Next Article in Journal
Epigenetic Dysregulations in Merkel Cell Polyomavirus-Driven Merkel Cell Carcinoma
Next Article in Special Issue
Signal Transduction of Mineralocorticoid and Angiotensin II Receptors in the Central Control of Sodium Appetite: A Narrative Review
Previous Article in Journal
Plants Saline Environment in Perception with Rhizosphere Bacteria Containing 1-Aminocyclopropane-1-Carboxylate Deaminase
Previous Article in Special Issue
Oxytocin, Erectile Function and Sexual Behavior: Last Discoveries and Possible Advances
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Complementary Role of Oxytocin and Vasopressin in Cardiovascular Regulation

Laboratory of Centre for Preclinical Research, Chair and Department of Experimental and Clinical Physiology, Medical University of Warsaw, 02-091 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(21), 11465; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222111465
Submission received: 30 September 2021 / Revised: 21 October 2021 / Accepted: 22 October 2021 / Published: 24 October 2021
(This article belongs to the Special Issue Advances in Oxytocin)

Abstract

:
The neurons secreting oxytocin (OXY) and vasopressin (AVP) are located mainly in the supraoptic, paraventricular, and suprachiasmatic nucleus of the brain. Oxytocinergic and vasopressinergic projections reach several regions of the brain and the spinal cord. Both peptides are released from axons, soma, and dendrites and modulate the excitability of other neuroregulatory pathways. The synthesis and action of OXY and AVP in the peripheral organs (eye, heart, gastrointestinal system) is being investigated. The secretion of OXY and AVP is influenced by changes in body fluid osmolality, blood volume, blood pressure, hypoxia, and stress. Vasopressin interacts with three subtypes of receptors: V1aR, V1bR, and V2R whereas oxytocin activates its own OXTR and V1aR receptors. AVP and OXY receptors are present in several regions of the brain (cortex, hypothalamus, pons, medulla, and cerebellum) and in the peripheral organs (heart, lungs, carotid bodies, kidneys, adrenal glands, pancreas, gastrointestinal tract, ovaries, uterus, thymus). Hypertension, myocardial infarction, and coexisting factors, such as pain and stress, have a significant impact on the secretion of oxytocin and vasopressin and on the expression of their receptors. The inappropriate regulation of oxytocin and vasopressin secretion during ischemia, hypoxia/hypercapnia, inflammation, pain, and stress may play a significant role in the pathogenesis of cardiovascular diseases.

1. Synthesis and Release of Oxytocin and Vasopressin

Oxytocin (Cys-Tyr-Ile-Gln-Asn-Cys-Pro-Leu-Gly-NH2; OXY) and arginine vasopressin (Cys-Tyr-Phe-Gln-Asn-Cys-Pro-Arg-Gly-NH2; AVP) are nonapeptides, differing only in two aminoacids [1,2,3]. In vertebrates, OXY and AVP are synthesized mainly in the hypothalamus as preprohormones, namely preprooxytocin and preprovasopressin. They are then processed along the axonal projections to the posterior lobe of the pituitary, where they are stored in secretory vesicles and released to the peripheral circulation as final products [4]. The neurons secreting OXY and AVP are located mainly in the supraoptic nucleus (SON), paraventricular nucleus (PVN), and the suprachiasmatic nucleus (SCN) [2,5,6,7,8]. In mammals, the preprohormone of vasopressin consists of vasopressin (most frequently arginine vasopressin, in some species lysine vasopressin), neurophysin and glycopeptide copeptin, which are cleaved and released in equimolar amounts. The stoichiometric generation of vasopressin and copeptin enables the measurement of the blood copeptin concentration instead of the AVP concentration [9,10,11,12]. The preproprotein of OXY cleavage results in the final release of oxytocin and its carrier protein, neurophysin [3].
In all species, OXY and AVP genes are located on the same chromosomes. They lie on opposite strands and are transcribed in opposite directions [3]. In human beings, mice, and rats, the genes encoding OXY and AVP prepropeptides are mapped correspondingly to chromosome 20p13 (www.ncbi.nlm.nih.gov/gene/551 (accessed on 15 October 2021)), chromosome 2 (www.ncbi.nlm.nih.gov/gene/11998 (accessed on 21 October 2021)) and chromosome 3 (www.ncbi.nlm.nih.gov/gene/24221 (accessed on 30 September 2021)). Under pathological conditions, the synthesis of inappropriate peptides is caused by polymorphisms of genes encoding specific elements of the OXY-AVP pathways [13].
Apart from AVP and OXY, vertebrates can synthesize lysine vasopressin (LVP, some mammals), vasotocin (birds, amphibians, fish), isotocin (teleost fish), and mesotocin (frogs) [1]. These compounds are released either to the systemic circulation and act as potent multifunctional substances or they are released locally and regulate the function of neighboring cells [14,15].
The main source of circulating oxytocin and vasopressin is the neurohypophysis, which is situated in the pituitary fossa of the sphenoid bone together with the adenohypophysis. In mammals, hypophysis is vascularized by the superior hypophyseal artery, which emerges from the circle of Willis. The branches forming the arteriae infundibularis vascularize the infundibular part of the adenohypophysis and form a capillary network on the external surface of the infundibulum, through which blood flows into the sinusoidal capillaries of the anterior lobe [16,17]. Neurohypophysis is pierced by multiple vessels. The inferior hypophyseal artery enters the organ of the posterior circumference of the distal neurohypophysis (neural lobe, posterior lobe). The blood capillaries of the posterior lobe possess a fenestrated endothelium with multiple pores, enabling the local release of peptides into the adenohypophysis and/or the cerebrospinal fluid. In the posterior lobe, the endings of the neurosecretory axons form a dense net together with blood capillaries and glial cells [16,17]. The process of fusion of the secretory vesicle with the membrane and the secretion of peptides involves activation of protein kinase A and protein kinase C and the influx of intracellular calcium.
Vasopressinergic and oxytocinergic projections also reach several regions of the central nervous system (CNS) and the spinal cord. Both peptides are released from axons, soma, and dendrites and modulate the excitability of other neuroregulatory pathways [15,18,19,20,21,22,23]. Several observations have revealed the release and/or action of OXY and AVP in the olfactory bulb, orbitofrontal cortex, anterior cingulate gyrus, amygdala, striatum, hippocampus, bed nucleus of the stria terminalis, suprachiasmatic nucleus, and the autonomic ganglia [7,15,18,19,24,25,26,27,28,29,30,31,32,33]. The local synthesis and action of vasopressin and oxytocin in the peripheral organs, such as the eye, heart, and gastrointestinal system has been observed (see below and [34,35,36,37,38]).
The secretion of oxytocin and vasopressin is enhanced by hyperosmolality, reduction of the circulating blood volume, hypotension, hypoxia, stress, and hyperthermia [17,39,40,41,42,43,44,45,46]. The effects of these signals are associated with the activation of multiple receptors and transmitting pathways. Neurons synthesizing oxytocin and vasopressin communicate with several neurotransmitter systems and participate in the neuromodulation and regulation of emotions and behavior [18,47,48,49,50,51,52]. The abundant innervation of oxytocinergic and vasopressinergic neurons enables the precise adjustment of the central and systemic secretion of OXY and AVP to specific physiological conditions and changes in the internal and external environment (see below and [34,35,36,37,38]).

2. Vasopressin and Oxytocin Receptors

Vasopressin operates through the activation of three subtypes of receptors: V1aR (AVPR1a), V1bR (AVPR1b, V3R), and V2R (AVPR2). Oxytocin acts through its own OXTR receptors, but it can also stimulate V1aR. The AVP/OXY receptors belong to different subtypes of the G protein-coupled receptors (GPCRs). Specifically, V1aR, V1bR, and OXTR belong to the Gq/11 protein family. The Gq/11 protein family participates in the activation of the phospholipase C signaling pathway, in the mobilization of Ca2+ from intracellular stores, in the activation of protein kinase type C (PKC), and in the generation of inositol 1,4,5-trisphosphate (Ins3P) and diacylglycerol (DAG). V2R belongs to the Gs protein receptors engaged in the activation of protein kinase A (PKA) and the adenylate cyclase signaling pathway, employing intracellular cAMP as the intracellular second messenger [4,46,53,54,55,56]. The PKA-independent intracellular effects of V2R stimulation on aquaporin 2 (AQP2) in the kidney have also been described [57].

Location of Vasopressin and Oxytocin Receptors

The genes for OXY and AVP as well as for AVPR1a, AVPR1b, AVPR2, and OXTR are located on different chromosomes [47,58,59,60]. These genes may manifest polymorphisms which are associated with inappropriate responsiveness to the ligand [61,62,63]. In humans, the OXTR gene is present as a single copy, which is mapped to the gene locus 3p25–3 (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/gene/5021 (accessed on 17 October 2021)), while AVP receptor genes are mapped to different loci, namely AVPR1a to 12q14.2 (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/gene/552 (accessed on 19 September 2021)), AVPR1b to 1q32.1 (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/gene/553 (accessed on 19 September 2021)) and AVPR2 to Xq28 (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/gene/554 (accessed on 8 July 2021)).
The synthesis and functional engagement of vasopressin V1aR, V1bR, and OXTR in several regions of the CNS as well as in the peripheral organs has been observed. Experiments have made it possible to detect V1aR, V1bR, and OXTR mRNAs and proteins in the cerebral cortex, hypothalamus, pons, medulla, and cerebellum [3,29,33,50,64,65,66,67,68,69,70,71,72]. V1R and OXTR are also present in the arteries, veins, heart, lungs, carotid bodies, kidneys, adrenal glands, pancreas, gastrointestinal tract, eyes, ovaries, uterus, thymus, and autonomic ganglia [30,35,37,38,46,69,73,74,75,76,77,78]. Vasopressin V2R mRNA and protein were found in the kidneys and the urinary bladder [58,79,80]. V2R mRNA was also detected in the human gastrointestinal tract [37]. It may be that V1aR, V2R, and OXTR form dimers or heteromers with other receptors and that this type of processing may have an impact on the responsiveness to the ligand [81,82,83,84,85].

3. Regulation of Oxytocin and Vasopressin Secretion

3.1. Neuronal and Non-Neuronal Oxytocin Transport and Secretion

The human oxytocin gene (20p13) consists of three exons. The first exon encodes the translocator signal, the oxytocin peptide, the processing signal (glycyl-lysyl-arginine—GRK), and the first nine residues of neurophysin I, the second exon encodes the central part of neurophysin, and the third exon encodes the COOH-terminal region of neurophysin. In most species, the OXY gene extends over 800 bp. Its regulatory region shows remarkable polymorphism [86,87,88]. After synthesis, oxytocin is either transported in large dense-core vesicles to the neurohypophysial terminals in the posterior pituitary or it is stored in the somato-dendritic regions of the magnocellular neurons of the supraoptic and paraventricular nuclei [21,89]. Transport of vesicles depends on the activation of protein kinase A (PKA) and protein kinase C (PKC). The activation of PKA results in an enhanced association of kinesin-2 with annexin A1 (ANXA1) and increased axon-localization of the OXY vesicles. The activation of PKC attenuates the binding of kinesin-2 to ANXA1 and reduces the axonal transport of OXY [90]. Oxytocin is also secreted from the axon terminals into the hypophyseal portal blood circulation of the median eminence [91]. The axonal exocytosis of oxytocin into the portal capillaries involves conventional mechanisms of neuronal exocytosis with the activation of the primary calcium voltage-dependent channels and subsequent intracellular Ca2+ influx. This is followed by the formation of a core complex (known as the soluble N-ethylmaleimide sensitive fusion protein attachment protein receptor, SNARE), which enables the fusion of the membrane of the vesicle to the cellular membrane, and the subsequent release of the cargo into the extracellular space [92,93,94,95,96]. The somato-dendritic release of oxytocin also depends on the entry of the extracellular Ca2+. The magnocellular neurons express various types of calcium channels (L, P/Q, N, and R) located on the somata and dendrites [97,98,99]. The excitability of the OXY neurons is regulated by several factors operating in the excitatory and inhibitory synapses, such as adhesion molecules, neurotransmitters, regulatory peptides, and gasotransmitters [100,101,102]. It has been shown that the somato-dendritic release of OXY can occur independently of depolarization, and that the dendritically released oxytocin can exert both local autocrine effects to modulate the activity of the mother neurons and paracrine effects on the surrounding neurons and glial cells [100].
It has been established that OXY is synthesized and secreted in the peripheral organs, including the heart, the kidneys, adrenals, testes, and uterus [3,34,103,104]. Radioimmunoassay and molecular studies on rats have revealed a high expression of OXY and OXTR in the cardiomyocytes of the right atrium and in the left and right ventricles [35,105]. OXTRs are also present in human vascular endothelial cells, large vessels, and cardiac microvessels [35,103,106]. The synthesis of oxytocin in cultured beta cell lines originating from murine, rodent, and human pancreatic islets has been observed [36].

3.2. Regulation of Oxytocin Secretion

The secretion of OXY is triggered by classical reproductive and maternal stimuli such as birth, maternal care of the offspring, suckling and lactation, and mating and sexual stimulation in males and females. In addition, it is modulated by dehydration, hyperosmolarity, hypovolemia, hemorrhage, and exposure to stressors (Figure 1). Oxytocin neurons are abundantly innervated and the secretion of OXY is stimulated by several neurotransmitters and neuropeptides (α-melanocyte-stimulating hormone—α-MSH; angiotensins, corticotropin releasing factor—CRF, dopamine, glucocorticoids, norepinephrine, serotonin), and inhibited by GABA, orexins, prolactin, and central opioids [52,101,107,108,109,110,111,112].
In estrogen sensitive loci, such as the uterus, hypothalamus, heart, smooth muscles, and mammary glands, estrogen and its receptors are essential for OXY gene transcription and the induction of OXTR. It has been found that ERα-mediated activity modulates OXTR transcription, whereas ERβ regulates OXY mRNA and OXY levels. In addition, treatment with estradiol increases OXY mRNA expression in the brains of wild-type mice, but not in ERβ knockout (ERβKO) male and female mice [113,114,115].

3.3. Regulation of Vasopressin Secretion

The synthesis and secretion of vasopressin and oxytocin show a remarkable similarity. In humans, the AVP gene, linked on chromosome 20p13, consists of three exons and encodes the nonapeptide vasopressin, neurophysin II, and copeptin [88,116,117,118]. Under physiological conditions, vasopressin regulates the water balance, body fluid osmolality, arterial blood pressure, body temperature, susceptibility to stress, and emotional challenges, and its secretion is adjusted to fluctuations of these parameters [14,119,120,121,122] (Figure 1).
Vasopressinergic neurons are particularly sensitive to changes in sodium concentration. The threshold level of sodium concentration for AVP release is estimated at 140 mEqNa+/l, and an increase by 1–2% is sufficient to stimulate further release. Other substances increasing osmolality, such as glucose, urea, or ATP are less effective stimulants [8,19,44,116,123,124,125,126]. However, it should be noted that, in the explants of the supraoptic nucleus, the effectiveness of glucose in releasing vasopressin is noticeably potentiated by insulin [124]. It is postulated that the increase of osmolality causes a shrinkage of the hypothalamic magnocellular neurons. This results in cytoskeletal alterations and an opening of the delta-N variant of the transient receptor potential vanniloid (TRPV1) channels, which serve as the molecular detectors of osmotic stress [127,128]. The opening of these channels results in depolarization of the neuronal soma (from −60 mV to −55 mV) which is followed by the generation of action potentials and the release of AVP [14]. TRPV1 activity is potentiated by an influx of calcium ions through L-type Ca2+ channels and by an activation of the protein kinase C-dependent pathway [129]. The osmoregulation of AVP secretion can be inhibited by anticipatory stimuli engaging the anterior cingulate area [14]. The hypothalamic vasopressinergic neurons receive a rich input from the neurons located in multiple regions of the brain and the spinal cord and are abundantly furnished with receptors for neurochemical compounds that modulate the sensitivity of these neurons to osmotic stimuli [130,131]. They are also located in the median eminence of the anteroventral third ventricle region (AV3V). The destruction of this region results in the inhibition of AVP secretion, and thus results in hyperosmolality and hypernatremia [132].
The release of vasopressin is also regulated by signals generated in the osmotic receptors of the gastrointestinal tract and liver, which may play an essential role in the anticipative adjustment of AVP secretion to the osmolality of an ingested meal [133,134].
Vasopressin secretion is significantly influenced by changes in blood volume, blood pressure, and oxygen content that are signaled to the vasopressinergic neurons by afferent inputs from the baroreceptors, low pressure mechanoreceptors, and arterial chemoreceptors, respectively [135,136,137]. The neurochemical background of the processes linking the transmission of signals from the cardiovascular receptors and chemoreceptors to the vasopressinergic neurons is still being investigated. The secretion of AVP is under the control of several groups of neurons engaged in the central regulation of blood volume and pressure, such as the rostral ventrolateral medulla (RVLM), the caudal ventrolateral medulla (CVLM) [138], and the circumventricular organs [116].
The mRNA encoding cAMP responsive element-binding protein 3-like 1 (CREB3L1)—a transcription factor of the CREB/activating transcription factor (ATF) family—may play an essential role in the enhancement of AVP transcription in the PVN and SON neurons [139]. Therefore, it is probable that all neuroactive compounds affecting CREB3L1 activity may influence the synthesis of AVP in the hypothalamus. Potent stimulators of AVP release during hypovolemia and hypotension include catecholamines, angiotensin II, and mineralocorticoids [140,141]. It has been observed that atrial natriuretic peptide (ANP) participates in the inhibition of AVP secretion during hypervolemia [50,142,143,144].
AVP secretion may be significantly modulated by mineralocorticoids. The SON and PVN neurons express 11β-HSD2 and mineralocorticoid receptor (MR) immunoreactivity, and the neurons immunoreactive for AVP and MR colocalize with those expressing epithelial sodium channels (ENaC). Moreover, it has been shown that the administration of mineralocorticoids induces the upregulation of AVP mRNA [145,146]. Particularly interesting is the interaction of estrogens with vasopressin. Estrogen receptors are expressed in the hypothalamic vasopressinergic neurons and ERβ inhibits the release of vasopressin in the PVN and SON. Because estrogen expression decreases during dehydration and high salt intake, it has been postulated that estrogen availability may have an impact on the regulation of the sensitivity of vasopressinergic neurons to osmotic stress [147].
There is strong evidence that the vasopressinergic system is activated during stress and that the blockade of V1 AVP receptors significantly reduces stress-induced anxiety and depression behavior [147,148,149,150,151,152,153]. Serotonin, Ang II, and cytokines may be engaged in the stimulation of AVP release during various types of stress [44,45,48].
Vasopressin secretion is elevated during hyperthermia, particularly during the heating of the brain thermosensitive region in the preoptic area [154]. In addition, a prominent direct effect of temperature on the temperature-sensitive ion current mediated by the activation of the TRPV1 channels (a six-fold increase in the firing rate of the action potentials) has been demonstrated in vasopressin neurons [155].

4. Oxytocin and Vasopressin Secretion during Cardiovascular Disorders

4.1. Oxytocin Secretion in Cardiovascular Diseases

Extensive experimental data have shown that arterial hypertension, myocardial infarction, and heart failure, or their consequences, such as pain and stress, have a significant impact on the oxytocin system.
Observations on spontaneously hypertensive rats (SHR) and their normotensive counterparts WKY have revealed that SHR has a lower level of oxytocin in the hypothalamus, midbrain, and brainstem, and showed reduced expression of OXY mRNA in the PVN, SON, and the posterior pituitary. They also manifest reduced expression of oxytocin receptors mRNA in the nucleus of the solitary tract (NTS) [156,157,158,159]. Jameson and coworkers reported that the release of OXY from the PVN into the brainstem is reduced in rats exposed to chronic intermittent hypoxia/hypercapnia, and that this effect is associated with the symptoms of human obstructive sleep apnea syndrome (OSAS) and hypertension [27].
The secretion of oxytocin is significantly affected in various models of hypertension and heart failure. The activation of the OXY neurons in the PVN prevents the development of arterial hypertension in the animal model of OSAS [27]. Downregulation of the OXTR protein and reduction of OXTR mRNA expression was found in the right ventricle of rats with pulmonary hypertension [160]. Several studies have demonstrated a correlation between cardiac oxytocin system activity and the development of post-infarction heart failure. In rats with confirmed post-infarction heart failure, a higher expression of OXY mRNA and an elevated OXY protein level were found in the right ventricle, whereas the OXTR protein level was high in both the left and right ventricle [105]. Other authors reported the downregulation of OXTR in the cardiac left ventricle during the early post myocardial infarction period. However, with the development of heart failure, the OXTR expression markedly increased [161]. Interestingly, rats with myocardial infarction also expressed a higher level of OXTR protein in the superior cervical ganglion [162], and this finding suggests the involvement of OXY in the modulation of sympathetic transmission after myocardial infarction.
Pain is a frequent symptom of CVD and plays an essential role in the regulation of the oxytocin system [163,164,165,166]. For instance, neuropathic pain significantly upregulates OXY mRNA expression in the SON and PVN neurons [167]. Pain, induced by tissue incision, decreases OXY content in the PVN, but increases its release in the periaqueductal grey (PAG) and the caudate nucleus [168]. Oxytocin reduces the pain intensity [166] and changes in its secretion may play a role in the regulation of sensitivity to pain in CVD.
Stress is also inherently present in CVD. Experimental studies showed that short-lasting stress increases the expression of OXY mRNA in the hypothalamus of rats [169] and decreases expression of OXTR mRNA in the heart of Sprague Dawley and Lewis rats [170]. Normotensive WKY rats receiving short intracerebroventricular (ICV) applications of OXY responded to stress with greater increases in blood pressure [171]; however, the same study showed that the enhancement of the pressor response to stress by exogenous OXT in normotensive WKY rats resulted from the interaction of OXY with AVP V1a receptors [171]. In another study, chronically stressed WKY rats treated with continuous ICV infusions of OXTR antagonist manifested significantly greater increases in blood pressure after the application of acute stress [172].
To date, few researchers have investigated the secretion of oxytocin in cardiovascular disorders in humans. Sivukhina et al. reported no significant differences in OXY-positive neuron expression in PVN and SON in the brains of patients, who died from chronic heart failure [173]. More recent findings on patients undergoing aortic valve replacement surgery revealed that myocardial ischemia is associated with significant changes in the expression of several genes involved in the oxytocin signaling pathway. Importantly, myocardial ischemia suppressed oxytocin signaling more effectively in females than in males [174].

4.2. Vasopressin Secretion in Cardiovascular Diseases

In both humans and animals, hypertension and heart failure are associated with an elevated secretion of vasopressin [175,176,177]. This is partly caused by the altered action of signals incoming from the cardiovascular baroreceptors, volume receptors, and chemoreceptors, and partly by other challenges generated by the disease, such as water–electrolyte and acid–base balance disorders, pain, and stress.
In experimental models of arterial hypertension, AVP content and its gene expression are elevated in the PVN, SON, and NTS [178,179]. There is evidence that several components of the renin-angiotensin system (RAS) and mineralocorticoids are involved in the activation of the vasopressin system in hypertension [45,179,180]. The engagement of mineralocorticoids is suggested by studies showing an abundant expression of mineralocorticoid receptors (MR) and α, β, and γ subunits of the ENaC in the cardiovascular brain regions innervated by vasopressinergic neurons, as well as by an upregulation of AVP mRNA and V1aR after the administration of mineralocorticoids [145,180,181]. It is worth noting that mineralocorticoids induce a greater expression of AVP, V1aR, and MR in SHR than in WKY rats [146].
Moreover, experimental data revealed that heart failure is associated with significant changes of AVP content in cardiovascular brain regions. Specifically, rats with heart failure produced by an overloading of the left ventricle manifested significantly elevated AVP levels in the PVN, the SON, the parafascicular nucleus, and the NTS. Simultaneously, they reduced the AVP content in the locus coeruleus. These changes were significantly attenuated by the blockade of AT1 receptors (AT1R) or the inhibition of angiotensin-converting enzyme (ACE), suggesting the engagement of the RAS in the activation of the vasopressinergic system in heart failure [177,182].
Among the factors that may enhance the release of AVP in the heart is hypoxia, which stimulates both the central and the systemic components of the vasopressin system [176,183,184]. Hypoxia activates the catecholaminergic neurons of the CVLM projecting to the PVN, and the pressor response evoked from the carotid bodies involves the stimulation of adrenoreceptors in this nucleus [185,186]. Furthermore, the release of AVP in response to hypoxia is mediated by the activation of the carotid body [137].
Because pain and stress are strong stimulators of the vasopressinergic system [130,166,169,187,188,189,190,191,192], it is very likely that they play an essential role in the stimulation of vasopressin release in CVD. It has been suggested that the activation of hypothalamic vasopressinergic neurons by pain is regulated by norepinephrine and the stimulation of noradrenergic alpha receptors [192].
Patients with cardiovascular pathologies, predominantly heart failure, but also myocardial infarction and arterial hypertension, present elevated plasma AVP levels [175,193]. Population-based data have shown elevated plasma AVP levels and its surrogate marker copeptin [10] in various types of heart failure, acute myocardial infarction, diabetes development, and hypertension [11,12,194,195,196,197]. For this reason, copeptin can be used in clinical practice as a predictive marker in cardiovascular diseases. Corresponding to findings from animal-derived models, higher AVP immunoreactivity in the PVN, the SON, and the posterior pituitary was also found in patients who died from chronic heart failure [173].

5. Cardiovascular Effects of Vasopressin and Oxytocin

As mentioned above, the secretion of oxytocin and vasopressin is regulated by the same stimuli in many instances and both peptides act through receptors engaging the same intracellular pathways. As emphasized below, they can exert complementary actions, either through intensification or abatement of their own regulatory effects.

5.1. Effects of Oxytocin on the Cardiovascular System

Information about the effects of oxytocin on the cardiovascular system has been provided by a number of experimental studies. Oxytocin receptors are present in the heart, in the vessels, and in the cardiovascular regions of the brain [35,109,198,199,200,201]. In the vascular endothelial cells of large vessels and cardiac microvessels, they co-localize with endothelial nitric oxide (NO) synthase (eNOS) [106]. In the estrogen-sensitive regions of the brain, such as the hypothalamus and the amygdala, the induction and transcription of OXTR are regulated by estrogens and depend on the stimulation of estrogen receptors (ER). Stimulation of ERα enhances induction of OXTR, while activation of ERβ helps to maintain the transcription of OXTR [113,202].
The majority of experimental findings indicate that the administration of oxytocin exerts a hypotensive effect and that the magnitude of this effect depends on its central and peripheral actions (Figure 2). In normotensive rats remaining at rest, ICV administration of OXY significantly decreases blood pressure, but does not influence the heart rate [203,204]. It has been shown that the subchronic (5 days) subcutaneous application of OXY increases the hypotensive action of clonidine, and this action is abolished by the blockade of alpha-2 receptors (α2R). The latter finding suggests the involvement of α-2 noradrenergic receptors in the hypotensive action of OXY [205]. In Sprague Dawley rats, OXTR mRNA was detected in the vena cava, the pulmonary vein, and the aorta. Its presence was associated with a high expression of ANP mRNA in the vena cava and the pulmonary vein. OXTR mRNA was also identified in the mesenteric arteries and the uterine arcuate arteries of female Sprague Dawley rats. The expression of OXTR mRNA was enhanced by estrogen in the aorta and the vena cava [103,206] and it is worth noting that oxytocin did not cause vasodilation of the systemic arteries. In contrast, in high concentrations, OXY elicited vasoconstriction, which was mediated by the stimulation of V1aR [206].
Transcripts of OXTR were demonstrated both in the atrial and ventricular sections of the heart. It has been shown that activation of the oxytocin receptors plays an essential role in the release of ANP in the heart [34]. In vitro observations on pluripotent P19 embryonic stem cells have shown that OXY induces differentiation of these cells into cardiomyocytes [207].
Extensive findings from animal-derived models indicate that oxytocin may play a beneficial role in the regulation of cardiovascular functions in hypertension. In particular, peripheral subcutaneous injections of OXY in SHR rats resulted in hypotension and caused greater diurnal and nocturnal reductions in blood pressure [208,209]. It appears that the beneficial effect of oxytocin on blood pressure develops during the early postnatal period [210,211]. The buffering role of oxytocin in the regulation of blood pressure in SHR is impaired and even reversed during exposure to an alarming stress [171,172]. SHR rats respond to the application of acute stress with greater increases in blood pressure than WKY rats and this difference is abolished by the chronic ICV administration of OXY [172]. Hypertension is associated with the altered expression of oxytocin in the central nervous system and the adrenal glands of SHR [156]. Reduced expression of OXTR was found in SHR in the NTS and the dorsal brain stem [157,158]. More recently, it was shown that oxytocin released by neurons projecting from the PVN to the dorsal motor nucleus of the vagus prevents the development of hypertension in rats chronically exposed to hypoxia/hypercapnia—an animal model of human hypertension associated with OSAS [27].
Some studies suggest the engagement of oxytocin in hypertension-induced cardiac remodeling. In particular, it has been found that the blockade of cardiac OXTR facilitates the development of fibrosis and causes a reduction in the LV ejection fraction in SHR rats [200]. SHR also manifests elevated levels of OXY and OXTR in the left ventricle of the heart [212].
The cardiac oxytocinergic system plays an essential role in cardiac remodeling and cardioprotection during cardiac ischemia and post-infarct heart failure [161,213,214,215,216,217,218,219]. In the isolated heart rat model, submitted to a one hour ischemia-reperfusion-perfusion procedure, the administration of oxytocin during the early reperfusion phase reduced the size of the myocardial infarct and increased the coronary blood flow. This was associated with a decreased production of reactive oxygen species (ROS), a reduced expression of pro-inflammatory cytokines (TNF-α, IL-1β, IL-6), and reduced apoptosis, reduced arrhythmia, and reduced ventricular fibrillation [217,218,219,220]. In addition, OXY increased the phosphorylated Hsp27 protein level and the atrial natriuretic peptide level in the cardiac left ventricle [217]. It should be noted that the cardioprotective actions of OXY are mediated through OXTR activation. The infusion of oxytocin elevated cell death in siRNA-mediated knockout cells with reduced expression of OXTR via binding with vasopressin receptors [220]. Centrally administered OXY decreases both the pressor response and the tachycardic response after exposure to an alarming stress in rats with myocardial infarction, but not in the sham-operated controls [221].
In the rat model of heart failure induced by pressure overloading of the left ventricle lasting four weeks, chronic stimulation of PVN neurons in adult Sprague Dawley rats resulted in several beneficial effects, such as a decrease in blood pressure and heart rate, a reduction of cellular hypertrophy, a reduction of fibrosis, a reduction of IL-1β concentration, and an improvement of the LV function [222]. Gutkowska et al. showed that oxytocin may prevent the development of diabetic cardiomyopathy. A downregulation of OXY gene expression in the heart of the db/db mouse model of type 2 diabetes mellitus has been reported [223]. Male db/db mice respond with a higher expression of cardiac OXTR, a lower fasting blood glucose level, reduced body fat accumulation, reduced ROS production, reduced cardiomyocyte hypertrophy, reduced fibrosis, and reduced apoptosis to prolonged (12 weeks) administration of oxytocin [215,224]. However, chronic overstimulation of OXTR may result in opposite and highly detrimental effects. It has been shown that 14-week-old transgenic α-MHC-Oxtr mice overexpressing oxytocin receptors manifest significantly greater left ventricle end-diastolic volume, prominent cardiac fibrosis, and enhanced expression of pro-fibrogenic genes. There is also a higher mortality rate among transgenic α-MHC-Oxtr mice than among the wild-type control strain [225].
To date, little is known about the cardiovascular effects of OXY in humans and further studies are needed. In line with experimental findings, OXY and its analogue carbetocin exert hypotensive effects during delivery. Rosseland et al. reported marked hypotension and an increase in the stroke volume after OXY/carbetocin administration during cesarean section [226]. Furthermore, in another study, the intravenous infusion of OXY during cesarean section resulted in stenocardia, transient profound tachycardia, hypotension, and electrocardiographic signs of myocardial ischemia [227]. It is worth noting that intranasal OXY administration caused a short-term increase in heart-rate variability in healthy volunteers [228].

5.2. Effects of Vasopressin on the Cardiovascular System

Vasopressin regulates blood pressure, blood flow, and body fluid volume through actions mediated by receptors located on vasopressin sensitive cardiovascular neurons in the brain and in the spinal cord, and on AVP receptors in numerous peripheral organs [29,60,69,73,123,125,126,130,175,229,230,231,232,233] (Figure 3). All types of vasopressin receptors are engaged in these actions. The central pressor action of vasopressin is mediated mainly by V1aR located in the RVLM and the rostral ventral respiratory column [234,235,236].
Stimulation of V1aR also plays an essential role in the vasoconstriction of systemic, coronary, and renal vessels [237,238,239,240]. However, stimulation of V1aR also causes vasodilation of the pulmonary arteries [241]. In the kidney, V1aR is present in the basolateral membranes of the thick ascending limbs of Henle’s loop (TAL) and the collecting ducts (CD), and in the juxtaglomerular apparatus [242,243]. Experiments on V1aR deficient mice (V1aR−/−) strongly suggest that activation of V1aR in the kidney plays an essential role in the stimulation of the RAS and in the activation of the V2R-AQP2 pathway. Elimination of V1aR in these mice resulted in polyuria and the diminution of renin-containing granular cells and the plasma renin level, as well as in the reduction of V2R expression in the collecting ducts [244,245]. Downregulation of V1a receptors is observed in the vascular smooth muscles in sepsis elicited by the application of lipopolysaccharide or proinflammatory cytokines, and this may account for the reduced responsiveness to vasopressin in septic shock [246].
V1bR is located mainly in the brain, the pituitary, the pancreas, and the kidneys [55,247,248]. In the brain, V1bR mediates the activation of the PVN presympathetic neurons by vasopressin [235,249]. They are also engaged in the stimulation of insulin release from the pancreatic islet cells, and the release of ACTH from the anterior pituitary [250,251]. Expression of V1b receptor mRNA in the anterior pituitary is mediated by glucocorticoids [250]. V1bR is located mainly in the inner medulla in the kidney. There is evidence that stimulation of renal V1bR can counterbalance the antidiuretic effect of vasopressin mediated by V2R [252]. It is likely that the counteractive effects of the stimulation of V1bR and V2 receptors by vasopressin may account for the potent natriuretic effect observed after the combined administration of insulin and vasopressin in rats [253].
Vasopressin has long been known as an antidiuretic hormone due to its water retaining action in the kidney, exerted by V2 receptors. In the renal tubules, V2R is located mainly in the collecting duct cells, where it is coupled to Gs proteins. The stimulation of V2R results in the activation of adenylate cyclase, the formation of cAMP, and the activation of protein kinase A (PKA). This leads to the fast activation of AQP2 molecules and their translocation from the cytoplasmic vesicles to the apical cellular membrane where they form water channels. The channels enable the shift of water from the tubules to the cells and subsequently to the basolateral membrane. PKA also intensifies the synthesis of AQP2 molecules [254,255,256]. Activation of V2R increases water permeability in the kidney and enhances the shift of urea and sodium. Acting on V2 receptors, AVP enhances the transcription of the epithelial sodium channels (ENaC) and urea transporters, and promotes the synthesis of renin [254,255,256,257,258]. Consequently, the stimulation of V2R reduces water and sodium excretion and their blockade results in diuresis and natriuresis. The effectiveness of these actions depends on the state of hydration of the body and the availability of other factors operating in the renal tubules, such as Ang II and aldosterone. Aldosterone cooperates with AVP in the regulation of sodium transport via the epithelial sodium channels in the distal portion of the nephron [259]. Experiments on rats with diabetes insipidus have shown that aldosterone reduces the expression of AQP2 in the inner stripe of the outer medulla and increases free water clearance and creatinine clearance, whereas the mineralocorticoid receptor antagonist, spironolactone, elevates AQP2 expression and decreases creatinine clearance [260]. The stimulation of V2 receptors causes their internalization and recycling. This recycling is associated with increased sensitivity of the collecting ducts to vasopressin. The process of recycling is intensified in cardiac failure induced by the ligation of the left anterior descending coronary artery and may account for increased sensitivity to the antidiuretic action of vasopressin in heart failure [261].
Although the kidney is the main target organ expressing V2R, studies exploring the expression of V2R mRNA and the functional consequences of V2R blockade have shown that these receptors may also play a role in the regulation of the gastrointestinal tract, the heart, and the brain [37,58,80,236,262]. Experiments on embryonic stem cells have shown that vasopressin contributes to cardiac differentiation and its action is accomplished by the stimulation of V2 receptors and NO synthesis [263].
The blockade of vasopressin receptors may exert beneficial effects in the treatment of various forms of cardiac failure. The application of antagonists selectively blocking V2 receptors in the rat induces a significant aquaphoric effect and improves hemodynamics [264]. Experiments on pigs with cardiac failure induced by rapid pacing showed that the systemic administration of the V1a receptor antagonist (SR49059) elicited beneficial effects illustrated by a reduction of the left ventricle dimension and peak wall stress, although it did not improve cardiac muscle contractility. The authors reported that the effectiveness of the treatment of heart failure was significantly greater after simultaneous application of V1aR and angiotensin II antagonists [265]. The altered expression of V1aR has been reported in rats with myocardial infarction and in rats exposed to chronic stress. The infarcted rats manifested a reduced V1aR mRNA expression in the preoptic, diencephalic, and mesocenphalopontine regions of the brain. In contrast, the expression of V1aR receptors was significantly elevated in the brain medulla and in the renal cortex of chronically stressed rats [266]. Taking into account the pressor effect of the stimulation of V1a receptors in the brain, it is likely that the downregulation of these receptors during the post-infarct state is a protective response, which is suppressed during stress. There is evidence that the stimulation of V1aR may reduce the cerebral blood flow during dehydration and that this response is associated with an increased production of ROS and can be effectively abolished by the blockade of V1aR [267].
The results from clinical observations are in line with the experimental data. Patients with heart failure manifest a significantly greater expression of V1aR in the left ventricle myocardium than subjects with non-failing hearts [268]. To date, the application of various types of V1aR and V2R antagonists in the treatment of heart failure has provided inconclusive results in human clinical trials [269,270,271]. Nonetheless, recent clinical guidelines on the diagnosis and the treatment of heart failure issued by the European Society of Cardiology indicate that tolvaptan, a selective V2R antagonist, can be considered for the treatment of heart failure patients with hyponatremia [272]. Recently, promising results have been obtained in a double-blind, randomized phase II AVANTI study following the effects of the application of pecavaptan, a dual V1a/V2 antagonist in patients with heart failure and fluid overload. This study revealed that treatment with pecavaptan significantly enhances decongestion of the heart and improves the maintenance of the body fluid balance in patients with acute heart failure [273].

6. Conclusions

  • The secretion of oxytocin is elevated by reproductive and maternal stimuli and by hyperosmolality, hypovolemia, hemorrhage, pain, and stress. The process of stimulation is related to the action of physiochemical factors, neurotransmitters (catecholamines, serotonin), neuropeptides, glucocorticoids, and estrogens. The secretion of oxytocin is inhibited by GABA, orexins, prolactin, and opioid peptides.
  • The secretion of vasopressin is stimulated by hypernatremia and the activation of TRPV1 channels, and by hypovolemia, hypotension, hypoxia, hyperthermia, pain, and stress. The stimulation is mediated by neurotransmitters (catecholamines, serotonin), neuropeptides (ANG II, cytokines), and mineralocorticoids. Hyponatremia, hypervolemia, and estrogens inhibit the secretion of vasopressin.
  • The cardiovascular system is regulated jointly by oxytocin and vasopressin and their action can be synergic or antagonistic.
  • Oxytocin exerts a hypotensive action, increases the production of vasodilatory compounds (ANP, NO), decreases pressor responses to hypoxia and stress, and exerts several cardioprotective effects in hypertension, cardiac ischemia, and cardiomyopathies. During post-infarct heart failure, oxytocin reduces the production of ROS and inflammatory cytokines, decreases apoptosis, decreases arrhythmia, decreases ventricular fibrillation, and reduces cardiovascular responses to stress.
  • Vasopressin increases the retention of water and sodium and exerts a pressor effect, which is mediated by the stimulation of the sympathetic nervous system and direct vasoconstrictive action exerted in the majority of the vascular beds. It also interacts with pressor and hypotensive compounds (angiotensins, mineralocorticoids, NO, ANP). Cardiovascular responses to vasopressin at rest and during stress are enhanced in hypertension and heart failure. Experimental and clinical studies show that the blockade of vasopressin receptors may exert beneficial effects in patients with heart failure and fluid overload.
  • To summarize, the present review emphasizes the important role of oxytocin and vasopressin in the regulation of the cardiovascular system. Vasopressin and oxytocin play complementary roles in the regulation of cardiovascular parameters, exerting synergic and antagonistic effects in the brain, in the vessels, and in the kidney. Evidence from experimental observations exploring the role of oxytocin and from experimental and clinical studies analyzing the role of vasopressin strongly suggests that the application of either oxytocin or vasopressin antagonists may exert favorable effects in cardiovascular diseases and extensive clinical studies are necessary to confirm this conclusion.

Funding

The study was supported by the Medical University of Warsaw (1MA/N/2021) and carried out with the use of CEPT infrastructure financed by the European Union—The European Regional Development Fund within the Operational Programme “Innovative Economy” for 2007–2013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors would like to express their gratitude to Marcin Kumosa for the preparation of the figures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boissonnas, R.A.; Guttman, S.T. Chemistry of the Neurohypophysial Hormones. In Handbook of Experimental Pharmakologie; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1968; Volume 23, pp. 40–66. [Google Scholar]
  2. Feldman, R.; Monakhov, M.; Pratt, M.; Ebstein, R.P. Oxytocin pathway genes: Evolutionary ancient system impacting on human affiliation, sociality, and psychopathology. Biol. Psychiatry 2016, 79, 174–184. [Google Scholar] [CrossRef]
  3. Gimpl, G.; Fahrenholz, F. The oxytocin receptor system: Structure, function, and regulation. Physiol. Rev. 2001, 81, 629–683. [Google Scholar] [CrossRef] [Green Version]
  4. Gruber, C.W. Physiology of invertebrate oxytocin and vasopressin neuropeptides. Exp. Physiol. 2014, 99, 55–61. [Google Scholar] [CrossRef]
  5. Armstrong, W.E.; Wang, L.; Li, C.; Teruyama, R. Performance, properties and plasticity of identified oxytocin and vasopressin neurones in vitro. J. Neuroendocrinol. 2010, 22, 330–342. [Google Scholar] [CrossRef] [Green Version]
  6. Boccia, M.L.; Petrusz, P.; Suzuki, K.; Marson, L.; Pedersen, C.A. Immunohistochemical localization of oxytocin receptors in human brain. Neuroscience 2013, 253, 155–164. [Google Scholar] [CrossRef]
  7. Buijs, R.M.; Swaab, D.F.; Dogterom, J.; van Leeuwen, F.W. Intra- and extrahypothalamic vasopressin and oxytocin pathways in the rat. Pathways to the limbic system, medulla oblongata and spinal cord. Cell Tissue Res. 1978, 192, 423–435. [Google Scholar] [CrossRef]
  8. Leng, G.; Pineda, R.; Sabatier, N.; Ludwig, M. 60 YEARS OF NEUROENDOCRINOLOGY: The posterior pituitary, from Geoffrey Harris to our present understanding. J. Endocrinol. 2015, 226, T173–T185. [Google Scholar] [CrossRef] [Green Version]
  9. Acher, R.; Chauvet, J. Structure, processing and evolution of the neurohypophysial hormone-neurophysin precursors. Biochimie 1988, 70, 1197–1207. [Google Scholar] [CrossRef]
  10. Morgenthaler, N.G.; Struck, J.; Alonso, C.; Bergmann, A. Assay for the measurement of copeptin, a stable peptide derived from the precursor of vasopressin. Clin. Chem. 2006, 52, 112–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Nickel, C.H.; Bingisser, R.; Morgenthaler, N.G. The role of copeptin as a diagnostic and prognostic biomarker for risk stratification in the emergency department. BMC Med. 2012, 20, 10–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Reinstadler, S.J.; Klug, G.; Feistritzer, H.J.; Metzler, B.; Mair, J. Copeptin testing in acute myocardial infarction: Ready for routine use? Dis. Markers 2015, 2015, 614145. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, R.; Zhang, H.F.; Han, J.S.; Han, S.P. Genes Related to Oxytocin and Arginine-Vasopressin Pathways: Associations with Autism Spectrum Disorders. Neurosci. Bull. 2017, 33, 238–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bankir, L.; Bichet, D.G.; Morgenthaler, N.G. Vasopressin: Physiology, assessment and osmosensation. J. Intern Med. 2017, 282, 284–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Bichet, D.G. Vasopressin at central levels and consequences of dehydration. Ann. Nutr. Metab. 2016, 68 (Suppl. 2), 19–23. [Google Scholar] [CrossRef]
  16. Bergmann, W. Neurohypophysis. Structure and Function. In Handbook of Experimental Pharmakologie; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1968; Volume 23, pp. 1–39. [Google Scholar]
  17. Ball, S.G.; Barber, T.; Baylis, P.H. Tests of posterior pituitary function. J. Endocrinol. Investig. 2003, 26 (Suppl. 7), 15–24. [Google Scholar]
  18. Bealer, S.L.; Armstrong, W.E.; Crowley, W.R. Oxytocin release in magnocellular nuclei: Neurochemical mediators and functional significance during gestation. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2010, 299, R452–R458. [Google Scholar] [CrossRef]
  19. Bichet, D.G. Central vasopressin: Dendritic and axonal secretion and renal actions. Clin. Kidney J. 2014, 7, 242–247. [Google Scholar] [CrossRef] [Green Version]
  20. Burnstock, G. Purinergic signalling in endocrine organs. Purinergic. Signal. 2014, 10, 189–231. [Google Scholar] [CrossRef] [Green Version]
  21. de Kock, C.P.; Wierda, K.D.; Bosman, L.W.; Min, R.; Koksma, J.J.; Mansvelder, H.D.; Verhage, M.; Brussaard, A.B. Somatodendritic secretion in oxytocin neurons is upregulated during the female reproductive cycle. J. Neurosci. 2003, 23, 2726–2734. [Google Scholar] [CrossRef] [Green Version]
  22. Mohr, E.; Richter, D. Local synthesis of the rat Vasopressin precursor in dendrites of in vitro cultured nerve cells. Brain Res. Mol. Brain Res. 2003, 114, 115–122. [Google Scholar] [CrossRef]
  23. Ventura, R.R.; Gomes, D.A.; Reis, W.L.; Elias, L.L.; Castro, M.; Valença, M.M.; Carnio, E.C.; Rettori, V.; McCann, S.M.; Antunes-Rodrigues, J. Nitrergic modulation of vasopressin, oxytocin and atrial natriuretic peptide secretion in response to sodium intake and hypertonic blood volume expansion. Braz. J. Med. Biol. Res. 2002, 35, 1101–1109. [Google Scholar] [CrossRef] [PubMed]
  24. Freeman, S.M.; Smith, A.L.; Goodman, M.M.; Bales, K. Selective localization of oxytocin receptors and vasopressin 1a receptors in the human brainstem. Soc. Neurosci. 2017, 12, 113–123. [Google Scholar] [CrossRef] [Green Version]
  25. Freeman, S.M.; Young, L.J. Comparative Perspectives on Oxytocin and Vasopressin Receptor Research in Rodents and Primates: Translational Implications. J. Neuroendocrinol. 2016, 28. [Google Scholar] [CrossRef] [PubMed]
  26. Gilbey, M.P.; Coote, J.H.; Fleetwood-Walker, S.; Peterson, D.F. The influence of the paraventriculo-spinal pathway, and oxytocin and vasopressin on sympathetic preganglionic neurones. Brain Res. 1982, 251, 283–290. [Google Scholar] [CrossRef]
  27. Jameson, H.; Bateman, R.; Byrne, P.; Dyavanapalli, J.; Wang, X.; Jain, V.; Mendelowitz, D. Oxytocin neuron activation prevents hypertension that occurs with chronic intermittent hypoxia/hypercapnia in rats. Am. J. Physiol. Heart Circ. Physiol. 2016, 310, H1549–H1557. [Google Scholar] [CrossRef] [PubMed]
  28. Motawei, K.; Pyner, S.; Ranson, R.N.; Kamel, M.; Coote, J.H. Terminals of paraventricular spinal neurones are closely associated with adrenal medullary sympathetic preganglionic neurones: Immunocytochemical evidence for vasopressin as a possible neurotransmitter in this pathway. Exp. Brain Res. 1999, 126, 68–76. [Google Scholar] [CrossRef]
  29. Murphy, D.; Konopacka, A.; Hindmarch, C.; Paton, J.F.; Sweedler, J.V.; Gillette, M.U.; Ueta, Y.; Grinevich, V.; Lozic, M.; Japundzic-Zigon, N. The hypothalamic-neurohypophyseal system: From genome to physiology. J. Neuroendocrinol. 2012, 24, 539–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Peters, S.; Kreulen, D.L. Vasopressin-mediated slow EPSPs in a mammalian sympathetic ganglion. Brain Res. 1985, 339, 126–129. [Google Scholar] [CrossRef]
  31. Piñol, R.A.; Jameson, H.; Popratiloff, A.; Lee, N.H.; Mendelowitz, D. Visualization of oxytocin release that mediates paired pulse facilitation in hypothalamic pathways to brainstem autonomic neurons. PLoS ONE. 2014, 9, e112138. [Google Scholar] [CrossRef] [Green Version]
  32. Rogers, C.N.; Ross, A.P.; Sahu, S.P.; Siegel, E.R.; Dooyema, J.M.; Cree, M.A.; Stopa, E.G.; Young, L.J.; Rilling, J.K.; Albers, H.E.; et al. Oxytocin- and arginine vasopressin-containing fibers in the cortex of humans, chimpanzees, and rhesus macaques. Am. J. Primatol. 2018, 80, e22875. [Google Scholar] [CrossRef] [PubMed]
  33. Young, L.J.; Toloczko, D.; Insel, T.R. Localization of vasopressin (V1a) receptor binding and mRNA in the rhesus monkey brain. J. Neuroendocrinol. 1999, 11, 291–297. [Google Scholar] [CrossRef]
  34. Gutkowska, J.; Jankowski, M.; Lambert, C.; Mukaddam-Daher, S.; Zingg, H.H.; McCann, S.M. Oxytocin releases atrial natriuretic peptide by combining with oxytocin receptors in the heart. Proc. Natl. Acad. Sci. USA 1997, 94, 11704–11709. [Google Scholar] [CrossRef] [Green Version]
  35. Jankowski, M.; Hajjar, F.; Kawas, S.A.; Mukaddam-Daher, S.; Hoffman, G.; McCann, S.M.; Gutkowska, J. Rat heart: A site of oxytocin production and action. Proc. Natl. Acad. Sci. USA 1998, 95, 14558–14563. [Google Scholar] [CrossRef] [Green Version]
  36. Mohan, S.; Khan, D.; Moffett, R.C.; Irwin, N.; Flatt, P.R. Oxytocin is present in islets and plays a role in beta-cell function and survival. Peptides 2018, 100, 260–268. [Google Scholar] [CrossRef]
  37. Monstein, H.J.; Truedsson, M.; Ryberg, A.; Ohlsson, B. Vasopressin receptor mRNA expression in the human gastrointestinal tract. Eur. Surg. Res. 2008, 40, 34–40. [Google Scholar] [CrossRef] [PubMed]
  38. Moritoh, S.; Sato, K.; Okada, Y.; Koizumi, A. Endogenous arginine vasopressin-positive retinal cells in arginine vasopressin-eGFP transgenic rats identified by immunohistochemistry and reverse transcriptase-polymerase chain reaction. Mol. Vis. 2011, 17, 3254–3261. [Google Scholar] [PubMed]
  39. Boyle, W.A., 3rd; Segel, L.D. Attenuation of vasopressin-mediated coronary constriction and myocardial depression in the hypoxic heart. Circ. Res. 1990, 66, 710–721. [Google Scholar] [CrossRef] [Green Version]
  40. Coldren, K.M.; Li, D.P.; Kline, D.D.; Hasser, E.M.; Heesch, C.M. Acute hypoxia activates neuroendocrine, but not presympathetic, neurons in the paraventricular nucleus of the hypothalamus: Differential role of nitric oxide. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2017, 312, R982–R995. [Google Scholar] [CrossRef]
  41. McCann, S.M.; Antunes-Rodrigues, J.; Franci, C.R.; Anselmo-Franci, J.A.; Karanth, S.; Rettori, V. Role of the hypothalamic pituitary adrenal axis in the control of the response to stress and infection. Braz. J. Med. Biol. Res. 2000, 33, 1121–1131. [Google Scholar] [CrossRef] [Green Version]
  42. Ruyle, B.C.; Klutho, P.J.; Baines, C.P.; Heesch, C.M.; Hasser, E.M. Hypoxia activates a neuropeptidergic pathway from the paraventricular nucleus of the hypothalamus to the nucleus tractus solitarii. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2018, 315, R1167–R1182. [Google Scholar] [CrossRef]
  43. Stocker, S.D.; Wilson, M.E.; Madden, C.J.; Lone, U.; Sved, A.F. Intravenous 6-hydroxydopamine attenuates vasopressin and oxytocin secretion stimulated by hemorrhage and hypotension but not hyperosmolality in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006, 291, R59–R67. [Google Scholar] [CrossRef] [Green Version]
  44. Szczepanska-Sadowska, E.; Czarzasta, K.; Cudnoch-Jedrzejewska, A. Dysregulation of the Renin-Angiotensin System and the Vasopressinergic System Interactions in Cardiovascular Disorders. Curr. Hypertens. Rep. 2018, 20, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Szczepanska-Sadowska, E.; Zera, T.; Sosnowski, P.; Cudnoch-Jedrzejewska, A.; Puszko, A.; Misicka, A. Vasopressin and Related Peptides; Potential Value in Diagnosis, Prognosis and Treatment of Clinical Disorders. Curr. Drug Metab. 2017, 18, 306–345. [Google Scholar] [CrossRef] [PubMed]
  46. Tanoue, A. New topics in vasopressin receptors and approach to novel drugs: Effects of vasopressin receptor on regulations of hormone secretion and metabolisms of glucose, fat, and protein. J. Pharmacol. Sci. 2009, 109, 50–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Ebstein, R.P.; Knafo, A.; Mankuta, D.; Chew, S.H.; Lai, P.S. The contributions of oxytocin and vasopressin pathway genes to human behavior. Horm. Behav. 2012, 61, 359–379. [Google Scholar] [CrossRef]
  48. Jørgensen, H.; Knigge, U.; Kjaer, A.; Warberg, J. Serotonergic involvement in stress-induced vasopressin and oxytocin secretion. Eur. J. Endocrinol. 2002, 147, 815–824. [Google Scholar] [CrossRef] [Green Version]
  49. McDonald, N.A.; Kuzmiski, J.B.; Naderi, N.; Schwab, Y.; Pittman, Q.J. Endogenous modulators of synaptic transmission: Cannabinoid regulation in the supraoptic nucleus. Prog. Brain Res. 2008, 170, 129–136. [Google Scholar] [CrossRef] [Green Version]
  50. Sladek, C.D.; Song, Z. Diverse roles of G-protein coupled receptors in the regulation of neurohypophyseal hormone secretion. J. Neuroendocrinol. 2012, 24, 554–565. [Google Scholar] [CrossRef] [Green Version]
  51. Stoop, R. Neuromodulation by oxytocin and vasopressin. Neuron 2012, 76, 142–159. [Google Scholar] [CrossRef] [Green Version]
  52. Vuong, C.; Van Uum, S.H.; O’Dell, L.E.; Lutfy, K.; Friedman, T.C. The effects of opioids and opioid analogs on animal and human endocrine systems. Endocr. Rev. 2010, 31, 98–132. [Google Scholar] [CrossRef] [Green Version]
  53. Gruber, C.W.; Muttenthaler, M.; Freissmuth, M. Ligand-based peptide design and combinatorial peptide libraries to target G protein-coupled receptors. Curr. Pharm. Des. 2010, 16, 3071–3088. [Google Scholar] [CrossRef] [Green Version]
  54. Liu, J.; Wess, J. Different single receptor domains determine the distinct G protein coupling profiles of members of the vasopressin receptor family. J. Biol. Chem. 1996, 271, 8772–8778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lolait, S.J.; O’Carroll, A.M.; Brownstein, M.J. Molecular biology of vasopressin receptors. Ann. N. Y. Acad. Sci. 1995, 771, 273–292. [Google Scholar] [CrossRef] [PubMed]
  56. Wettschureck, N.; Moers, A.; Hamalainen, T.; Lemberger, T.; Schütz, G.; Offermanns, S. Heterotrimeric G proteins of the Gq/11 family are crucial for the induction of maternal behavior in mice. Mol. Cell Biol. 2004, 24, 8048–8054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Datta, A.; Yang, C.R.; Limbutara, K.; Chou, C.L.; Rinschen, M.M.; Raghuram, V.; Knepper, M.A. PKA-independent vasopressin signaling in renal collecting duct. FASEB J. 2020, 34, 6129–6146. [Google Scholar] [CrossRef]
  58. Juul, K.V.; Bichet, D.G.; Nielsen, S.; Nørgaard, J.P. The physiological and pathophysiological functions of renal and extrarenal vasopressin V2 receptors. Am. J. Physiol. Renal. Physiol. 2014, 306, F931–F940. [Google Scholar] [CrossRef]
  59. Michelini, S.; Urbanek, M.; Dean, M.; Goldman, D. Polymorphism and genetic mapping of the human oxytocin receptor gene on chromosome 3. Am. J. Med. Genet. 1995, 60, 183–187. [Google Scholar] [CrossRef]
  60. Thibonnier, M.; Graves, M.K.; Wagner, M.S.; Auzan, C.; Clauser, E.; Willard, H.F. Structure, sequence, expression, and chromosomal localization of the human V1a vasopressin receptor gene. Genomics 1996, 31, 327–334. [Google Scholar] [CrossRef]
  61. Friedman, E.; Carson, E.; Larsson, C.; DeMarco, L. A polymorphism in the coding region of the vasopressin type 2 receptor (AVPR2) gene. Hum. Mol. Genet. 1993, 2, 1746. [Google Scholar] [CrossRef]
  62. Thibonnier, M.; Graves, M.K.; Wagner, M.S.; Chatelain, N.; Soubrier, F.; Corvol, P.; Willard, H.F.; Jeunemaitre, X. Study of V(1)-vascular vasopressin receptor gene microsatellite polymorphisms in human essential hypertension. J. Mol. Cell Cardiol. 2000, 32, 557–564. [Google Scholar] [CrossRef]
  63. Wilczyński, K.M.; Siwiec, A.; Janas-Kozik, M. Systematic Review of Literature on Single-Nucleotide Polymorphisms Within the Oxytocin and Vasopressin Receptor Genes in the Development of Social Cognition Dysfunctions in Individuals Suffering From Autism Spectrum Disorder. Front. Psychiatry 2019, 10, 380. [Google Scholar] [CrossRef]
  64. Barberis, C.; Tribollet, E. Vasopressin and oxytocin receptors in the central nervous system. Crit. Rev. Neurobiol. 1996, 10, 119–154. [Google Scholar] [CrossRef]
  65. Campos-Lira, E.; Kelly, L.; Seifi, M.; Jackson, T.; Giesecke, T.; Mutig, K.; Koshimizu, T.A.; Hernandez, V.S.; Zhang, L.; Swinny, J.D. Dynamic Modulation of Mouse Locus Coeruleus Neurons by Vasopressin 1a and 1b Receptors. Front. Neurosci. 2018, 12, 919. [Google Scholar] [CrossRef] [PubMed]
  66. Góźdź, A.; Szczepańska-Sadowska, E.; Maśliński, W.; Kumosa, M.; Szczepańska, K.; Dobruch, J. Differential expression of vasopressin V1a and V1b receptors mRNA in the brain of renin transgenic TGR(mRen2)27 and Sprague-Dawley rats. Brain Res. Bull. 2003, 59, 399–403. [Google Scholar] [CrossRef]
  67. Griffiths, P.R.; Lolait, S.J.; Harris, L.E.; Paton, J.F.R.; O’Carroll, A.M. Vasopressin V1a receptors mediate the hypertensive effects of [Pyr(1)]apelin-13 in the rat rostral ventrolateral medulla. J. Physiol. 2017, 595, 3303–3318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Jackiewicz, E.; Szczepanska-Sadowska, E.; Dobruch, J. Altered expression of angiotensin AT1a and vasopressin V1a receptors and nitric oxide synthase mRNA in the brain of rats with renovascular hypertension. J. Physiol. Pharmacol. 2004, 55, 725–737. [Google Scholar]
  69. Lolait, S.J.; O’Carroll, A.M.; Mahan, L.C.; Felder, C.C.; Button, D.C.; Young, W.S., 3rd; Mezey, E.; Brownstein, M.J. Extrapituitary expression of the rat V1b vasopressin receptor gene. Proc. Natl. Acad. Sci. USA 1995, 92, 6783–6787. [Google Scholar] [CrossRef] [Green Version]
  70. Ostrowski, N.L.; Lolait, S.J.; Bradley, D.J.; O’Carroll, A.M.; Brownstein, M.J.; Young, W.S., 3rd. Distribution of V1a and V2 vasopressin receptor messenger ribonucleic acids in rat liver, kidney, pituitary and brain. Endocrinology 1992, 131, 533–535. [Google Scholar] [CrossRef]
  71. Vaccari, C.; Lolait, S.J.; Ostrowski, N.L. Comparative distribution of vasopressin V1b and oxytocin receptor messenger ribonucleic acids in brain. Endocrinology 1998, 139, 5015–5033. [Google Scholar] [CrossRef] [PubMed]
  72. Young, W.S.; Li, J.; Wersinger, S.R.; Palkovits, M. The vasopressin 1b receptor is prominent in the hippocampal area CA2 where it is unaffected by restraint stress or adrenalectomy. Neuroscience 2006, 143, 1031–1039. [Google Scholar] [CrossRef] [Green Version]
  73. Gutkowska, J.; Miskurka, M.; Danalache, B.; Gassanov, N.; Wang, D.; Jankowski, M. Functional arginine vasopressin system in early heart maturation. Am. J. Physiol. Heart Circ. Physiol. 2007, 293, H2262–H2270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Horn, A.M.; Lightman, S.L. Vasopressin-stimulated turnover of phosphatidylinositol in the decentralised superior cervical ganglion of the rat. Brain Res. 1988, 455, 18–23. [Google Scholar] [CrossRef]
  75. Kiraly, M.; Audigier, S.; Tribollet, E.; Barberis, C.; Dolivo, M.; Dreifuss, J.J. Biochemical and electrophysiological evidence of functional vasopressin receptors in the rat superior cervical ganglion. Proc. Natl. Acad. Sci. USA 1986, 83, 5335–5339. [Google Scholar] [CrossRef] [Green Version]
  76. Klenerova, V.; Chottova-Dvorakova, M.; Skopek, P.; Sida, P.; Mistrova, E.; Slavikova, J.; Hynie, S. Expression of heart oxytocin receptor and its mRNA in two rat strains with different activity of HPA axis. Neuro Endocrinol. Lett. 2011, 32, 805–810. [Google Scholar] [PubMed]
  77. Qiu, F.; Qiu, C.Y.; Cai, H.; Liu, T.T.; Qu, Z.W.; Yang, Z.; Li, J.D.; Zhou, Q.Y.; Hu, W.P. Oxytocin inhibits the activity of acid-sensing ion channels through the vasopressin, V1A receptor in primary sensory neurons. Br. J. Pharmacol. 2014, 171, 3065–3076. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Zera, T.; Przybylski, J.; Grygorowicz, T.; Kasarello, K.; Podobińska, M.; Mirowska-Guzel, D.; Cudnoch-Jedrzejewska, A. Vasopressin V1a receptors are present in the carotid body and contribute to the control of breathing in male Sprague-Dawley rats. Peptides 2018, 102, 68–74. [Google Scholar] [CrossRef] [PubMed]
  79. Birder, L.A.; Wolf-Johnston, A.S.; Jackson, E.K.; Wein, A.J.; Dmochowski, R. Aging increases the expression of vasopressin receptors in both the kidney and urinary bladder. Neurourol. Urodyn. 2019, 38, 393–397. [Google Scholar] [CrossRef]
  80. Robertson, G.L. Differential diagnosis of familial diabetes insipidus. Handb. Clin. Neurol. 2021, 181, 239–248. [Google Scholar] [CrossRef]
  81. Dekan, Z.; Kremsmayr, T.; Keov, P.; Godin, M.; Teakle, N.; Dürrauer, L.; Xiang, H.; Gharib, D.; Bergmayr, C.; Hellinger, R.; et al. Nature-inspired dimerization as a strategy to modulate neuropeptide pharmacology exemplified with vasopressin and oxytocin. Chem. Sci. 2021, 12, 4057–4062. [Google Scholar] [CrossRef]
  82. Devost, D.; Zingg, H.H. Homo- and hetero-dimeric complex formations of the human oxytocin receptor. J. Neuroendocrinol. 2004, 16, 372–377. [Google Scholar] [CrossRef]
  83. Kim, S.H.; Bennett, P.R.; Terzidou, V. Advances in the role of oxytocin receptors in human parturition. Mol. Cell Endocrinol. 2017, 449, 56–63. [Google Scholar] [CrossRef] [PubMed]
  84. Murat, B.; Devost, D.; Andrés, M.; Mion, J.; Boulay, V.; Corbani, M.; Zingg, H.H.; Guillon, G. V1b and CRHR1 receptor heterodimerization mediates synergistic biological actions of vasopressin and CRH. Mol. Endocrinol. 2012, 26, 502–520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Terrillon, S.; Durroux, T.; Mouillac, B.; Breit, A.; Ayoub, M.A.; Taulan, M.; Jockers, R.; Barberis, C.; Bouvier, M. Oxytocin and vasopressin V1a and V2 receptors form constitutive homo- and heterodimers during biosynthesis. Mol. Endocrinol. 2003, 17, 677–6791. [Google Scholar] [CrossRef] [Green Version]
  86. Cosenza, G.; Iannaccone, M.; Pico, B.A.; Gallo, D.; Capparelli, R.; Pauciullo, A. Molecular characterisation, genetic variability and detection of a functional polymorphism influencing the promoter activity of OXT gene in goat and sheep. J. Dairy Res. 2017, 84, 165–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Pauciullo, A.; Ogah, D.M.; Iannaccone, M.; Erhardt, G.; Di Stasio, L.; Cosenza, G. Genetic characterization of the oxytocin-neurophysin I gene (OXT) and its regulatory regions analysis in domestic Old and New World camelids. PLoS ONE 2018, 13, e0195407. [Google Scholar] [CrossRef] [PubMed]
  88. Sausville, E.; Carney, D.; Battey, J. The human vasopressin gene is linked to the oxytocin gene and is selectively expressed in a cultured lung cancer cell line. J. Biol. Chem. 1985, 260, 10236–10241. [Google Scholar] [CrossRef]
  89. Pow, D.V.; Morris, J.F. Dendrites of hypothalamic magnocellular neurons release neurohypophysial peptides by exocytosis. Neuroscience 1989, 32, 435–439. [Google Scholar] [CrossRef]
  90. Makani, V.; Sultana, R.; Sie, K.S.; Orjiako, D.; Tatangelo, M.; Dowling, A.; Cai, J.; Pierce, W.; Butterfield, D.A.; Hill, J.; et al. Annexin A1 complex mediates oxytocin vesicle transport. J. Neuroendocrinol. 2013, 25, 1241–1254. [Google Scholar] [CrossRef] [Green Version]
  91. Jirikowski, G.F. Diversity of central oxytocinergic projections. Cell Tissue Res. 2019, 375, 41–48. [Google Scholar] [CrossRef]
  92. Jahn, R.; Südhof, T.C. Synaptic vesicles and exocytosis. Annu. Rev. Neurosci. 1994, 17, 219–246. [Google Scholar] [CrossRef]
  93. Llinás, R.; Sugimori, M.; Hillman, D.E.; Cherksey, B. Distribution and functional significance of the Ptype, voltage-dependent Ca2+ channels in the mammalian central nervous system. Trends Neurosci. 1992, 15, 351–355. [Google Scholar] [CrossRef]
  94. Stojilkovic, S.S. Ca2+-regulated exocytosis and SNARE function. Trends Endocrinol. Metab. 2005, 16, 81–83. [Google Scholar] [CrossRef] [PubMed]
  95. Tobin, V.; Schwab, Y.; Lelos, N.; Onaka, T.; Pittman, Q.J.; Ludwig, M. Expression of exocytosis proteins in rat supraoptic nucleus neurones. J. Neuroendocrinol. 2012, 24, 629–641. [Google Scholar] [CrossRef] [PubMed]
  96. Zhou, Q.; Zhou, P.; Wang, A.L.; Wu, D.; Zhao, M.; Südhof, T.C.; Brunger, A.T. The primed SNARE-complexin-synaptotagmin complex for neuronal exocytosis. Nature 2017, 548, 420–425. [Google Scholar] [CrossRef] [Green Version]
  97. Fisher, T.E.; Bourque, C.W. Calcium-channel subtypes in the somata and axon terminals of magnocellular neurosecretory cells. Trends Neurosci. 1996, 19, 440–444. [Google Scholar] [CrossRef]
  98. Foehring, R.C.; Armstrong, W.E. Pharmacological dissection of high-voltage-activated Ca2+ current types in acutely dissociated rat supraoptic magnocellular neurons. J. Neurophysiol. 1996, 76, 977–983. [Google Scholar] [CrossRef] [PubMed]
  99. Ludwig, M.; Landgraf, R. Does the release of vasopressin within the supraoptic nucleus of the rat brain depend upon changes in osmolality and Ca2+/K+? Brain Res. 1992, 576, 231–234. [Google Scholar] [CrossRef]
  100. Brown, C.H.; Ludwig, M.; Tasker, J.G.; Stern, J.E. Somato-dendritic vasopressin and oxytocin secretion in endocrine and autonomic regulation. J. Neuroendocrinol. 2020, 32, e12856. [Google Scholar] [CrossRef]
  101. Ginsberg, S.D.; Hof, P.R.; Young, W.G.; Morrison, J.H. Noradrenergic innervation of vasopressin- and oxytocin-containing neurons in the hypothalamic paraventricular nucleus of the macaque monkey: Quantitative analysis using double-label immunohistochemistry and confocal laser microscopy. J. Comp. Neurol. 1994, 341, 476–491. [Google Scholar] [CrossRef]
  102. Theodosis, D.T. Oxytocin-secreting neurons: A physiological model of morphological neuronal and glial plasticity in the adult hypothalamus. Front. Neuroendocrinol. 2002, 23, 101–135. [Google Scholar] [CrossRef]
  103. Jankowski, M.; Wang, D.; Hajjar, F.; Mukaddam-Daher, S.; McCann, S.M.; Gutkowska, J. Oxytocin and its receptors are synthesized in the rat vasculature. Proc. Natl. Acad. Sci. USA 2000, 97, 6207–6211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Nicholson, H.D.; Pickering, B.T. Oxytocin, a male intragonadal hormone. Regul. Pept. 1993, 45, 253–256. [Google Scholar] [CrossRef]
  105. Wsol, A.; Kasarello, K.; Kuch, M.; Gala, K.; Cudnoch-Jedrzejewska, A. Increased Activity of the Intracardiac Oxytocinergic System in the Development of Postinfarction Heart Failure. Biomed Res. Int. 2016, 2016, 3652068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Thibonnier, M.; Conarty, D.M.; Preston, J.A.; Plesnicher, C.L.; Dweik, R.A.; Erzurum, S.C. Human vascular endothelial cells express oxytocin receptors. Endocrinology 1999, 140, 1301–1309. [Google Scholar] [CrossRef]
  107. Ebner, K.; Wotjak, C.T.; Landgraf, R.; Engelmann, M. A single social defeat experience selectively stimulates the release of oxytocin, but not vasopressin, within the septal brain area of male rats. Brain Res. 2000, 872, 87–92. [Google Scholar] [CrossRef]
  108. Ivell, R.; Russell, J.A. Oxytocin: Cellular and molecular approaches in medicine and research. Rev. Reprod. 1996, 1, 13–18. [Google Scholar] [CrossRef]
  109. Jurek, B.; Neumann, I.D. The Oxytocin Receptor: From Intracellular Signaling to Behavior. Physiol. Rev. 2018, 98, 1805–1908. [Google Scholar] [CrossRef]
  110. Leng, G.; Caquineau, C.; Sabatier, N. Regulation of oxytocin secretion. Vitam. Horm. 2005, 71, 27–58. [Google Scholar] [CrossRef] [PubMed]
  111. Nishioka, T.; Anselmo-Franci, J.A.; Li, P.; Callahan, M.F.; Morris, M. Stress increases oxytocin release within the hypothalamic paraventricular nucleus. Brain Res. 1998, 781, 57–61. [Google Scholar] [CrossRef]
  112. Tobin, V.A.; Leng, G.; Ludwig, M.; Douglasm, A.J. Increased sensitivity of monoamine release in the supraoptic nucleus in late pregnancy: Region- and stimulus-dependent responses. J. Neuroendocrinol. 2010, 22, 430–437. [Google Scholar] [CrossRef]
  113. Nomura, M.; McKenna, E.; Korach, K.S.; Pfaff, D.W.; Ogawa, S. Estrogen receptor-beta regulates transcript levels for oxytocin and arginine vasopressin in the hypothalamic paraventricular nucleus of male mice. Brain Res. Mol. Brain Res. 2002, 109, 84–94. [Google Scholar] [CrossRef]
  114. Patisaul, H.B.; Scordalakes, E.M.; Young, L.J.; Rissman, E.F. Oxytocin, but not oxytocin receptor, is regulated by oestrogen receptor beta in the female mouse hypothalamus. J. Neuroendocrinol. 2003, 15, 787–793. [Google Scholar] [CrossRef] [PubMed]
  115. Sharma, D.; Handa, R.J.; Uht, R.M. The ERβ ligand 5α-androstane, 3β,17β-diol (3β-diol) regulates hypothalamic oxytocin (Oxt) gene expression. Endocrinology 2012, 153, 2353–2361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Arima, H.; Kondo, K.; Murase, T.; Yokoi, H.; Iwasaki, Y.; Saito, H.; Oiso, Y. Regulation of vasopressin synthesis and release by area postrema in rats. Endocrinology 1998, 139, 1481–1486. [Google Scholar] [CrossRef] [PubMed]
  117. Aspé-Sánchez, M.; Moreno, M.; Rivera, M.I.; Rossi, A.; Ewer, J. Oxytocin and vasopressin receptor gene polymorphisms: Role in social and psychiatric traits. Front. Neurosci. 2016, 9, 510. [Google Scholar] [CrossRef] [Green Version]
  118. Francis, S.M.; Sagar, A.; Levin-Decanini, T.; Liu, W.; Carter, C.S.; Jacob, S. Oxytocin and vasopressin systems in genetic syndromes and neurodevelopmental disorders. Brain Res. 2014, 1580, 199–218. [Google Scholar] [CrossRef] [Green Version]
  119. Caldwell, H.K.; Aulino, E.A.; Rodriguez, K.M.; Witchey, S.K.; Yaw, A.M. Social Context, Stress, Neuropsychiatric Disorders, and the Vasopressin 1b Receptor. Front. Neurosci. 2017, 11, 567. [Google Scholar] [CrossRef]
  120. Daikoku, R.; Kunitake, T.; Kato, K.; Tanoue, A.; Tsujimoto, G.; Kannan, H. Body water balance and body temperature in vasopressin V1b receptor knockout mice. Auton. Neurosci. 2007, 136, 58–62. [Google Scholar] [CrossRef]
  121. Lozić, M.; Šarenac, O.; Murphy, D.; Japundžić-Žigon, N. Vasopressin, Central Autonomic Control and Blood Pressure Regulation. Curr. Hypertens. Rep. 2018, 20, 11. [Google Scholar] [CrossRef]
  122. Wu, N.; Shang, S.; Su, Y. The arginine vasopressin V1b receptor gene and prosociality: Mediation role of emotional empathy. Psych J. 2015, 4, 160–165. [Google Scholar] [CrossRef]
  123. Antunes-Rodrigues, J.; Ruginsk, S.G.; Mecawi, A.S.; Margatho, L.O.; Cruz, J.C.; Vilhena-Franco, T.; Reis, W.L.; Ventura, R.R.; Reis, L.C.; Vivas, L.M.; et al. Mapping and signaling of neural pathways involved in the regulation of hydromineral homeostasis. Braz. J. Med. Biol. Res. 2013, 46, 327–338. [Google Scholar] [CrossRef] [PubMed]
  124. Song, Z.; Levin, B.E.; Stevens, W.; Sladek, C.D. Supraoptic oxytocin and vasopressin neurons function as glucose and metabolic sensors. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2014, 306, R447–R456. [Google Scholar] [CrossRef] [Green Version]
  125. Szczepańska-Sadowska, E.; Simon-Oppermann, C.; Gray, D.; Simon, E. Control of central release of vasopressin. J. Physiol. (Paris) 1984, 79, 432–439. [Google Scholar]
  126. Szczepańska-Sadowska, E.; Simon-Oppermann, C.; Gray, D.A.; Simon, E. Plasma and cerebrospinal fluid vasopressin and osmolality in relation to thirst. Pflugers. Arch. 1984, 400, 294–299. [Google Scholar] [CrossRef]
  127. Sharif-Naeini, R.; Ciura, S.; Zhang, Z.; Bourque, C.W. Contribution of TRPV channels to osmosensory transduction, thirst, and vasopressin release. Kidney Int. 2008, 73, 811–815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Tasker, J.G.; Prager-Khoutorsky, M.; Teruyama, R.; Lemos, J.R.; Amstrong, W.E. Advances in the neurophysiology of magnocellular neuroendocrine cells. J. Neuroendocrinol. 2020, 32, e12826. [Google Scholar] [CrossRef]
  129. Bansal, V.; Fisher, T.E. Osmotic activation of a Ca(2+)-dependent phospholipase C pathway that regulates N TRPV1-mediated currents in rat supraoptic neurons. Physiol. Rep. 2017, 5, e13259. [Google Scholar] [CrossRef] [PubMed]
  130. Japundžić-Žigon, N.; Lozić, M.; Šarenac, O.; Murphy, D. Vasopressin & Oxytocin in Control of the Cardiovascular System: An Updated Review. Curr. Neuropharmacol. 2020, 18, 14–33. [Google Scholar] [CrossRef]
  131. Leng, G.; Brown, C.H.; Russell, J.A. Physiological pathways regulating the activity of magnocellular neurosecretory cells. Prog. Neurobiol. 1999, 57, 625–655. [Google Scholar] [CrossRef]
  132. Johnson, A.K. The periventricular anteroventral third ventricle (AV3V): Its relationship with the subfornical organ and neural systems involved in maintaining body fluid homeostasis. Brain Res. Bull. 1985, 15, 595–601. [Google Scholar] [CrossRef]
  133. Antunes-Rodrigues, J.; de Castro, M.; Elias, L.L.; Valença, M.M.; McCann, S.M. Neuroendocrine control of body fluid metabolism. Physiol. Rev. 2004, 84, 169–208. [Google Scholar] [CrossRef] [Green Version]
  134. Lechner, S.G.; Markworth, S.; Poole, K.; Smith, E.S.; Lapatsina, L.; Frahm, S.; May, M.; Pischke, S.; Suzuki, M.; Ibañez-Tallon, I.; et al. The molecular and cellular identity of peripheral osmoreceptors. Neuron 2011, 69, 332–344. [Google Scholar] [CrossRef] [Green Version]
  135. Norsk, P. Influence of low- and high-pressure baroreflexes on vasopressin release in humans. Acta Endocrinol. (Copenh.) 1989, 121 (Suppl. 1), 3–27. [Google Scholar] [CrossRef] [PubMed]
  136. Schrier, R.W.; Berl, T.; Anderson, R.J. Osmotic and nonosmotic control of vasopressin release. Am. J. Physiol. 1979, 236, F321–F332. [Google Scholar] [CrossRef] [PubMed]
  137. Zera, T.; Moraes, D.J.A.; da Silva, M.P.; Fisher, J.P.; Paton, J.F.R. The Logic of Carotid Body Connectivity to the Brain. Physiology (Bethesda) 2019, 34, 264–282. [Google Scholar] [CrossRef]
  138. Ross, C.A.; Ruggiero, D.A.; Park, D.H.; Joh, T.H.; Sved, A.F.; Fernandez-Pardal, J.; Saavedra, J.M.; Reis, D.J. 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. 1984, 4, 474–494. [Google Scholar] [CrossRef]
  139. Greenwood, M.; Bordieri, L.; Greenwood, M.; Rosso Melo, M.; Colombari, D.S.; Colombari, E.; Paton, J.F.; Murphy, D. Transcription factor CREB3L1 regulates vasopressin gene expression in the rat hypothalamus. J. Neurosci. 2014, 34, 3810–3820. [Google Scholar] [CrossRef] [Green Version]
  140. Dawson, C.A.; Jhamandas, J.H.; Krukoff, T.L. Activation by systemic angiotensin II of neurochemically identified neurons in rat hypothalamic paraventricular nucleus. J. Neuroendocrinol. 1998, 10, 453–459. [Google Scholar] [CrossRef] [PubMed]
  141. Sladek, C.D.; Kapoor, J.R. Neurotransmitter/neuropeptide interactions in the regulation of neurohypophyseal hormone release. Exp. Neurol. 2001, 171, 200–209. [Google Scholar] [CrossRef]
  142. Coletti, R.; de Lima, J.B.M.; Vechiato, F.M.V.; de Oliveira, F.L.; Debarba, L.K.; Almeida-Pereira, G.; Elias, L.L.K.; Antunes-Rodrigues, J. Nitric oxide acutely modulates hypothalamic and neurohypophyseal carbon monoxide and hydrogen sulphide production to control vasopressin, oxytocin and atrial natriuretic peptide release in rats. J. Neuroendocrinol. 2019, 31, e12686. [Google Scholar] [CrossRef]
  143. Keeler, R.; Feuchuk, D.; Wilson, N. Atrial peptides and the renal response to hypervolemia in nephrotic rats. Can. J. Physiol. Pharmacol. 1987, 65, 2071–2075. [Google Scholar] [CrossRef] [PubMed]
  144. Reis, W.L.; Giusti-Paiva, A.; Ventura, R.R.; Margatho, L.O.; Gomes, D.A.; Elias, L.L.; Antunes-Rodrigues, J. Central nitric oxide blocks vasopressin, oxytocin and atrial natriuretic peptide release and antidiuretic and natriuretic responses induced by central angiotensin II in conscious rats. Exp. Physiol. 2007, 92, 903–911. [Google Scholar] [CrossRef] [PubMed]
  145. Haque, M.; Wilson, R.; Sharma, K.; Mills, N.J.; Teruyama, R. Localisation of 11β-Hydroxysteroid Dehydrogenase Type 2 in Mineralocorticoid Receptor Expressing Magnocellular Neurosecretory Neurones of the Rat Supraoptic and Paraventricular Nuclei. J. Neuroendocrinol. 2015, 27, 835–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Pietranera, L.; Saravia, F.; Roig, P.; Lima, A.; De Nicola, A.F. Mineralocorticoid treatment upregulates the hypothalamic vasopressinergic system of spontaneously hypertensive rats. Neuroendocrinology 2004, 80, 100–110. [Google Scholar] [CrossRef] [PubMed]
  147. Sladek, C.D.; Somponpun, S.J. Estrogen receptors: Their roles in regulation of vasopressin release for maintenance of fluid and electrolyte homeostasis. Front. Neuroendocrinol. 2008, 29, 114–127. [Google Scholar] [CrossRef] [Green Version]
  148. Huang, S.; Li, G.; Pan, Y.; Song, M.; Zhao, J.; Wan, X.; Krebs, C.J.; Wang, Z.; Han, W.; Zhang, Z. Density-induced social stress alters oxytocin and vasopressin activities in the brain of a small rodent species. Integr. Zool. 2021, 16, 149–159. [Google Scholar] [CrossRef] [PubMed]
  149. Lightman, S.L. The neuroendocrinology of stress: A never ending story. J. Neuroendocrinol. 2008, 20, 880–884. [Google Scholar] [CrossRef]
  150. Liu, J.P.; Clarke, I.J.; Funder, J.W.; Engler, D. Studies of the secretion of corticotropin-releasing factor and arginine vasopressin into the hypophysial-portal circulation of the conscious sheep. II. The central noradrenergic and neuropeptide Y pathways cause immediate and prolonged hypothalamic-pituitary-adrenal activation. Potential involvement in the pseudo-Cushing’s syndrome of endogenous depression and anorexia nervosa. J. Clin. Investig. 1994, 93, 1439–1450. [Google Scholar] [CrossRef] [Green Version]
  151. Nair, B.B.; Khant Aung, Z.; Porteous, R.; Prescott, M.; Glendining, K.A.; Jenkins, D.E.; Augustine, R.A.; Silva, M.S.B.; Yip, S.H.; Bouwer, G.T.; et al. Impact of chronic variable stress on neuroendocrine hypothalamus and pituitary in male and female C57BL/6J mice. J. Neuroendocrinol. 2021, 33, e12972. [Google Scholar] [CrossRef] [PubMed]
  152. Veenema, A.H.; Blume, A.; Niederle, D.; Buwalda, B.; Neumann, I.D. Effects of early life stress on adult male aggression and hypothalamic vasopressin and serotonin. Eur. J. Neurosci. 2006, 24, 1711–1720. [Google Scholar] [CrossRef] [Green Version]
  153. Wigger, A.; Sánchez, M.M.; Mathys, K.C.; Ebner, K.; Frank, E.; Liu, D.; Kresse, A.; Neumann, I.D.; Holsboer, F.; Plotsky, P.M.; et al. Alterations in central neuropeptide expression, release, and receptor binding in rats bred for high anxiety: Critical role of vasopressin. Neuropsychopharmacology 2004, 29, 1–14. [Google Scholar] [CrossRef]
  154. Szczepańska-Sadowska, E. Plasma ADH increase and osmotic thirst suppression elicited by preoptic heating in the dog. Am. J. Physiol. 1974, 226, 155–161. [Google Scholar] [CrossRef]
  155. Sharif-Naeini, R.; Ciura, S.; Bourque, C.W. TRPV1 gene required for thermosensory transduction and anticipatory secretion from vasopressin neurons during hyperthermia. Neuron 2008, 58, 179–185. [Google Scholar] [CrossRef] [Green Version]
  156. Gaida, W.; Lang, R.E.; Kraft, K.; Unger, T.; Ganten, D. Altered neuropeptide concentrations in spontaneously hypertensive rats: Cause or consequence? Clin. Sci. (Lond.) 1985, 68, 35–43. [Google Scholar] [CrossRef] [PubMed]
  157. Higa-Taniguchi, K.T.; Felix, J.V.; Michelini, L.C. Brainstem oxytocinergic modulation of heart rate control in rats: Effects of hypertension and exercise training. Exp. Physiol. 2009, 94, 1103–1113. [Google Scholar] [CrossRef]
  158. Martins, A.S.; Crescenzi, A.; Stern, J.E.; Bordin, S.; Michelini, L.C. Hypertension and exercise training differentially affect oxytocin and oxytocin receptor expression in the brain. Hypertension 2005, 46, 1004–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. van Tol, H.H.; van den Buuse, M.; de Jong, W.; Burbach, J.P. Vasopressin and oxytocin gene expression in the supraoptic and paraventricular nucleus of the spontaneously hypertensive rat (SHR) during development of hypertension. Brain Res. 1988, 464, 303–311. [Google Scholar] [CrossRef] [Green Version]
  160. Broderick, T.L.; Wang, Y.; Gutkowska, J.; Wang, D.; Jankowski, M. Downregulation of oxytocin receptors in right ventricle of rats with monocrotaline-induced pulmonary hypertension. Acta Physiol. (Oxf.) 2010, 200, 147–158. [Google Scholar] [CrossRef]
  161. Indrambarya, T.; Boyd, J.H.; Wang, Y.; McConechy, M.; Walley, K.R. Low-dose vasopressin infusion results in increased mortality and cardiac dysfunction following ischemia-reperfusion injury in mice. Crit. Care 2009, 13, R98. [Google Scholar] [CrossRef] [Green Version]
  162. Yu, R.; Wang, F.; Yin, J.; Shi, Y.; Wang, Y.; Wen, S.; Hu, H.; Yan, S. Expression of oxytocin receptor in the rat superior cervical ganglion after myocardial infarction. Int. J. Clin. Exp. Pathol. 2018, 11, 739–747. [Google Scholar]
  163. Althammer, F.; Eliava, M.; Grinevich, V. Central and peripheral release of oxytocin: Relevance of neuroendocrine and neurotransmitter actions for physiology and behavior. Handb. Clin. Neurol. 2021, 180, 25–44. [Google Scholar] [CrossRef]
  164. Boll, S.; Almeida de Minas, A.C.; Raftogianni, A.; Herpertz, S.C.; Grinevich, V. Oxytocin and Pain Perception: From Animal Models to Human Research. Neuroscience 2018, 387, 149–161. [Google Scholar] [CrossRef]
  165. Nersesyan, Y.; Demirkhanyan, L.; Cabezas-Bratesco, D.; Oakes, V.; Kusudam, R.; Dawsonm, T.; Sun, X.; Cao, C.; Cohen, A.M.; Chelluboina, B.; et al. Oxytocin Modulates Nociception as an Agonist of Pain-Sensing TRPV1. Cell Rep. 2017, 21, 1681–1691. [Google Scholar] [CrossRef] [Green Version]
  166. Szczepanska-Sadowska, E.; Cudnoch-Jedrzejewska, A.; Sadowski, B. Differential role of specific cardiovascular neuropeptides in pain regulation: Relevance to cardiovascular diseases. Neuropeptides 2020, 81, 102046. [Google Scholar] [CrossRef] [PubMed]
  167. Nishimura, H.; Kawasaki, M.; Suzuki, H.; Matsuura, T.; Motojima, Y.; Ohnishi, H.; Yamanaka, Y.; Yoshimura, M.; Maruyama, T.; Saito, R.; et al. Neuropathic Pain Up-Regulates Hypothalamo-Neurohypophysial and Hypothalamo-Spinal Oxytocinergic Pathways in Oxytocin-Monomeric Red Fluorescent Protein 1 Transgenic Rat. Neuroscience 2019, 406, 50–61. [Google Scholar] [CrossRef] [PubMed]
  168. Pan, Y.J.; Wang, D.X.; Yang, J.; He, X.L.; Xiao, N.M.; Ma, R.Q.; Wang, C.H.; Lin, B.C. Oxytocin in hypothalamic supraoptic nucleus is transferred to the caudate nucleus to influence pain modulation. Neuropeptides 2016, 58, 61–65. [Google Scholar] [CrossRef] [PubMed]
  169. Jezova, D.; Skultetyova, I.; Tokarev, D.I.; Bakos, P.; Vigas, M. Vasopressin and oxytocin in stress. Ann. N. Y. Acad. Sci. 1995, 771, 192–203. [Google Scholar] [CrossRef] [PubMed]
  170. Skopek, P.; Hynie, S.; Chottova-Dvorakova, M.; Sida, P.; Slavikova, J.; Mistrova, E.; Klenerova, V. Effects of acute stressors on the expression of oxytocin receptor mRNA in hearts of rats with different activity of HPA axis. Neuro Endocrinol. Lett. 2012, 33, 124–132. [Google Scholar]
  171. Wsol, A.; Szczepanska-Sadowska, E.; Kowalewski, S.; Puchalska, L.; Cudnoch-Jedrzejewska, A. Oxytocin differently regulates pressor responses to stress in WKY and SHR rats: The role of central oxytocin and V1a receptors. Stress 2014, 17, 117–125. [Google Scholar] [CrossRef]
  172. Wsol, A.; Wojno, O.; Puchalska, L.; Wrzesien, R.; Szczepanska-Sadowska, E.; Cudnoch-Jedrzejewska, A. Impaired hypotensive effects of centrally acting oxytocin in SHR and WKY rats exposed to chronic mild stress. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2020, 318, R160–R172. [Google Scholar] [CrossRef]
  173. Sivukhina, E.V.; Morozov, I.E.; Dolzhikov, A.A.; Jirikowski, G.F.; Grinevich, V. Comparison of vasopressin and oxytocin expressions in the hypothalamo-neurohypophysial system of patients with chronic heart failure. Horm. Metab. Res. 2010, 42, 56–60. [Google Scholar] [CrossRef]
  174. Stone, G.; Choi, A.; Meritxell, O.; Gorham, J.; Heydarpour, M.; Seidman, C.E.; Seidman, J.G.; Aranki, S.F.; Body, S.C.; Carey, V.J.; et al. Sex differences in gene expression in response to ischemia in the human left ventricular myocardium. Hum. Mol. Genet. 2019, 28, 1682–1693. [Google Scholar] [CrossRef]
  175. Cowley, A.W., Jr.; Cushman, W.C.; Quillen, E.W., Jr.; Skelton, M.M.; Langford, H.G. Vasopressin elevation in essential hypertension and increased responsiveness to sodium intake. Hypertension 1981, 3, I93–I100. [Google Scholar] [CrossRef] [Green Version]
  176. Iovino, M.; Guastamacchia, E.; Giagulli, V.A.; Fiore, G.; Licchelli, B.; Iovino, E.; Triggiani, V. Role of central and peripheral chemoreceptors in vasopressin secretion control. Endocr. Metab. Immune Disord. Drug Targets 2013, 13, 250–255. [Google Scholar] [CrossRef] [PubMed]
  177. Muders, F.; Riegger, G.A.; Bahner, U.; Palkovits, M. The central vasopressinergic system in experimental left ventricular hypertrophy and dysfunction. Prog. Brain Res. 2002, 139, 275–279. [Google Scholar] [CrossRef] [PubMed]
  178. Feuerstein, G.; Zerbe, R.L.; Faden, A.I. Central cardiovascular effects of vasotocin, oxytocin and vasopressin in conscious rats. J. Pharmacol. Exp. Ther. 1984, 228, 348–353. [Google Scholar] [PubMed]
  179. Moriguchi, A.; Ferrario, C.M.; Brosnihan, K.B.; Ganten, D.; Morris, M. Differential regulation of central vasopressin in transgenic rats harboring the mouse Ren-2 gene. Am. J. Physiol. 1994, 267 Pt 2, R786–R791. [Google Scholar] [CrossRef] [PubMed]
  180. Teruyama, R.; Sakuraba, M.; Wilson, L.L.; Wandrey, N.E.; Armstrong, W.E. Epithelial Na⁺ sodium channels in magnocellular cells of the rat supraoptic and paraventricular nuclei. Am. J. Physiol. Endocrinol. Metab. 2012, 302, E273–E285. [Google Scholar] [CrossRef] [Green Version]
  181. Amin, M.S.; Wang, H.W.; Reza, E.; Whitman, S.C.; Tuana, B.S.; Leenen, F.H. Distribution of epithelial sodium channels and mineralocorticoid receptors in cardiovascular regulatory centers in rat brain. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2005, 289, R1787–R1797. [Google Scholar] [CrossRef] [PubMed]
  182. Muders, F.; Elsner, D.; Schunkert, H.; Riegger, G.A.; Palkovits, M. Central vasopressin is modulated by chronic blockade of the renin-angiotensin system in experimental left ventricular hypertrophy. Am. J. Hypertens. 1999, 12, 311–314. [Google Scholar] [CrossRef] [Green Version]
  183. Fujimori, K.; Endo, C.; Kin, S.; Funata, Y.; Araki, T.; Sato, A.; Murata, Y. Endocrinologic and biophysical responses to prolonged (24-hour) hypoxemia in fetal goats. Am. J. Obstet. Gynecol. 1994, 171, 470–477. [Google Scholar] [CrossRef]
  184. Raff, H.; Shinsako, J.; Keil, L.C.; Dallman, M.F. Vasopressin, ACTH, and corticosteroids during hypercapnia and graded hypoxia in dogs. Am. J. Physiol. 1983, 244, E453–E458. [Google Scholar] [CrossRef] [PubMed]
  185. King, T.L.; Kline, D.D.; Ruyle, B.C.; Heesch, C.M.; Hasser, E.M. Acute systemic hypoxia activates hypothalamic paraventricular nucleus-projecting catecholaminergic neurons in the caudal ventrolateral medulla. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2013, 305, R1112–R1123. [Google Scholar] [CrossRef] [Green Version]
  186. Kubo, T.; Yanagihara, Y.; Yamaguchi, H.; Fukumori, R. Excitatory amino acid receptors in the paraventricular hypothalamic nucleus mediate pressor response induced by carotid body chemoreceptor stimulation in rats. Clin. Exp. Hypertens. 1997, 19, 1117–1134. [Google Scholar] [CrossRef]
  187. Aguilera, G. HPA axis responsiveness to stress: Implications for healthy aging. Exp. Gerontol. 2011, 46, 90–95. [Google Scholar] [CrossRef] [Green Version]
  188. Suzuki, H.; Kawasaki, M.; Ohnishi, H.; Otsubo, H.; Ohbuchi, T.; Katoh, A.; Hashimoto, H.; Yokoyama, T.; Fujihara, H.; Dayanithi, G.; et al. Exaggerated response of a vasopressin-enhanced green fluorescent protein transgene to nociceptive stimulation in the rat. J. Neurosci. 2009, 29, 13182–13189. [Google Scholar] [CrossRef]
  189. Szczepanska-Sadowska, E.; Cudnoch-Jedrzejewska, A.; Wsol, A. The role of oxytocin and vasopressin in the pathophysiology of heart failure in pregnancy and in fetal and neonatal life. Am. J. Physiol. Heart. Circ. Physiol. 2020, 318, H639–H651. [Google Scholar] [CrossRef]
  190. Yang, J.; Yang, Y.; Chen, J.M.; Xu, H.T.; Liu, W.Y.; Lin, B.C. Arginine vasopressin in periaqueductal gray, which relates to antinociception, comes from hypothalamic paraventricular nucleus in the rat. Neurosci. Lett. 2007, 412, 154–158. [Google Scholar] [CrossRef]
  191. Yang, J.; Yang, Y.; Xu, H.T.; Chen, J.M.; Liu, W.Y.; Lin, B.C. Arginine vasopressin induces periaqueductal gray release of enkephalin and endorphin relating to pain modulation in the rat. Regul. Pept. 2007, 142, 29–36. [Google Scholar] [CrossRef]
  192. Yang, J.; Yuan, H.F.; Liu, W.Y.; Zhang, X.X.; Feng, J.P.; Ni, N.; Yang, D.W.; Song, C.Y.; Xu, H.T.; Wang, G.; et al. Norepinephrine regulates arginine vasopressin secretion in hypothalamic paraventricular nucleus relating with pain modulation. Neuropeptides 2009, 43, 259–265. [Google Scholar] [CrossRef]
  193. McAlpine, H.M.; Cobbe, S.M. Neuroendocrine changes in acute myocardial infarction. Am. J. Med. 1988, 84, 61–66. [Google Scholar] [CrossRef] [PubMed]
  194. Goldsmith, S.R.; Francis, G.S.; Cowley, A.W., Jr.; Levine, T.B.; Cohn, J.N. Increased plasma arginine vasopressin levels in patients with congestive heart failure. J. Am. Coll. Cardiol. 1983, 1, 1385–1390. [Google Scholar] [CrossRef] [Green Version]
  195. Khan, S.Q.; Dhillon, O.S.; O’Brien, R.J.; Struck, J.; Quinn, P.A.; Morgenthaler, N.G.; Squire, I.B.; Davies, J.E.; Bergmannm, A.; Ng, L.L. C-terminal provasopressin (copeptin) as a novel and prognostic marker in acute myocardial infarction: Leicester Acute Myocardial Infarction Peptide (LAMP) study. Circulation 2007, 115, 2103–2110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Reinstadler, S.J.; Klug, G.; Feistritzer, H.J.; Mayr, A.; Harrasser, B.; Mair, J.; Bader, K.; Streil, K.; Hammerer-Lercher, A.; Esterhammer, R.; et al. Association of copeptin with myocardial infarct size and myocardial function after ST segment elevation myocardial infarction. Heart 2013, 99, 1525–1529. [Google Scholar] [CrossRef] [PubMed]
  197. Szatalowicz, V.L.; Arnold, P.E.; Chaimovitz, C.; Bichet, D.; Berl, T.; Schrier, R.W. Radioimmunoassay of plasma arginine vasopressin in hyponatremic patients with congestive heart failure. N. Engl. J. Med. 1981, 305, 263–266. [Google Scholar] [CrossRef]
  198. Gimpl, G.; Reitz, J.; Brauer, S.; Trossen, C. Oxytocin receptors: Ligand binding, signalling and cholesterol dependence. Prog. Brain Res. 2008, 170, 193–204. [Google Scholar] [CrossRef] [PubMed]
  199. Gutkowska, J.; Jankowski, M. Oxytocin revisited: It is also a cardiovascular hormone. J. Am. Soc. Hypertens. 2008, 2, 318–325. [Google Scholar] [CrossRef]
  200. Jankowski, M.; Bissonauth, V.; Gaom, L.; Gangal, M.; Wang, D.; Danalache, B.; Wang, Y.; Stoyanova, E.; Cloutier, G.; Blaise, G.; et al. Anti-inflammatory effect of oxytocin in rat myocardial infarction. Basic. Res. Cardiol. 2010, 105, 205–218. [Google Scholar] [CrossRef]
  201. Sofroniew, M.V. Vasopressin and oxytocin in the mammalian brain and spinal cord. Trends Neurosci. 1983, 6, 467–472. [Google Scholar] [CrossRef]
  202. Young, L.J.; Wang, Z.; Donaldson, R.; Rissman, E.F. Estrogen receptor alpha is essential for induction of oxytocin receptor by estrogen. Neuroreport 1998, 9, 933–936. [Google Scholar] [CrossRef]
  203. Petersson, M.; Uvnäs-Moberg, K. Effects of an acute stressor on blood pressure and heart rate in rats pretreated with intracerebroventricular oxytocin injections. Psychoneuroendocrinology 2007, 32, 959–965. [Google Scholar] [CrossRef]
  204. Petersson, M.; Alster, P.; Lundeberg, T.; Uvnäs-Moberg, K. Oxytocin causes a long-term decrease of blood pressure in female and male rats. Physiol. Behav. 1996, 60, 1311–1315. [Google Scholar] [CrossRef]
  205. Petersson, M.; Lundeberg, T.; Uvnäs-Moberg, K. Oxytocin enhances the effects of clonidine on blood pressure and locomotor activity in rats. J. Auton. Nerv. Syst. 1999, 78, 49–56. [Google Scholar] [CrossRef]
  206. Miller, M.E.; Davidge, S.T.; Mitchell, B.F. Oxytocin does not directly affect vascular tone in vessels from nonpregnant and pregnant rats. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, H1223–H1228. [Google Scholar] [CrossRef] [Green Version]
  207. Paquin, J.; Danalache, B.A.; Jankowski, M.; McCann, S.M.; Gutkowska, J. Oxytocin induces differentiation of P19 embryonic stem cells to cardiomyocytes. Proc. Natl. Acad. Sci. USA 2002, 99, 9550–9555. [Google Scholar] [CrossRef] [Green Version]
  208. Gutkowska, J.; Aliou, Y.; Lavoie, J.L.; Gaab, K.; Jankowski, M.; Broderick, T.L. Oxytocin decreases diurnal and nocturnal arterial blood pressure in the conscious unrestrained spontaneously hypertensive rat. Pathophysiology 2016, 23, 111–121. [Google Scholar] [CrossRef] [PubMed]
  209. Petersson, M.; Lundeberg, T.; Uvnäs-Moberg, K. Oxytocin decreases blood pressure in male but not in female spontaneously hypertensive rats. J. Auton. Nerv. Syst. 1997, 66, 15–18. [Google Scholar] [CrossRef]
  210. Holst, S.; Uvnäs-Moberg, K.; Petersson, M. Postnatal oxytocin treatment and postnatal stroking of rats reduce blood pressure in adulthood. Auton. Neurosci. 2002, 99, 85–90. [Google Scholar] [CrossRef]
  211. Petersson, M.; Uvnäs-Moberg, K. Postnatal oxytocin treatment of spontaneously hypertensive male rats decreases blood pressure and body weight in adulthood. Neurosci. Lett. 2008, 440, 166–169. [Google Scholar] [CrossRef] [PubMed]
  212. Wsol, A.; Gondek, A.; Podobinska, M.; Chmielewski, M.; Sajdel-Sułkowska, E.; Cudnoch-Jędrzejewska, A. Increased oxytocinergic system activity in the cardiac muscle in spontaneously hypertensive SHR rats. Arch. Med. Sci. 2019, 2019, 85446. [Google Scholar] [CrossRef]
  213. Faghihi, M.; Alizadeh, A.M.; Khori, V.; Latifpour, M.; Khodayari, S. The role of nitric oxide, reactive oxygen species, and protein kinase C in oxytocin-induced cardioprotection in ischemic rat heart. Peptides 2012, 37, 314–319. [Google Scholar] [CrossRef] [PubMed]
  214. Jankowski, M.; Wang, D.; Danalache, B.; Gangal, M.; Gutkowska, J. Cardiac oxytocin receptor blockade stimulates adverse cardiac remodeling in ovariectomized spontaneously hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H265–H274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Jankowski, M.; Broderick, T.L.; Gutkowska, J. Oxytocin and cardioprotection in diabetes and obesity. BMC Endocr. Disord. 2016, 16, 34. [Google Scholar] [CrossRef] [Green Version]
  216. Jankowski, M.; Broderick, T.L.; Gutkowska, J. The Role of Oxytocin in Cardiovascular Protection. Front. Psychol. 2020, 11, 2139. [Google Scholar] [CrossRef]
  217. Ondrejcakova, M.; Barancik, M.; Bartekova, M.; Ravingerova, T.; Jezova, D. Prolonged oxytocin treatment in rats affects intracellular signaling and induces myocardial protection against infarction. Gen. Physiol. Biophys. 2012, 31, 261–270. [Google Scholar] [CrossRef] [Green Version]
  218. Polshekan, M.; Jamialahmadi, K.; Khori, V.; Alizadeh, A.M.; Saeidi, M.; Ghayour-Mobarhan, M.; Jand, Y.; Ghahremani, M.H.; Yazdani, Y. RISK pathway is involved in oxytocin postconditioning in isolated rat heart. Peptides 2016, 86, 55–62. [Google Scholar] [CrossRef]
  219. Polshekan, M.; Khori, V.; Alizadeh, A.M.; Ghayour-Mobarhan, M.; Saeidi, M.; Jand, Y.; Rajaei, M.; Farnoosh, G.; Jamialahmadi, K. The SAFE pathway is involved in the postconditioning mechanism of oxytocin in isolated rat heart. Peptides 2019, 111, 142–151. [Google Scholar] [CrossRef]
  220. Gonzalez-Reyes, A.; Menaouar, A.; Yip, D.; Danalache, B.; Plante, E.; Noiseux, N.; Gutkowska, J.; Jankowski, M. Molecular mechanisms underlying oxytocin-induced cardiomyocyte protection from simulated ischemia-reperfusion. Mol. Cell Endocrinol. 2015, 412, 170–181. [Google Scholar] [CrossRef]
  221. Wsol, A.; Cudnoch-Jedrzejewska, 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, 517–525. [Google Scholar] [CrossRef]
  222. Garrott, K.; Dyavanapalli, J.; Cauley, E.; Dwyer, M.K.; Kuzmiak-Glancy, S.; Wang, X.; Mendelowitz, D.; Kay, M.W. Chronic activation of hypothalamic oxytocin neurons improves cardiac function during left ventricular hypertrophy-induced heart failure. Cardiovasc. Res. 2017, 113, 1318–1328. [Google Scholar] [CrossRef] [Green Version]
  223. Gutkowska, J.; Broderick, T.L.; Bogdan, D.; Wang, D.; Lavoie, J.M.; Jankowski, M. Downregulation of oxytocin and natriuretic peptides in diabetes: Possible implications in cardiomyopathy. J. Physiol. 2009, 587 Pt 19, 4725–4736. [Google Scholar] [CrossRef] [PubMed]
  224. Plante, E.; Menaouar, A.; Danalache, B.A.; Yip, D.; Broderick, T.L.; Chiasson, J.L.; Jankowski, M.; Gutkowska, J. Oxytocin treatment prevents the cardiomyopathy observed in obese diabetic male db/db mice. Endocrinology 2015, 156, 1416–1428. [Google Scholar] [CrossRef] [PubMed]
  225. Jung, C.; Wernly, B.; Bjursell, M.; Wiseman, J.; Admyre, T.; Wikström, J.; Palmér, M.; Seeliger, F.; Lichtenauer, M.; Franz, M.; et al. Cardiac-Specific Overexpression of Oxytocin Receptor Leads to Cardiomyopathy in Mice. J. Card Fail. 2018, 24, 470–478. [Google Scholar] [CrossRef] [PubMed]
  226. Rosseland, L.A.; Hauge, T.H.; Grindheim, G.; Stubhaug, A.; Langesæter, E. Changes in blood pressure and cardiac output during cesarean delivery: The effects of oxytocin and carbetocin compared with placebo. Anesthesiology 2013, 119, 541–551. [Google Scholar] [CrossRef] [PubMed]
  227. Svanström, M.C.; Biber, B.; Hanes, M.; Johansson, G.; Näslund, U.; Bålfors, E.M. Signs of myocardial ischaemia after injection of oxytocin: A randomized double-blind comparison of oxytocin and methylergometrine during Caesarean section. Br. J. Anaesth. 2008, 100, 683–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Kemp, A.H.; Quintana, D.S.; Kuhnert, R.L.; Griffiths, K.; Hickie, I.B.; Guastella, A.J. Oxytocin increases heart rate variability in humans at rest: Implications for social approach-related motivation and capacity for social engagement. PLoS ONE 2012, 7, e44014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Cowley, A.W., Jr.; Szczepanska-Sadowska, E.; Stepniakowski, K.; Mattson, D. Chronic intravenous administration of V1 arginine vasopressin agonist results in sustained hypertension. Am. J. Physiol. 1994, 267 Pt 2, H751–H756. [Google Scholar] [CrossRef] [PubMed]
  230. Góźdź, A.; Szczepańska-Sadowska, E.; Szczepańska, K.; Maśliński, W.; Luszczyk, B. Vasopressin V1a, V1b and V2 receptors mRNA in the kidney and heart of the renin transgenic TGR(mRen2)27 and Sprague Dawley rats. J. Physiol. Pharmacol. 2002, 53, 349–357. [Google Scholar]
  231. Gruber, C.W.; Koehbach, J.; Muttenthaler, M. Exploring bioactive peptides from natural sources for oxytocin and vasopressin drug discovery. Future Med. Chem. 2012, 4, 1791–1798. [Google Scholar] [CrossRef] [Green Version]
  232. Guillon, G.; Trueba, M.; Joubert, D.; Grazzini, E.; Chouinard, L.; Côté, M.; Payet, M.D.; Manzoni, O.; Barberis, C.; Robert, M. Vasopressin stimulates steroid secretion in human adrenal glands: Comparison with angiotensin-II effect. Endocrinology 1995, 136, 1285–1295. [Google Scholar] [CrossRef]
  233. Szczepanska-Sadowska, E.; Stepniakowski, K.; Skelton, M.M.; Cowley, A., Jr. Prolonged stimulation of intrarenal V1 vasopressin receptors results in sustained hypertension. Am. J. Physiol. 1994, 267, R1217–R1225. [Google Scholar] [CrossRef] [PubMed]
  234. Kc, P.; Haxhiu, M.A.; Tolentino-Silva, F.P.; Wu, M.; Trouth, C.O.; Mack, S.O. Paraventricular vasopressin-containing neurons project to brain stem and spinal cord respiratory-related sites. Respir. Physiol. Neurobiol. 2002, 133, 75–88. [Google Scholar] [CrossRef]
  235. Komnenov, D.; Quaal, H.; Rossi, N.F. V(1a) and V(1b) 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, R213–R225. [Google Scholar] [CrossRef] [PubMed]
  236. Yang, Z.; Coote, J.H. The role of supraspinal vasopressin and glutamate neurones in an increase in renal sympathetic activity in response to mild haemorrhage in the rat. Exp. Physiol. 2006, 91, 791–797. [Google Scholar] [CrossRef]
  237. Lankhuizen, I.M.; van Veghel, R.; Saxena, P.R.; Schoemaker, R.G. [Arg8]-vasopressin-induced responses on coronary and mesenteric arteries of rats with myocardial infarction: The effects of V1a- and V2-receptor antagonists. J. Cardiovasc. Pharmacol. 2000, 36, 38–44. [Google Scholar] [CrossRef]
  238. Lankhuizen, I.M.; van Veghel, R.; Saxena, P.R.; Schoemaker, R.G. Vascular and renal effects of vasopressin and its antagonists in conscious rats with chronic myocardial infarction; evidence for receptor shift. Eur. J. Pharmacol. 2001, 423, 195–202. [Google Scholar] [CrossRef]
  239. Sellke, N.; Kuczmarski, A.; Lawandy, I.; Cole, V.L.; Ehsan, A.; Singh, A.K.; Liu, Y.; Sellke, F.W.; Feng, J. Enhanced coronary arteriolar contraction to vasopressin in patients with diabetes after cardiac surgery. J. Thorac. Cardiovasc. Surg. 2018, 156, 2098–2107. [Google Scholar] [CrossRef]
  240. Serradeil-Le Gal, C.; Villanova, G.; Boutin, M.; Maffrand, J.P.; Le Fur, G. Effects of SR 49059, a non-peptide antagonist of vasopressin V1a receptors, on vasopressin-induced coronary vasoconstriction in conscious rabbits. Fundam. Clin. Pharmacol. 1995, 9, 17–24. [Google Scholar] [CrossRef] [PubMed]
  241. Walker, B.R.; Haynes, J.J.; Wang, H.L.; Voelkel, N.F. Vasopressin-induced pulmonary vasodilation in rats. Am. J. Physiol. 1989, 257, H415–H422. [Google Scholar] [CrossRef]
  242. Izumi, Y.; Nakayama, Y.; Mori, T.; Miyazaki, H.; Inoue, H.; Kohda, Y.; Inoue, T.; Nonoguchi, H.; Tomita, K. Downregulation of vasopressin V2 receptor promoter activity via V1a receptor pathway. Am. J. Physiol. Renal. Physiol. 2007, 292, F1418–F1426. [Google Scholar] [CrossRef]
  243. Terada, Y.; Tomita, K.; Nonoguchi, H.; Yang, T.; Marumo, F. Different localization and regulation of two types of vasopressin receptor messenger RNA in microdissected rat nephron segments using reverse transcription polymerase chain reaction. J. Clin. Investig. 1993, 92, 2339–2345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Aoyagi, T.; Koshimizu, T.A.; Tanoue, A. Vasopressin regulation of blood pressure and volume: Findings from V1a receptor-deficient mice. Kidney Int. 2009, 76, 1035–1039. [Google Scholar] [CrossRef] [Green Version]
  245. Izumi, Y.; Hori, K.; Nakayama, Y.; Kimura, M.; Hasuike, Y.; Nanami, M.; Kohda, Y.; Otaki, Y.; Kuragano, T.; Obinata, M.; et al. Aldosterone requires vasopressin V1a receptors on intercalated cells to mediate acid-base homeostasis. J. Am. Soc. Nephrol. 2011, 22, 673–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Bucher, M.; Hobbhahn, J.; Taeger, K.; Kurtz, A. Cytokine-mediated downregulation of vasopressin V(1A) receptors during acute endotoxemia in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002, 282, R979–R984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Pena, A.; Murat, B.; Trueba, M.; Ventura, M.A.; Bertrand, G.; Cheng, L.L.; Stoev, S.; Szeto, H.H.; Wo, N.; Brossard, G.; et al. Pharmacological and physiological characterization of d[Leu4, Lys8]vasopressin, the first V1b-selective agonist for rat vasopressin/oxytocin receptors. Endocrinology 2007, 148, 4136–4146. [Google Scholar] [CrossRef] [Green Version]
  248. Saito, M.; Tahara, A.; Sugimoto, T.; Abe, K.; Furuichi, K. Evidence that atypical vasopressin V(2) receptor in inner medulla of kidney is V(1B) receptor. Eur. J. Pharmacol. 2000, 401, 289–296. [Google Scholar] [CrossRef]
  249. El-Werfali, W.; Toomasian, C.; Maliszewska-Scislo, M.; Li, C.; Rossi, N.F. Haemodynamic and renal sympathetic responses to V1b vasopressin receptor activation within the paraventricular nucleus. Exp. Physiol. 2015, 100, 553–565. [Google Scholar] [CrossRef]
  250. Aguilera, G.; Rabadan-Diehl, C. Regulation of vasopressin V1b receptors in the anterior pituitary gland of the rat. Exp. Physiol. 2000, 85, 19S–26S. [Google Scholar] [CrossRef]
  251. Oshikawa, S.; Tanoue, A.; Koshimizu, T.A.; Kitagawa, Y.; Tsujimoto, G. Vasopressin stimulates insulin release from islet cells through V1b receptors: A combined pharmacological/knockout approach. Mol. Pharmacol. 2004, 65, 623–629. [Google Scholar] [CrossRef] [Green Version]
  252. Hus-Citharel, A.; Bouby, N.; Corbani, M.; Mion, J.; Mendre, C.; Darusi, J.; Tomboly, C.; Trueba, M.; Serradeil-Le Gal, C.; Llorens-Cortes, C.; et al. Characterization of a functional V(1B) vasopressin receptor in the male rat kidney: Evidence for cross talk between V(1B) and V(2) receptor signaling pathways. Am. J. Physiol. Renal. Physiol. 2021, 321, F305–F321. [Google Scholar] [CrossRef]
  253. Szczepanska-Sadowska, E.; Brzezinski, M. Interaction between effects of insulin and vasopressin on renal excretion of water and sodium in rats. Horm. Metab. Res. 1982, 14, 175–179. [Google Scholar] [CrossRef]
  254. Ishikawa, S.E. Hyponatremia Associated with Heart Failure: Pathological Role of Vasopressin-Dependent Impaired Water Excretion. J. Clin. Med. 2015, 4, 933–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Qian, Q. Salt, water and nephron: Mechanisms of action and link to hypertension and chronic kidney disease. Nephrology (Carlton) 2018, 23, 44–49. [Google Scholar] [CrossRef] [Green Version]
  256. Robertson, G.L. The regulation of vasopressin function in health and disease. Recent Prog. Horm. Res. 1976, 33, 333–385. [Google Scholar] [CrossRef]
  257. Gonzalez, A.A.; Cifuentes-Araneda, F.; Ibaceta-Gonzalez, C.; Gonzalez-Vergara, A.; Zamora, L.; Henriquez, R.; Rosales, C.B.; Navar, L.G.; Prieto, M.C. Vasopressin/V2 receptor stimulates renin synthesis in the collecting duct. Am. J. Physiol. Renal. Physiol. 2016, 310, F284–F293. [Google Scholar] [CrossRef] [Green Version]
  258. Schrier, R.W. Vasopressin and aquaporin 2 in clinical disorders of water homeostasis. Semin Nephrol. 2008, 28, 289–296. [Google Scholar] [CrossRef] [Green Version]
  259. Stockand, J.D. Vasopressin regulation of renal sodium excretion. Kidney Int. 2010, 78, 849–856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Nielsen, J.; Kwon, T.H.; Praetorius, J.; Frøkiaer, J.; Knepper, M.A.; Nielsen, S. Aldosterone increases urine production and decreases apical AQP2 expression in rats with diabetes insipidus. Am. J. Physiol. Renal. Physiol. 2006, 290, F438–F449. [Google Scholar] [CrossRef] [Green Version]
  261. Brønd, L.; Müllertz, K.M.; Torp, M.; Nielsen, J.; Graebe, M.; Hadrup, N.; Nielsen, S.; Christensen, S.; Jonassen, T.E. Congestive heart failure in rats is associated with increased collecting duct vasopressin sensitivity and vasopressin type 2 receptor reexternalization. Am. J. Physiol. Renal. Physiol. 2013, 305, F1547–F1554. [Google Scholar] [CrossRef] [Green Version]
  262. Cowley, A.W., Jr. Control of the renal medullary circulation by vasopressin V1 and V2 receptors in the rat. Exp. Physiol. 2000, 85, 223S–231S. [Google Scholar] [CrossRef] [Green Version]
  263. Gassanov, N.; Jankowski, M.; Danalache, B.; Wang, D.; Grygorczyk, R.; Hoppe, U.C.; Gutkowska, J. Arginine vasopressin-mediated cardiac differentiation: Insights into the role of its receptors and nitric oxide signaling. J. Biol. Chem. 2007, 282, 11255–11265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Burrell, L.M.; Phillips, P.A.; Stephenson, J.M.; Risvanis, J.; Johnston, C.I. Vasopressin and a nonpeptide antidiuretic hormone receptor antagonist (OPC-31260). Blood Press. 1994, 3, 137–141. [Google Scholar] [CrossRef] [PubMed]
  265. Clair, M.J.; King, M.K.; Goldberg, A.T.; Hendrick, J.W.; Nisato, R.; Gay, D.M.; Morrison, A.E.; McElmurray, J.H., 3rd; Krombach, R.S.; Bond, B.R.; et al. Selective vasopressin, angiotensin II, or dual receptor blockade with developing congestive heart failure. J. Pharmacol. Exp. Ther. 2000, 293, 852–860. [Google Scholar] [PubMed]
  266. Milik, E.; Szczepanska-Sadowska, E.; Dobruch, J.; Cudnoch-Jedrzejewska, A.; Maslinski, W. Altered expression of V1a receptors mRNA in the brain and kidney after myocardial infarction and chronic stress. Neuropeptides 2014, 48, 257–266. [Google Scholar] [CrossRef] [PubMed]
  267. Faraco, G.; Wijasa, T.S.; Park, L.; Moore, J.; Anrather, J.; Iadecola, C. Water deprivation induces neurovascular and cognitive dysfunction through vasopressin-induced oxidative stress. J. Cereb. Blood Flow Metab. 2014, 34, 852–860. [Google Scholar] [CrossRef] [Green Version]
  268. Zhu, W.; Tilley, D.G.; Myers, V.D.; Tsai, E.J.; Feldman, A.M. Increased vasopressin 1A receptor expression in failing human hearts. J. Am. Coll. Cardiol. 2014, 63, 375–376. [Google Scholar] [CrossRef] [Green Version]
  269. Gheorghiade, M.; Konstam, M.A.; Burnett, J.C., Jr.; Grinfeld, L.; Maggioni, A.P.; Swedberg, K.; Udelson, J.E.; Zannad, F.; Cook, T.; Ouyang, J.; et al. Short-term clinical effects of tolvaptan, an oral vasopressin antagonist, in patients hospitalized for heart failure: The EVEREST Clinical Status Trials. Efficacy of Vasopressin Antagonism in Heart Failure Outcome Study with Tolvaptan (EVEREST) Investigators. JAMA 2007, 297, 1332–1343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Udelson, J.E.; McGrew, F.A.; Flores, E.; Ibrahim, H.; Katz, S.; Koshkarian, G.; O’Brien, T.; Kronenberg, M.W.; Zimmer, C.; Orlandi, C.; et al. Multicenter, randomized, double-blind, placebo-controlled study on the effect of oral tolvaptan on left ventricular dilation and function in patients with heart failure and systolic dysfunction. J. Am. Coll. Cardiol. 2007, 49, 2151–2159. [Google Scholar] [CrossRef] [Green Version]
  271. Urbach, J.; Goldsmith, S.R. Vasopressin antagonism in heart failure: A review of the hemodynamic studies and major clinical trials. Ther. Adv. Cardiovasc. Dis. 2021, 15, 1753944720977741. [Google Scholar] [CrossRef]
  272. McDonagh, T.A.; Metra, M.; Adamo, M.; Gardner, R.S.; Baumbach, A.; Böhm, M.; Burri, H.; Butler, J.; Čelutkienė, J.; Chioncel, O.; et al. 2021 ESC Guidelines for the diagnosis and treatment of acute and chronic heart failure. Eur. Heart J. 2021, 42, 3599–3726. [Google Scholar] [CrossRef] [PubMed]
  273. Goldsmith, S.R.; Burkhoff, D.; Gustafsson, F.; Voors, A.; Zannad, F.; Kolkhof, P.; Staedtler, G.; Colorado, P.; Dinh, W.; Udelson, J.E. Dual Vasopressin Receptor Antagonism to Improve Congestion in Patients With Acute Heart Failure: Design of the AVANTI Trial. J. Card Fail. 2021, 27, 233–241. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Main factors operating in the brain and peripheral organs participating in the regulation of the synthesis and release of oxytocin and vasopressin. Abbreviations: AP—area postrema; AVP—arginine vasopressin; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; OXY—oxytocin; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Figure 1. Main factors operating in the brain and peripheral organs participating in the regulation of the synthesis and release of oxytocin and vasopressin. Abbreviations: AP—area postrema; AVP—arginine vasopressin; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; OXY—oxytocin; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Ijms 22 11465 g001
Figure 2. The role of oxytocin in cardiovascular regulation. Oxytocin stimulates specific receptors (OXTR) and vasopressin receptors (V1aR) in different organs. In the central nervous system, it regulates the activity of several groups of neurons and may exert either stimulatory or inhibitory effects, depending on the site of action. The predominant effects of the action of oxytocin in the central nervous system involve reduced activation of the sympathetic division of the autonomic nervous system and enhanced activation of the parasympathetic division of the autonomic nervous system. Abbreviations: AP—area postrema; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Figure 2. The role of oxytocin in cardiovascular regulation. Oxytocin stimulates specific receptors (OXTR) and vasopressin receptors (V1aR) in different organs. In the central nervous system, it regulates the activity of several groups of neurons and may exert either stimulatory or inhibitory effects, depending on the site of action. The predominant effects of the action of oxytocin in the central nervous system involve reduced activation of the sympathetic division of the autonomic nervous system and enhanced activation of the parasympathetic division of the autonomic nervous system. Abbreviations: AP—area postrema; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Ijms 22 11465 g002
Figure 3. The role of vasopressin in cardiovascular regulation. Vasopressin stimulates specific types of receptors (V1aR, V1bR, and V2R) in various organs. In the central nervous system, it regulates the activity of several groups of neurons and exerts either stimulatory or inhibitory effects, depending on the site of action. Consequently, the effects of activation of both divisions of the autonomic nervous system (sympathetic and parasympathetic) can be observed, depending on specific factors provoking activation of the vasopressinergic system. Abbreviations: AP—area postrema; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Figure 3. The role of vasopressin in cardiovascular regulation. Vasopressin stimulates specific types of receptors (V1aR, V1bR, and V2R) in various organs. In the central nervous system, it regulates the activity of several groups of neurons and exerts either stimulatory or inhibitory effects, depending on the site of action. Consequently, the effects of activation of both divisions of the autonomic nervous system (sympathetic and parasympathetic) can be observed, depending on specific factors provoking activation of the vasopressinergic system. Abbreviations: AP—area postrema; MnPO—median preoptic nucleus; NTS—nucleus tractus solitarii; OVLT—organum vasculosum of the lamina terminalis; PVN—paraventricular nucleus; RVLM—rostral ventrolateral medulla; 3rd—third ventricle; 4th—fourth ventricle.
Ijms 22 11465 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Szczepanska-Sadowska, E.; Wsol, A.; Cudnoch-Jedrzejewska, A.; Żera, T. Complementary Role of Oxytocin and Vasopressin in Cardiovascular Regulation. Int. J. Mol. Sci. 2021, 22, 11465. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222111465

AMA Style

Szczepanska-Sadowska E, Wsol A, Cudnoch-Jedrzejewska A, Żera T. Complementary Role of Oxytocin and Vasopressin in Cardiovascular Regulation. International Journal of Molecular Sciences. 2021; 22(21):11465. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222111465

Chicago/Turabian Style

Szczepanska-Sadowska, Ewa, Agnieszka Wsol, Agnieszka Cudnoch-Jedrzejewska, and Tymoteusz Żera. 2021. "Complementary Role of Oxytocin and Vasopressin in Cardiovascular Regulation" International Journal of Molecular Sciences 22, no. 21: 11465. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222111465

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop