulation to affect the peripheral cardiovascular function or the local parenchyma changing autonomic function (49). Changes in gene expression and possibly epigenetic mechanisms may further influence the excitatory/inhibitory balance of the MNCs or alter their responses to subsequent challenges. While this plasticity may normally function to help maintain homeostasis, it may contribute to cardiovascular disease (Fig. 2). The outcome of these processes may ultimately be determined by the sex of the individual. For example, AVP has been identified as contributing to the pathogenesis of pre-eclampsia (182) which is characterized by hypertension in late pregnancy.

Figure 2.

Figure 2.

Homeostatic challenges activate central pathways that induce plasticity in the MNCs of the SON and PVN. This plasticity influences the release of AVP and OXY both centrally and into systemic circulation and contributes to cardiovascular regulation. SON, supraoptic nucleus; mPVN, paraventricular nucleus magnocellular division; pPVN, paraventricular nucleus parvocellular division. The sex of the individual would influence each of these processes. (Created with BioRender.com).

Blocking AVP receptors improves fluid balance in patients with either liver disease or congestive heart failure (183–186), but these effects have not always translated into better clinical outcomes (184). As indicated above, the dendritic release of peptides from MNCs influences the activity of autonomic networks and possibly other neuroendocrine systems (49, 187). This suggests that plasticity in MNCs influences other physiological endpoints that impact cardiovascular function. In male rats, V1 receptors in the PVN contribute to hypertension associated with chronic mild stress paradigm (50). Activation of the V1 receptors associated with mild chronic stress in this paradigm could be due to dendritic release of AVP, and intermittent activation of these receptors could cause sensitization of the cardiovascular and endocrine responses that ultimately lead to hypertension. It is known that female rats tend to have greater neuroendocrine responses to acute stress than male rats, but sex differences in chronic stress are not as well characterized (88). Improving clinical outcomes might require a better understanding of how homeostatic mechanisms are converted into pathophysiology by neuroplasticity and how these processes are determined by the sex of the organism in addition to targeting receptors that regulate end-organ function.

Acknowledgments

Financial Support: This work was supported by the National Institutes of Health/ National Institute on Aging T32 grant AG020494 and National Institutes of Health/National Heart Lung and Blood Institute grants P01 088052, R01 HL142341, and R01 155977.

Glossary

Abbreviations

AON

anterior olfactory nucleus

AVP

arginine vasopressin

BDNF

brain-derived neurotrophic factor

DBB

diagonal band of Broca

ENaC

epithelial sodium channel

HNS

hypothalamo-neurohypophyseal system

HPA

hypothalamic–pituitary–adrenal

ICV

intracerebroventricular

MNC

magnocellular neurosecretory cell

OXY

oxytocin

PNZ

perinuclear zone

PVN

paraventricular nucleus

SON

supraoptic nucleus

Additional Information

Disclosures : The authors have nothing to disclose.

Data Availability

Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

  • 1. Burbach  JP, Luckman  SM, Murphy  D, Gainer  H. Gene regulation in the magnocellular hypothalamo-neurohypophysial system. Physiol Rev.  2001;81(3):1197-1267. [DOI] [PubMed] [Google Scholar]
  • 2. Oliver  G, Schäfer  EA. On the physiological action of extracts of pituitary body and certain other glandular organs: preliminary communication. J Physiol.  1895;18(3):277-279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Share  L. Role of vasopressin in cardiovascular regulation. Physiol Rev.  1988;68(4):1248-1284. [DOI] [PubMed] [Google Scholar]
  • 4. Farini  F. Diabete insipido ed opoterapia. Gazz Osped Clin. 1913;34:1135-1139. [Google Scholar]
  • 5. Dale  HH. On some physiological actions of ergot. J Physiol.  1906;34(3):163-206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Bell  WB. The pituitary body and the therapeutic value of the infundibular extract in shock, uterine atony, and intestinal paresis. Br Med J.  1909;2(2553):1609-1613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Neumann  ID, Landgraf  R. Balance of brain oxytocin and vasopressin: implications for anxiety, depression, and social behaviors. Trends Neurosci.  2012;35(11):649-659. [DOI] [PubMed] [Google Scholar]
  • 8. Walum  H, Young  LJ. The neural mechanisms and circuitry of the pair bond. Nat Rev Neurosci.  2018;19(11):643-654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Cilz  NI, Cymerblit-Sabba  A, Young  WS. Oxytocin and vasopressin in the rodent hippocampus. Genes Brain Behav.  2019;18(1):e12535. [DOI] [PubMed] [Google Scholar]
  • 10. Japundzic-Zigon  N. Effects of nonpeptide V1a and V2 antagonists on blood pressure fast oscillations in conscious rats. Clin Exp Hypertens.  2001;23(4):277-292. [DOI] [PubMed] [Google Scholar]
  • 11. Milutinović  S, Murphy  D, Japundzić-Zigon  N. The role of central vasopressin receptors in the modulation of autonomic cardiovascular controls: a spectral analysis study. Am J Physiol Regul Integr Comp Physiol.  2006;291(6):R1579-R1591. [DOI] [PubMed] [Google Scholar]
  • 12. Petersson  M. Cardiovascular effects of oxytocin. Prog Brain Res.  2002;139(1):281-288. [DOI] [PubMed] [Google Scholar]
  • 13. Sladek  CD, Michelini  LC, Stachenfeld  NS, Stern  JE, Urban  JH. Endocrine-autonomic linkages. Compr Physiol.  2015;5(3):1281-1323. [DOI] [PubMed] [Google Scholar]
  • 14. Prager-Khoutorsky  M, Choe  KY, Levi  DI, Bourque  CW. Role of vasopressin in rat models of salt-dependent hypertension. Curr Hypertens Rep.  2017;19(5):42. [DOI] [PubMed] [Google Scholar]
  • 15. Ukor  IF, Walley  KR. Vasopressin in vasodilatory shock. Crit Care Clin.  2019;35(2):247-261. [DOI] [PubMed] [Google Scholar]
  • 16. Bourque  CW. Central mechanisms of osmosensation and systemic osmoregulation. Nat Rev Neurosci.  2008;9(7):519-531. [DOI] [PubMed] [Google Scholar]
  • 17. Brown  CH. Magnocellular neurons and posterior pituitary function. Compr Physiol.  2016;6(4):1701-1741. [DOI] [PubMed] [Google Scholar]
  • 18. Manning  M, Stoev  S, Chini  B, Durroux  T, Mouillac  B, Guillon  G. Peptide and non-peptide agonists and antagonists for the vasopressin and oxytocin V1a, V1b, V2 and OT receptors: research tools and potential therapeutic agents. Prog Brain Res.  2008;170(1):473-512. [DOI] [PubMed] [Google Scholar]
  • 19. Zingg  HH. Vasopressin and oxytocin receptors. Baillieres Clin Endocrinol Metab.  1996;10(1):75-96. [DOI] [PubMed] [Google Scholar]
  • 20. Gimpl  G, Fahrenholz  F. The oxytocin receptor system: structure, function, and regulation. Physiol Rev.  2001;81(2):629-683. [DOI] [PubMed] [Google Scholar]
  • 21. Koshimizu  TA, Nakamura  K, Egashira  N, Hiroyama  M, Nonoguchi  H, Tanoue  A. Vasopressin V1a and V1b receptors: from molecules to physiological systems. Physiol Rev.  2012;92(4):1813-1864. [DOI] [PubMed] [Google Scholar]
  • 22. Aoyagi  T, Koshimizu  TA, Tanoue  A. Vasopressin regulation of blood pressure and volume: findings from V1a receptor-deficient mice. Kidney Int.  2009;76(10):1035-1039. [DOI] [PubMed] [Google Scholar]
  • 23. Zerbe  RL, Feuerstein  G, Meyer  DK, Kopin  IJ. Cardiovascular, sympathetic, and renin-angiotensin system responses to hemorrhage in vasopressin-deficient rats. Endocrinology.  1982;111(2):608-613. [DOI] [PubMed] [Google Scholar]
  • 24. Barrett  LK, Singer  M, Clapp  LH. Vasopressin: mechanisms of action on the vasculature in health and in septic shock. Crit Care Med.  2007;35(1):33-40. [DOI] [PubMed] [Google Scholar]
  • 25. Rodriguez  R, Cucci  M, Kane  S, Fernandez  E, Benken  S. Novel vasopressors in the treatment of vasodilatory shock: a systematic review of angiotensin II, selepressin, and terlipressin. J Intensive Care Med.  2020;35(4):327-337. [DOI] [PubMed] [Google Scholar]
  • 26. Guyenet  PG. The sympathetic control of blood pressure. Nat Rev Neurosci.  2006;7(5):335-346. [DOI] [PubMed] [Google Scholar]
  • 27. Dampney  RA. Central neural control of the cardiovascular system: current perspectives. Adv Physiol Educ.  2016;40(3):283-296. [DOI] [PubMed] [Google Scholar]
  • 28. Alexander  N, Morris  M. Increased plasma vasopressin in sinoaortic denervated rats. Neuroendocrinology.  1986;42(5):361-367. [DOI] [PubMed] [Google Scholar]
  • 29. Morris  M, Alexander  N. Baroreceptor influences on oxytocin and vasopressin secretion. Hypertension.  1989;13(2):110-114. [DOI] [PubMed] [Google Scholar]
  • 30. Thames  MD, Schmid  PG. Cardiopulmonary receptors with vagal afferents tonically inhibit ADH release in the dog. Am J Physiol.  1979;237(3):H299-H304. [DOI] [PubMed] [Google Scholar]
  • 31. Share  L, Levy  MN. Cardiovascular receptors and blood titer of antidiuretic hormone. Am J Physiol.  1962;203(3):425-428. [DOI] [PubMed] [Google Scholar]
  • 32. Jhamandas  JH, Renaud  LP. A gamma-aminobutyric-acid-mediated baroreceptor input to supraoptic vasopressin neurones in the rat. J Physiol.  1986;381:595-606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Jhamandas  JH, Renaud  LP. Diagonal band neurons may mediate arterial baroreceptor input to hypothalamic vasopressin-secreting neurons. Neurosci Lett.  1986;65(2):214-218. [DOI] [PubMed] [Google Scholar]
  • 34. Cunningham  JT, Nissen  R, Renaud  LP. Ibotenate lesions of the diagonal band of broca attenuate baroreceptor sensitivity of rat supraoptic vasopressin neurons. J Neuroendocrinol.  1992;4(3):303-309. [DOI] [PubMed] [Google Scholar]
  • 35. Jhamandas  JH, Raby  W, Rogers  J, Buijs  RM, Renaud  LP. Diagonal band projection towards the hypothalamic supraoptic nucleus: light and electron microscopic observations in the rat. J Comp Neurol.  1989;282(1):15-23. [DOI] [PubMed] [Google Scholar]
  • 36. Tribollet  E, Armstrong  WE, Dubois-Dauphin  M, Dreifuss  JJ. Extra-hypothalamic afferent inputs to the supraoptic nucleus area of the rat as determined by retrograde and anterograde tracing techniques. Neuroscience.  1985;15(1):135-148. [DOI] [PubMed] [Google Scholar]
  • 37. Jhamandas  JH, Renaud  LP. Bicuculline blocks an inhibitory baroreflex input to supraoptic vasopressin neurons. Am J Physiol.  1987;252(5 Pt 2):R947-R952. [DOI] [PubMed] [Google Scholar]
  • 38. Kasai  M, Osaka  T, Inenaga  K, Kannan  H, Yamashita  H. gamma-Aminobutyric acid antagonist blocks baroreceptor-activated inhibition of neurosecretory cells in the hypothalamic paraventricular nucleus of rats. Neurosci Lett.  1987;81(3):319-324. [DOI] [PubMed] [Google Scholar]
  • 39. Grindstaff  RJ, Grindstaff  RR, Sullivan  MJ, Cunningham  JT. Role of the locus ceruleus in baroreceptor regulation of supraoptic vasopressin neurons in the rat. Am J Physiol Regul Integr Comp Physiol.  2000;279(1):R306-R319. [DOI] [PubMed] [Google Scholar]
  • 40. Choe  KY, Han  SY, Gaub  P, et al.  High salt intake increases blood pressure via BDNF-mediated downregulation of KCC2 and impaired baroreflex inhibition of vasopressin neurons. Neuron.  2015;85(3):549-560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Kim  YB, Kim  YS, Kim  WB, et al.  GABAergic excitation of vasopressin neurons: possible mechanism underlying sodium-dependent hypertension. Circ Res.  2013;113(12):1296-1307. [DOI] [PubMed] [Google Scholar]
  • 42. Levi  DI, Wyrosdic  JC, Hicks  AI, et al.  High dietary salt amplifies osmoresponsiveness in vasopressin-releasing neurons. Cell Rep.  2021;34(11):108866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Cancelliere  NM, Black  EA, Ferguson  AV. Neurohumoral integration of cardiovascular function by the lamina terminalis. Curr Hypertens Rep.  2015;17(12):93. [DOI] [PubMed] [Google Scholar]
  • 44. Frazier  CJ, Harden  SW, Alleyne  AR, et al.  An angiotensin-responsive connection from the lamina terminalis to the paraventricular nucleus of the hypothalamus evokes vasopressin secretion to increase blood pressure in mice. J Neurosci.  2021;41(7):1429-1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Sandgren  JA, Linggonegoro  DW, Zhang  SY, et al.  Angiotensin AT1A receptors expressed in vasopressin-producing cells of the supraoptic nucleus contribute to osmotic control of vasopressin. Am J Physiol Regul Integr Comp Physiol.  2018;314(6):R770-R780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Porter  JP, Brody  MJ. Spinal vasopressin mechanisms of cardiovascular regulation. Am J Physiol.  1986;251(3 Pt 2):R510-R517. [DOI] [PubMed] [Google Scholar]
  • 47. Riphagen  CL, Pittman  QJ. Mechanisms underlying the cardiovascular responses to intrathecal vasopressin administration in rats. Can J Physiol Pharmacol.  1989;67(4):269-275. [DOI] [PubMed] [Google Scholar]
  • 48. Yang  Z, Coote  JH. Paraventricular nucleus influence on renal sympathetic activity in vasopressin gene-deleted rats. Exp Physiol.  2007;92(1):109-117. [DOI] [PubMed] [Google Scholar]
  • 49. Son  SJ, Filosa  JA, Potapenko  ES, et al.  Dendritic peptide release mediates interpopulation crosstalk between neurosecretory and preautonomic networks. Neuron.  2013;78(6):1036-1049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Komnenov  D, Quaal  H, Rossi  NF. V1a and V1b vasopressin receptors within the paraventricular nucleus contribute to hypertension in male rats exposed to chronic mild unpredictable stress. Am J Physiol Regul Integr Comp Physiol.  2021;320(3):R213-R225. [DOI] [PubMed] [Google Scholar]
  • 51. Gutkowska  J, Jankowski  M, Mukaddam-Daher  S, McCann  SM. Oxytocin is a cardiovascular hormone. Braz J Med Biol Res.  2000;33(6):625-633. [DOI] [PubMed] [Google Scholar]
  • 52. Birnbaumer  M. Vasopressin receptors. Trends Endocrinol Metab.  2000;11(10):406-410. [DOI] [PubMed] [Google Scholar]
  • 53. Gutkowska  J, Jankowski  M, Antunes-Rodrigues  J. The role of oxytocin in cardiovascular regulation. Braz J Med Biol Res.  2014;47(3):206-214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Gutkowska  J, Jankowski  M, Lambert  C, Mukaddam-Daher  S, Zingg  HH, McCann  SM. Oxytocin releases atrial natriuretic peptide by combining with oxytocin receptors in the heart. Proc Natl Acad Sci U S A.  1997;94(21):11704-11709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. de Bold  AJ, Borenstein  HB, Veress  AT, Sonnenberg  H. A rapid and potent natriuretic response to intravenous injection of atrial myocardial extract in rats. Life Sci.  1981;28(1):89-94. [DOI] [PubMed] [Google Scholar]
  • 56. Haanwinckel  MA, Elias  LK, Favaretto  AL, Gutkowska  J, McCann  SM, Antunes-Rodrigues  J. Oxytocin mediates atrial natriuretic peptide release and natriuresis after volume expansion in the rat. Proc Natl Acad Sci U S A.  1995;92(17):7902-7906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Soares  TJ, Coimbra  TM, Martins  AR, et al.  Atrial natriuretic peptide and oxytocin induce natriuresis by release of cGMP. Proc Natl Acad Sci U S A.  1999;96(1):278-283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. John  SW, Krege  JH, Oliver  PM, et al.  Genetic decreases in atrial natriuretic peptide and salt-sensitive hypertension. Science.  1995;267(5198):679-681. [DOI] [PubMed] [Google Scholar]
  • 59. Field  LJ, Veress  AT, Steinhelper  ME, Cochrane  K, Sonnenberg  H. Kidney function in ANF-transgenic mice: effect of blood volume expansion. Am J Physiol.  1991;260(1 Pt 2):R1-R5. [DOI] [PubMed] [Google Scholar]
  • 60. Bernatova  I, Rigatto  KV, Key  MP, Morris  M. Stress-induced pressor and corticosterone responses in oxytocin-deficient mice. Exp Physiol.  2004;89(5):549-557. [DOI] [PubMed] [Google Scholar]
  • 61. Antunes-Rodrigues  J, Ramalho  MJ, Reis  LC, et al.  Lesions of the hypothalamus and pituitary inhibit volume-expansion-induced release of atrial natriuretic peptide. Proc Natl Acad Sci U S A.  1991;88(7):2956-2960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Higa  KT, Mori  E, Viana  FF, Morris  M, Michelini  LC. Baroreflex control of heart rate by oxytocin in the solitary-vagal complex. Am J Physiol Regul Integr Comp Physiol.  2002;282(2):R537-R545. [DOI] [PubMed] [Google Scholar]
  • 63. Buemann  B, Uvnäs-Moberg  K. Oxytocin may have a therapeutical potential against cardiovascular disease. Possible pharmaceutical and behavioral approaches. Med Hypotheses.  2020;138(5):109597. [DOI] [PubMed] [Google Scholar]
  • 64. Jhamandas  JH, Renaud  LP. Neurophysiology of a central baroreceptor pathway projecting to hypothalamic vasopressin neurons. Can J Neurol Sci.  1987;14(1):17-24. [DOI] [PubMed] [Google Scholar]
  • 65. Quail  AW, Woods  RL, Korner  PI. Cardiac and arterial baroreceptor influences in release of vasopressin and renin during hemorrhage. Am J Physiol.  1987;252(6 Pt 2):H1120-H1126. [DOI] [PubMed] [Google Scholar]
  • 66. Callahan  MF, Ludwig  M, Tsai  KP, Sim  LJ, Morris  M. Baroreceptor input regulates osmotic control of central vasopressin secretion. Neuroendocrinology.  1997;65(4):238-245. [DOI] [PubMed] [Google Scholar]
  • 67. Verbalis  JG, Mangione  MP, Stricker  EM. Oxytocin produces natriuresis in rats at physiological plasma concentrations. Endocrinology.  1991;128(3):1317-1322. [DOI] [PubMed] [Google Scholar]
  • 68. Rasmussen  MS, Simonsen  JA, Sandgaard  NC, Høilund-Carlsen  PF, Bie  P. Effects of oxytocin in normal man during low and high sodium diets. Acta Physiol Scand.  2004;181(2):247-257. [DOI] [PubMed] [Google Scholar]
  • 69. McCann  SM, Franci  C, Gutkowska  J, Favaretto  AL, Antunes-Rodrigues  J. Neural control of atrial natriuretic peptide actions on fluid intake and excretion. Proc Soc Exp Biol Med.  1996;213(2):117-127. [DOI] [PubMed] [Google Scholar]
  • 70. Cunningham  JT, Bruno  SB, Grindstaff  RR, et al.  Cardiovascular regulation of supraoptic vasopressin neurons. Prog Brain Res.  2002;139(1):257-273. [PubMed] [Google Scholar]
  • 71. Goldsmith  SR, Francis  GS, Cowley  AW  Jr. Arginine vasopressin and the renal response to water loading in congestive heart failure. Am J Cardiol.  1986;58(3):295-299. [DOI] [PubMed] [Google Scholar]
  • 72. Ball  SG. Vasopressin and disorders of water balance: the physiology and pathophysiology of vasopressin. Ann Clin Biochem.  2007;44(Pt 5):417-431. [DOI] [PubMed] [Google Scholar]
  • 73. Grantham  J, Linshaw  M. The effect of hyponatremia on the regulation of intracellular volume and solute composition. Circ Res.  1984;54(5):483-491. [DOI] [PubMed] [Google Scholar]
  • 74. Balapattabi  K, Farmer  GE, Knapp  BA, et al.  Effects of salt-loading on supraoptic vasopressin neurones assessed by ClopHensorN chloride imaging. J Neuroendocrinol.  2019;31(8):e12752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Balapattabi  K, Little  JT, Farmer  GE, Cunningham  JT. High salt loading increases brain derived neurotrophic factor in supraoptic vasopressin neurones. J Neuroendocrinol.  2018;30(11):e12639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Dumais  KM, Veenema  AH. Vasopressin and oxytocin receptor systems in the brain: Sex differences and sex-specific regulation of social behavior. Front Neuroendocrinol.  2016;40(1):1-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Wang  Z, Zhou  L, Hulihan  TJ, Insel  TR. Immunoreactivity of central vasopressin and oxytocin pathways in microtine rodents: a quantitative comparative study. J Comp Neurol.  1996;366(4):726-737. [DOI] [PubMed] [Google Scholar]
  • 78. Wang  Y, Xu  L, Pan  Y, Wang  Z, Zhang  Z. Species differences in the immunoreactive expression of oxytocin, vasopressin, tyrosine hydroxylase and estrogen receptor alpha in the brain of Mongolian gerbils (Meriones unguiculatus) and Chinese striped hamsters (Cricetulus barabensis). PLoS One.  2013;8(6):e65807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Stone  JD, Crofton  JT, Share  L. Sex differences in central adrenergic control of vasopressin release. Am J Physiol.  1989;257(5 Pt 2):R1040-R1045. [DOI] [PubMed] [Google Scholar]
  • 80. Stone  JD, Crofton  JT, Share  L. Sex differences in central cholinergic and angiotensinergic control of vasopressin release. Am J Physiol.  1992;263(5 Pt 2):R1030-R1034. [DOI] [PubMed] [Google Scholar]
  • 81. Scott  N, Prigge  M, Yizhar  O, Kimchi  T. A sexually dimorphic hypothalamic circuit controls maternal care and oxytocin secretion. Nature.  2015;525(7570):519-522. [DOI] [PubMed] [Google Scholar]
  • 82. Williams  TD, Carter  DA, Lightman  SL. Sexual dimorphism in the posterior pituitary response to stress in the rat. Endocrinology.  1985;116(2):738-740. [DOI] [PubMed] [Google Scholar]
  • 83. Carter  DA, Lightman  SL. Modulation of oxytocin secretion by ascending noradrenergic pathways: sexual dimorphism in rats. Brain Res.  1987;406(1-2):313-316. [DOI] [PubMed] [Google Scholar]
  • 84. Minhas  S, Liu  C, Galdamez  J, So  VM, Romeo  RD. Stress-induced oxytocin release and oxytocin cell number and size in prepubertal and adult male and female rats. Gen Comp Endocrinol.  2016;234(1):103-109. [DOI] [PubMed] [Google Scholar]
  • 85. Zoladz  PR, Conrad  CD, Fleshner  M, Diamond  DM. Acute episodes of predator exposure in conjunction with chronic social instability as an animal model of post-traumatic stress disorder. Stress.  2008;11(4):259-281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Rosmond  R, Björntorp  P. The hypothalamic-pituitary-adrenal axis activity as a predictor of cardiovascular disease, type 2 diabetes and stroke. J Intern Med.  2000;247(2):188-197. [DOI] [PubMed] [Google Scholar]
  • 87. Kudielka  BM, Kirschbaum  C. Sex differences in HPA axis responses to stress: a review. Biol Psychol.  2005;69(1):113-132. [DOI] [PubMed] [Google Scholar]
  • 88. Heck  AL, Handa  RJ. Sex differences in the hypothalamic-pituitary-adrenal axis’ response to stress: an important role for gonadal hormones. Neuropsychopharmacology.  2019;44(1):45-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Dempster  KS, O’Leary  DD, MacNeil  AJ, Hodges  GJ, Wade  TJ. Linking the hemodynamic consequences of adverse childhood experiences to an altered HPA axis and acute stress response. Brain Behav Immun.  2021;93(1):254-263. [DOI] [PubMed] [Google Scholar]
  • 90. Balapattabi  K, Little  JT, Bachelor  ME, Cunningham  RL, Cunningham  JT. Sex Differences in the Regulation of Vasopressin and Oxytocin Secretion in Bile Duct-Ligated Rats. Neuroendocrinology.  2021;111(3):237-248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Taka-aki Koshimizu  YN, Tanoue  A, Oikawa  R, et al.  V1a vasopressin receptors maintain normal blood pressure by regulating circulating blood volume and baroreflex sensitivity. Proc Natl Acad Sci U S A. 2006;103(20):7807-7812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92. Hirsch  AT, Majzoub  JA, Ren  CJ, Scales  KM, Creager  MA. Contribution of vasopressin to blood pressure regulation during hypovolemic hypotension in humans. J Appl Physiol (1985).  1993;75(5):1984-1988. [DOI] [PubMed] [Google Scholar]
  • 93. Li  X, Chan  TO, Myers  V, et al.  Controlled and cardiac-restricted overexpression of the arginine vasopressin V1A receptor causes reversible left ventricular dysfunction through Gαq-mediated cell signaling. Circulation.  2011;124(5):572-581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. Wasilewski  MA, Myers  VD, Recchia  FA, Feldman  AM, Tilley  DG. Arginine vasopressin receptor signaling and functional outcomes in heart failure. Cell Signal.  2016;28(3):224-233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Zhu  W, Tilley  DG, Myers  VD, Tsai  EJ, Feldman  AM. Increased vasopressin 1A receptor expression in failing human hearts. J Am Coll Cardiol.  2014;63(4):375-376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Phillips  PA, Abrahams  JM, Kelly  JM, Mooser  V, Trinder  D, Johnston  CI. Localization of vasopressin binding sites in rat tissues using specific V1 and V2 selective ligands. Endocrinology.  1990;126(3):1478-1484. [DOI] [PubMed] [Google Scholar]
  • 97. Rabow  S, Hjorth  U, Schönbeck  S, Olofsson  P. Effects of oxytocin and anaesthesia on vascular tone in pregnant women: a randomised double-blind placebo-controlled study using non-invasive pulse wave analysis. BMC Pregnancy Childbirth.  2018;18(1):453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Thibonnier  M, Conarty  DM, Preston  JA, Plesnicher  CL, Dweik  RA, Erzurum  SC. Human vascular endothelial cells express oxytocin receptors. Endocrinology.  1999;140(3):1301-1309. [DOI] [PubMed] [Google Scholar]
  • 99. Uhl-Bronner  S, Waltisperger  E, Martínez-Lorenzana  G, Condes Lara  M, Freund-Mercier  MJ. Sexually dimorphic expression of oxytocin binding sites in forebrain and spinal cord of the rat. Neuroscience.  2005;135(1):147-154. [DOI] [PubMed] [Google Scholar]
  • 100. Dumais  KM, Bredewold  R, Mayer  TE, Veenema  AH. Sex differences in oxytocin receptor binding in forebrain regions: correlations with social interest in brain region- and sex- specific ways. Horm Behav.  2013;64(4):693-701. [DOI] [PubMed] [Google Scholar]
  • 101. Smith  CJW, Poehlmann  ML, Li  S, Ratnaseelan  AM, Bredewold  R, Veenema  AH. Age and sex differences in oxytocin and vasopressin V1a receptor binding densities in the rat brain: focus on the social decision-making network. Brain Struct Funct.  2017;222(2):981-1006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Davisson  RL, Yang  G, Beltz  TG, Cassell  MD, Johnson  AK, Sigmund  CD. The brain renin-angiotensin system contributes to the hypertension in mice containing both the human renin and human angiotensinogen transgenes. Circ Res.  1998;83(10):1047-1058. [DOI] [PubMed] [Google Scholar]
  • 103. Martins  AS, Crescenzi  A, Stern  JE, Bordin  S, Michelini  LC. Hypertension and exercise training differentially affect oxytocin and oxytocin receptor expression in the brain. Hypertension.  2005;46(4):1004-1009. [DOI] [PubMed] [Google Scholar]
  • 104. Cudnoch-Jedrzejewska  A, Dobruch  J, Puchalska  L, Szczepańska-Sadowska  E. Interaction of AT1 receptors and V1a receptors-mediated effects in the central cardiovascular control during the post-infarct state. Regul Pept.  2007;142(3):86-94. [DOI] [PubMed] [Google Scholar]
  • 105. Wsół  A, Cudnoch-Je drzejewska  A, Szczepanska-Sadowska  E, Kowalewski  S, Dobruch  J. Central oxytocin modulation of acute stress-induced cardiovascular responses after myocardial infarction in the rat. Stress.  2009;12(6):517-525. [DOI] [PubMed] [Google Scholar]
  • 106. Cudnoch-Jedrzejewska  A, Szczepanska-Sadowska  E, Dobruch  J, Gomolka  R, Puchalska  L. Brain vasopressin V(1) receptors contribute to enhanced cardiovascular responses to acute stress in chronically stressed rats and rats with myocardial infarcton. Am J Physiol Regul Integr Comp Physiol.  2010;298(3):R672-R680. [DOI] [PubMed] [Google Scholar]
  • 107. Dobruch  J, Cudnoch-Jedrzejewska  A, Szczepanska-Sadowska  E. Enhanced involvement of brain vasopressin V1 receptors in cardiovascular responses to stress in rats with myocardial infarction. Stress.  2005;8(4):273-284. [DOI] [PubMed] [Google Scholar]
  • 108. McCann  SM, Antunes-Rodrigues  J, Jankowski  M, Gutkowska  J. Oxytocin, vasopressin and atrial natriuretic peptide control body fluid homeostasis by action on their receptors in brain, cardiovascular system and kidney. Prog Brain Res.  2002;139(1):309-328. [DOI] [PubMed] [Google Scholar]
  • 109. Tribollet  E, Audigier  S, Dubois-Dauphin  M, Dreifuss  JJ. Gonadal steroids regulate oxytocin receptors but not vasopressin receptors in the brain of male and female rats. An autoradiographical study. Brain Res.  1990;511(1):129-140. [DOI] [PubMed] [Google Scholar]
  • 110. Ostrowski  NL, Lolait  SJ. Oxytocin receptor gene expression in female rat kidney. The effect of estrogen. Adv Exp Med Biol.  1995;395(1):329-340. [PubMed] [Google Scholar]
  • 111. Ostrowski  NL, Young  WS  3rd, Lolait  SJ. Estrogen increases renal oxytocin receptor gene expression. Endocrinology.  1995;136(4):1801-1804. [DOI] [PubMed] [Google Scholar]
  • 112. Liu  J, Sharma  N, Zheng  W, et al.  Sex differences in vasopressin V₂ receptor expression and vasopressin-induced antidiuresis. Am J Physiol Renal Physiol.  2011;300(2):F433-F440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Juul  KV, Bichet  DG, Nielsen  S, Nørgaard  JP. The physiological and pathophysiological functions of renal and extrarenal vasopressin V2 receptors. Am J Physiol Renal Physiol.  2014;306(9):F931-F940. [DOI] [PubMed] [Google Scholar]
  • 114. Schrier  RW. Water and sodium retention in edematous disorders: role of vasopressin and aldosterone. Am J Med.  2006;119(7 Suppl 1):S47-S53. [DOI] [PubMed] [Google Scholar]
  • 115. Tasker  JG, Prager-Khoutorsky  M, Teruyama  R, Lemos  JR, Amstrong  WE. Advances in the neurophysiology of magnocellular neuroendocrine cells. J Neuroendocrinol.  2020;32(4):e12826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Mason  WT. Supraoptic neurones of rat hypothalamus are osmosensitive. Nature.  1980;287(5778):154-157. [DOI] [PubMed] [Google Scholar]
  • 117. Bourque  CW, Renaud  LP. Activity patterns and osmosensitivity of rat supraoptic neurones in perfused hypothalamic explants. J Physiol.  1984;349(1):631-642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Bansal  V, Fisher  TE. Osmotic activation of a Ca(2+)-dependent phospholipase C pathway that regulates N TRPV1-mediated currents in rat supraoptic neurons. Physiol Rep. 2017;5(8):e13259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Sharif Naeini  R, Witty  MF, Séguéla  P, Bourque  CW. An N-terminal variant of Trpv1 channel is required for osmosensory transduction. Nat Neurosci.  2006;9(1):93-98. [DOI] [PubMed] [Google Scholar]
  • 120. Nedungadi  TP, Carreño  FR, Walch  JD, Bathina  CS, Cunningham  JT. Region-specific changes in transient receptor potential vanilloid channel expression in the vasopressin magnocellular system in hepatic cirrhosis-induced hyponatraemia. J Neuroendocrinol.  2012;24(4):642-652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Carreño  FR, Ji  LL, Cunningham  JT. Altered central TRPV4 expression and lipid raft association related to inappropriate vasopressin secretion in cirrhotic rats. Am J Physiol Regul Integr Comp Physiol.  2009;296(2):R454-R466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Saxena  A, Bachelor  M, Park  YH, Carreno  FR, Nedungadi  TP, Cunningham  JT. Angiotensin II induces membrane trafficking of natively expressed transient receptor potential vanilloid type 4 channels in hypothalamic 4B cells. Am J Physiol Regul Integr Comp Physiol.  2014;307(8):R945-R955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Zhang  Z, Kindrat  AN, Sharif-Naeini  R, Bourque  CW. Actin filaments mediate mechanical gating during osmosensory transduction in rat supraoptic nucleus neurons. J Neurosci.  2007;27(15):4008-4013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Prager-Khoutorsky  M, Bourque  CW. Osmosensation in vasopressin neurons: changing actin density to optimize function. Trends Neurosci.  2010;33(2):76-83. [DOI] [PubMed] [Google Scholar]
  • 125. Prager-Khoutorsky  M, Khoutorsky  A, Bourque  CW. Unique interweaved microtubule scaffold mediates osmosensory transduction via physical interaction with TRPV1. Neuron.  2014;83(4):866-878. [DOI] [PubMed] [Google Scholar]
  • 126. Zhang  Z, Bourque  CW. Amplification of transducer gain by angiotensin II-mediated enhancement of cortical actin density in osmosensory neurons. J Neurosci.  2008;28(38):9536-9544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127. Goswami  C, Dreger  M, Jahnel  R, Bogen  O, Gillen  C, Hucho  F. Identification and characterization of a Ca2+-sensitive interaction of the vanilloid receptor TRPV1 with tubulin. J Neurochem.  2004;91(5):1092-1103. [DOI] [PubMed] [Google Scholar]
  • 128. Goswami  C, Hucho  TB, Hucho  F. Identification and characterisation of novel tubulin-binding motifs located within the C-terminus of TRPV1. J Neurochem.  2007;101(1):250-262. [DOI] [PubMed] [Google Scholar]
  • 129. Hicks  AI, Barad  Z, Sobrero  A, et al.  Effects of salt loading on the organisation of microtubules in rat magnocellular vasopressin neurones. J Neuroendocrinol.  2020;32(2):e12817. [DOI] [PubMed] [Google Scholar]
  • 130. Oliet  SH, Bonfardin  VD. Morphological plasticity of the rat supraoptic nucleus–cellular consequences. Eur J Neurosci.  2010;32(12):1989-1994. [DOI] [PubMed] [Google Scholar]
  • 131. Tasker  JG, Oliet  SH, Bains  JS, Brown  CH, Stern  JE. Glial regulation of neuronal function: from synapse to systems physiology. J Neuroendocrinol.  2012;24(4):566-576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Theodosis  DT, el Majdoubi  M, Gies  U, Poulain  DA. Physiologically-linked structural plasticity of inhibitory and excitatory synaptic inputs to oxytocin neurons. Adv Exp Med Biol.  1995;395(1):155-171. [PubMed] [Google Scholar]
  • 133. Hatton  GI, Perlmutter  LS, Salm  AK, Tweedle  CD. Dynamic neuronal-glial interactions in hypothalamus and pituitary: implications for control of hormone synthesis and release. Peptides.  1984;5(Suppl 1):121-138. [DOI] [PubMed] [Google Scholar]
  • 134. Miyata  S, Nakashima  T, Kiyohara  T. Structural dynamics of neural plasticity in the supraoptic nucleus of the rat hypothalamus during dehydration and rehydration. Brain Res Bull.  1994;34(3):169-175. [DOI] [PubMed] [Google Scholar]
  • 135. Di  S, Tasker  JG. Dehydration-induced synaptic plasticity in magnocellular neurons of the hypothalamic supraoptic nucleus. Endocrinology.  2004;145(11):5141-5149. [DOI] [PubMed] [Google Scholar]
  • 136. Currás-Collazo  MC, Dao  J. Osmotic activation of the hypothalamo-neurohypophysial system reversibly downregulates the NMDA receptor subunit, NR2B, in the supraoptic nucleus of the hypothalamus. Brain Res Mol Brain Res.  1999;70(2):187-196. [DOI] [PubMed] [Google Scholar]
  • 137. Di  S, Jiang  Z, Wang  S, et al.  Labile calcium-permeable AMPA receptors constitute new glutamate synapses formed in hypothalamic neuroendocrine cells during salt loading. eNeuro. 2019;6(4):ENEURO.0112-19.2019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Hu  B, Cunningham  JT, Nissen  R, Renaud  LP, Bourque  CW. Rat supraoptic neurons are resistant to glutamate neurotoxicity. Neuroreport.  1992;3(1):87-90. [DOI] [PubMed] [Google Scholar]
  • 139. Carreño  FR, Walch  JD, Dutta  M, Nedungadi  TP, Cunningham  JT. Brain-derived neurotrophic factor-tyrosine kinase B pathway mediates NMDA receptor NR2B subunit phosphorylation in the supraoptic nuclei following progressive dehydration. J Neuroendocrinol.  2011;23(10):894-905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140. Miyata  S, Itoh  T, Matsushima  O, Nakashima  T, Kiyohara  T. Not only osmotic stress but also repeated restraint stress causes structural plasticity in the supraoptic nucleus of the rat hypothalamus. Brain Res Bull.  1994;33(6):669-675. [DOI] [PubMed] [Google Scholar]
  • 141. Kim  JS, Kim  WB, Kim  YB, et al.  Chronic hyperosmotic stress converts GABAergic inhibition into excitation in vasopressin and oxytocin neurons in the rat. J Neurosci.  2011;31(37):13312-13322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142. Konopacka  A, Qiu  J, Yao  ST, et al.  Osmoregulation requires brain expression of the renal Na-K-2Cl cotransporter NKCC2. J Neurosci.  2015;35(13):5144-5155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143. Balapattabi  K, Little  JT, Bachelor  M, Cunningham  JT. Brain-derived neurotrophic factor and supraoptic vasopressin neurons in hyponatremia. Neuroendocrinology.  2020;110(7-8):630-641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. Korpal  AK, Han  SY, Schwenke  DO, Brown  CH. A switch from GABA inhibition to excitation of vasopressin neurons exacerbates the development angiotensin II-dependent hypertension. J Neuroendocrinol. 2018;30(8):e12564. [DOI] [PubMed] [Google Scholar]
  • 145. Han  SY, Bouwer  GT, Seymour  AJ, Korpal  AK, Schwenke  DO, Brown  CH. Induction of hypertension blunts baroreflex inhibition of vasopressin neurons in the rat. Eur J Neurosci.  2015;42(9):2690-2698. [DOI] [PubMed] [Google Scholar]
  • 146. Ye  ZY, Li  DP, Byun  HS, Li  L, Pan  HL. NKCC1 upregulation disrupts chloride homeostasis in the hypothalamus and increases neuronal activity-sympathetic drive in hypertension. J Neurosci.  2012;32(25):8560-8568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Garty  H, Palmer  LG. Epithelial sodium channels: function, structure, and regulation. Physiol Rev.  1997;77(2):359-396. [DOI] [PubMed] [Google Scholar]
  • 148. Amin  MS, Wang  HW, Reza  E, Whitman  SC, Tuana  BS, Leenen  FH. Distribution of epithelial sodium channels and mineralocorticoid receptors in cardiovascular regulatory centers in rat brain. Am J Physiol Regul Integr Comp Physiol.  2005;289(6):R1787-R1797. [DOI] [PubMed] [Google Scholar]
  • 149. Teruyama  R, Sakuraba  M, Wilson  LL, Wandrey  NE, Armstrong  WE. Epithelial Na⁺ sodium channels in magnocellular cells of the rat supraoptic and paraventricular nuclei. Am J Physiol Endocrinol Metab.  2012;302(3):E273-E285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Sharma  K, Haque  M, Guidry  R, Ueta  Y, Teruyama  R. Effect of dietary salt intake on epithelial Na+ channels (ENaC) in vasopressin magnocellular neurosecretory neurons in the rat supraoptic nucleus. J Physiol.  2017;595(17):5857-5874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Lebrun  CJ, Rohmeiss  P, Demmert  G, Unger  T. Sensitization to pressor effects by repeated central injections of vasopressin in conscious rats. Brain Res.  1989;505(1):39-43. [DOI] [PubMed] [Google Scholar]
  • 152. Poulin  P, Szot  P, Dorsa  DM, Pittman  QJ. Vasopressin-induced sensitization: involvement of neurohypophyseal peptide receptors. Eur J Pharmacol.  1995;294(1):29-39. [DOI] [PubMed] [Google Scholar]
  • 153. Pelton  GH, Lee  Y, Davis  M. Repeated stress, like vasopressin, sensitizes the excitatory effects of corticotropin releasing factor on the acoustic startle reflex. Brain Res.  1997;778(2):381-387. [DOI] [PubMed] [Google Scholar]
  • 154. Poulin  P, Pittman  QJ. Arginine vasopressin-induced sensitization in brain: facilitated inositol phosphate production without changes in receptor number. J Neuroendocrinol.  1993;5(1):23-31. [DOI] [PubMed] [Google Scholar]
  • 155. Landgraf  R, Ludwig  M. Vasopressin release within the supraoptic and paraventricular nuclei of the rat brain: osmotic stimulation via microdialysis. Brain Res.  1991;558(2):191-196. [DOI] [PubMed] [Google Scholar]
  • 156. Ludwig  M, Callahan  MF, Neumann  I, Landgraf  R, Morris  M. Systemic osmotic stimulation increases vasopressin and oxytocin release within the supraoptic nucleus. J Neuroendocrinol.  1994;6(4):369-373. [DOI] [PubMed] [Google Scholar]
  • 157. Neumann  I, Ludwig  M, Engelmann  M, Pittman  QJ, Landgraf  R. Simultaneous microdialysis in blood and brain: oxytocin and vasopressin release in response to central and peripheral osmotic stimulation and suckling in the rat. Neuroendocrinology.  1993;58(6):637-645. [DOI] [PubMed] [Google Scholar]
  • 158. Dutra  SGV, Paterson  A, Monteiro  LRN, et al.  Physiological and transcriptomic changes in the hypothalamic-neurohypophysial system after 24 h of furosemide-induced sodium depletion. Neuroendocrinology.  2021;111(1-2):70-86. [DOI] [PubMed] [Google Scholar]
  • 159. Greenwood  MP, Mecawi  AS, Hoe  SZ, et al.  A comparison of physiological and transcriptome responses to water deprivation and salt loading in the rat supraoptic nucleus. Am J Physiol Regul Integr Comp Physiol.  2015;308(7):R559-R568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160. Greenwood  M, Greenwood  MP, Mecawi  AS, et al.  Transcription factor CREB3L1 mediates cAMP and glucocorticoid regulation of arginine vasopressin gene transcription in the rat hypothalamus. Mol Brain.  2015;8(1):68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Greenwood  M, Paterson  A, Rahman  PA, et al.  Transcription factor Creb3l1 regulates the synthesis of prohormone convertase enzyme PC1/3 in endocrine cells. J Neuroendocrinol.  2020;32(4):e12851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Gottlieb  HB, Ji  LL, Jones  H, Penny  ML, Fleming  T, Cunningham  JT. Differential effects of water and saline intake on water deprivation-induced c-Fos staining in the rat. Am J Physiol Regul Integr Comp Physiol.  2006;290(5):R1251-R1261. [DOI] [PubMed] [Google Scholar]
  • 163. Ji  LL, Fleming  T, Penny  ML, Toney  GM, Cunningham  JT. Effects of water deprivation and rehydration on c-Fos and FosB staining in the rat supraoptic nucleus and lamina terminalis region. Am J Physiol Regul Integr Comp Physiol.  2005;288(1):R311-R321. [DOI] [PubMed] [Google Scholar]
  • 164. Yao  ST, Gouraud  SS, Qiu  J, Cunningham  JT, Paton  JF, Murphy  D. Selective up-regulation of JunD transcript and protein expression in vasopressinergic supraoptic nucleus neurones in water-deprived rats. J Neuroendocrinol.  2012;24(12):1542-1552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165. Howe  BM, Bruno  SB, Higgs  KA, Stigers  RL, Cunningham  JT. FosB expression in the central nervous system following isotonic volume expansion in unanesthetized rats. Exp Neurol.  2004;187(1):190-198. [DOI] [PubMed] [Google Scholar]
  • 166. Cunningham  JT, Nedungadi  TP, Walch  JD, Nestler  EJ, Gottlieb  HB. ΔFosB in the supraoptic nucleus contributes to hyponatremia in rats with cirrhosis. Am J Physiol Regul Integr Comp Physiol.  2012;303(2):R177-R185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167. Walch  JD, Nedungadi  TP, Cunningham  JT. ANG II receptor subtype 1a gene knockdown in the subfornical organ prevents increased drinking behavior in bile duct-ligated rats. Am J Physiol Regul Integr Comp Physiol.  2014;307(6):R597-R607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168. Greenwood  MP, Greenwood  M, Gillard  BT, Loh  SY, Paton  JF, Murphy  D. Epigenetic control of the vasopressin promoter explains physiological ability to regulate vasopressin transcription in dehydration and salt loading states in the rat. J Neuroendocrinol. 2016;28(4):e12371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Murgatroyd  C, Patchev  AV, Wu  Y, et al.  Dynamic DNA methylation programs persistent adverse effects of early-life stress. Nat Neurosci.  2009;12(12):1559-1566. [DOI] [PubMed] [Google Scholar]
  • 170. Gao  Q, Fan  X, Xu  T, et al.  Promoter methylation changes and vascular dysfunction in pre-eclamptic umbilical vein. Clin Epigenetics.  2019;11(1):84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Gao  Q, Li  H, Ding  H, et al.  Hyper-methylation of AVPR1A and PKCΒ gene associated with insensitivity to arginine vasopressin in human pre-eclamptic placental vasculature. EBioMedicine.  2019;44(1):574-581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172. Ferreira-Neto  HC, Biancardi  VC, Stern  JE. A reduction in SK channels contributes to increased activity of hypothalamic magnocellular neurons during heart failure. J Physiol.  2017;595(20):6429-6442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Ferreira-Neto  HC, Stern  JE. Functional coupling between NMDA receptors and SK channels in rat hypothalamic magnocellular neurons: altered mechanisms during heart failure. J Physiol.  2021;599(2):507-520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174. Zhang  M, Biancardi  VC, Stern  JE. An increased extrasynaptic NMDA tone inhibits A-type K+ current and increases excitability of hypothalamic neurosecretory neurons in hypertensive rats. J Physiol.  2017;595(14):4647-4661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175. Zhang  M, Stern  JE. Altered NMDA receptor-evoked intracellular Ca2+ dynamics in magnocellular neurosecretory neurons of hypertensive rats. J Physiol.  2017;595(24):7399-7411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Pitra  S, Feng  Y, Stern  JE. Mechanisms underlying prorenin actions on hypothalamic neurons implicated in cardiometabolic control. Mol Metab.  2016;5(10):858-868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Pitra  S, Stern  JE. A-type K+ channels contribute to the prorenin increase of firing activity in hypothalamic vasopressin neurosecretory neurons. Am J Physiol Heart Circ Physiol.  2017;313(3):H548-H557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178. Pitra  S, Worker  CJ, Feng  Y, Stern  JE. Exacerbated effects of prorenin on hypothalamic magnocellular neuronal activity and vasopressin plasma levels during salt-sensitive hypertension. Am J Physiol Heart Circ Physiol.  2019;317(3):H496-H504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Bains  JS, Wamsteeker Cusulin  JI, Inoue  W. Stress-related synaptic plasticity in the hypothalamus. Nat Rev Neurosci.  2015;16(7):377-388. [DOI] [PubMed] [Google Scholar]
  • 180. Johnson  AK, Xue  B. Central nervous system neuroplasticity and the sensitization of hypertension. Nat Rev Nephrol.  2018;14(12):750-766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181. Hurley  SW, Beltz  TG, Guo  F, Xue  B, Johnson  AK. Amphetamine-induced sensitization of hypertension and lamina terminalis neuroinflammation. Am J Physiol Regul Integr Comp Physiol.  2020;318(3):R649-R656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Sandgren  JA, Scroggins  SM, Santillan  DA, et al.  Vasopressin: the missing link for preeclampsia?  Am J Physiol Regul Integr Comp Physiol.  2015;309(9):R1062-R1064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183. Rodriguez  M, Hernandez  M, Cheungpasitporn  W, et al.  Hyponatremia in heart failure: pathogenesis and management. Curr Cardiol Rev.  2019;15(4):252-261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184. Wang  L, Zhang  Q, Liu  M, et al.  Tolvaptan in reversing worsening acute heart failure: A systematic review and meta-analysis. J Int Med Res.  2019;47(11):5414-5425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185. Adachi  T, Takaki  A, Sato  S, et al.  High expression of a vascular stricture-related marker is predictive of an early response to tolvaptan, and a low fractional excretion of sodium is predictive of a poor long-term survival after tolvaptan administration for liver cirrhosis. Hepatol Res.  2020;50(12):1347-1354. [DOI] [PubMed] [Google Scholar]
  • 186. Kanayama  K, Chiba  T, Kobayashi  K, et al.  Long-term administration of Tolvaptan to patients with decompensated cirrhosis. Int J Med Sci.  2020;17(7):874-880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Brown  CH, Ludwig  M, Tasker  JG, Stern  JE. Somato-dendritic vasopressin and oxytocin secretion in endocrine and autonomic regulation. J Neuroendocrinol.  2020;32(6):e12856. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.


Articles from Endocrinology are provided here courtesy of The Endocrine Society

RESOURCES