Skip to main content

REVIEW article

Front. Psychiatry, 22 September 2015
Sec. Systems Biology Archive
This article is part of the Research Topic Frontiers in Synaptic Plasticity: Dendritic Spines, Circuitries and Behavior View all 11 articles

Examining the Role of Vasopressin in the Modulation of Parental and Sexual Behaviors

Josi Maria Zimmermann-PeruzattoJosi Maria Zimmermann-Peruzatto1Virgínia Meneghini LazzariVirgínia Meneghini Lazzari2Ana Carolina de MouraAna Carolina de Moura2Silvana AlmeidaSilvana Almeida2Mrcia Giovenardi,*\r\n   Márcia Giovenardi1,2*
  • 1Programa de Pós-Graduação em Fisiologia, Instituto de Ciências Básicas da Saúde (ICBS), Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre, Brazil
  • 2Programa de Pós-Graduação em Ciências da Saúde, Universidade Federal de Ciências da Saúde de Porto Alegre (UFCSPA), Porto Alegre, Brazil

Vasopressin (VP) and VP-like neuropeptides are evolutionarily stable peptides found in all vertebrate species. In non-mammalian vertebrates, vasotocin (VT) plays a role similar to mammalian VP, whereas mesotocin and isotocin are functionally similar to mammalian oxytocin (OT). Here, we review the involvement of VP in brain circuits, synaptic plasticity, evolution, and function, highlighting the role of VP in social behavior. In all studied species, VP is encoded on chromosome 20p13, and in mammals, VP is produced in specific hypothalamic nuclei and released by the posterior pituitary. The role of VP is mediated by the stimulation of the V1a, V1b, and V2 receptors as well as the oxytocinergic and purinergic receptors. VT and VP functions are usually related to osmotic and cardiovascular homeostasis when acting peripherally. However, these neuropeptides are also critically involved in the central modulation of social behavior displays, such as pairing recognition, pair-bonding, social memory, sexual behavior, parental care, and maternal and aggressive behavior. Evidence suggests that these effects are primarily mediated by V1a receptor in specific brain circuits that provide important information for the onset and control of social behaviors in normal and pathological conditions.

General Aspects

Vasopressin-like (VP-like) and oxytocin-like (OT-like) peptides have been isolated from invertebrates and vertebrates in more than 100 species (1). Approximately 700 million years ago, the ancestral gene encoding the precursor protein diverged between the invertebrate and vertebrate families (2). Generally, all vertebrate species express a VP-like and OT-like peptide.

The lineage of VP-like peptides is evolutionarily stable. In fishes, amphibians, reptiles, and birds, vasotocin (VT) shares similar roles with mammalian arginine vasopressin (VP), whereas mesotocin and teleost isotocin are functionally similar to mammalian oxytocin (OT) (3).

The gene expression and regulation of these peptides is conserved among vertebrates (3). The chemical structure of VP and OT differs by only two of nine amino acid residues (4, 5). The VP/OT superfamily can be traced back to different types of invertebrates, such as annelids and mollusks (6). There is a homology of 80% between VP and OT, but these neuropeptides have distinct physiological activities. The receptors for VP-like and OT-like peptides have been described in invertebrates (6) and vertebrates (7).

Both VP and OT are produced in the hypothalamus, released in the neurohypophysis, and distributed throughout the brain (8, 9). In 1895, the vasopressor effect of the neurohypophyseal extract was attributed to the neurohypophysis gland (10). However, two decades later, the antidiuretic effect of the pituitary extract was demonstrated (11). The isolation of VP in the fifties confirmed that the same neuropeptide is synthesized in the neurohypophysis gland and possesses antidiuretic and vasopressor effects (12).

This review focuses on the involvement of VP in brain circuits, synaptic plasticity, evolution, and function, highlighting the role of VP in parental and sexual behaviors.

Vasopressin and Vasopressin-Like Peptides

Gene Structure

Nucleotide sequences encoding the VP and OT hormones are highly homologous. In all species, VP and OT are located on the same chromosome, 20p13, but encoded by different genes separated by a segment of DNA only 12 kb long (13). The similarity in the intron–exon structures of the two genes and opposite orientation suggest recent gene duplication (14).

Synthesis and Release

Vasopressin is a nonapeptide with a disulfide bridge between two cysteine amino acids and is synthesized in a smaller amount by parvocellular neurons, primarily by magnocellular neurons of the hypothalamus in paraventricular nucleus (PVN) and supraoptic nucleus (SON) (15). These nuclei send axons to the neuro pituitary along the supraoptic–hypophyseal tract. In a prohormone state, VP migrates along the supraoptic–hypophyseal tract to the neurohypophysis gland, where it is released into circulation (16).

The PVN and SON receive afferent nerve impulses from receptors in the left atrium, aortic arch, and carotid sinuses via the vagus nerve. PVN and SON receive osmotic input from the lamina terminalis, which is excluded from the blood–brain barrier and is thus affected by systemic osmolality. Furthermore, Holmes et al. (16) suggested that in the rat brain, extrahypothalamic structures, such as the bed nucleus of the stria terminalis (BNST), the medial amygdala, nucleus of the locus coeruleus, hippocampus, and choroid plexus, in addition to the hypothalamus, are able to synthesize VP.

The anterior pituitary gland also releases VP but in smaller quantities. VP can activate the hypothalamic–pituitary–adrenal axis by stimulating the V1b receptor (V1bR) and controlling the liberation of adrenocorticotropic hormone (ACTH) (17), whereas the V1a receptor (V1aR) controls the synthesis and release of cortisol in the adrenal cortex (18).

Vasopressin Receptors

The functions of VP are modulated by stimulation of G-protein-coupled receptors (GPRCs) from each independent tissue. They are classified as V1aR, V1bR (also known as V3R), V2R, oxytocinergic (OTR), and P2 purinergic (P2R) receptors (19, 20). In mammals, the VPRs are widely distributed in the brain (21, 22).

Vasopressin can bind to all of these receptors but not with the same affinity. The receptors, which use G-proteins as transducer signals across the cell membranes, have seven hydrophobic transmembrane domains, four extracellular domains, and four intracellular domains (23). Neurotransmitters, hormones, and chemokines indicate the courses of VP action, whereas local mediators signal to the four main G-protein families to regulate metabolic enzymes, ion channels, and transcriptional regulators. The different types of VPR extracellular signals refer to specific G-proteins. Several important hormones interact with the Gi pathway, which is characterized by inhibition of adenylyl cyclase (24).

Agonist of VPR is a substance that initiates a physiological response through specific interactions with G-protein-coupled receptor kinases and protein kinase C present in the carboxyl termini of the receptors (25). The VP signal is transmitted through guanine nucleotide-binding proteins (G-proteins) (26), such as Gs and Gq/11 subtypes (24).

V1aR Receptor

V1aR is primarily found on vascular smooth muscle and causes vasoconstriction by an increase in intracellular calcium via the phosphatidyl-inositol-bisphosphonate cascade. Studies in rats have shown that V1aR is also located in the brain, myocardium, gonads, cervical ganglion, liver, blood vessels, kidney, spleen, renal medulla, and platelets (2022, 2729); however, the physiologic roles of VP remain unknown in many tissues.

V1bR Receptor

V1bR is localized in pituitary gland, olfactory bulb, septum, hippocampus, pancreatic beta cells, and adrenal medulla and induces the release of hormones (20, 3032). The phylogenetic analysis showed that V1bR diverged early from the V1aR sequences and presented the closest relationship with the OTR (31). V1bR is highly expressed in the anterior pituitary where it is thought to play a role in costimulating the neuroendocrine response to stress (33). The VP causes secretion of ACTH, which is important for the induction and phenotype maintenance of ACTH-secreting tumors mediated through Gs, Gi, and Gq/11 (34). Studies have shown that V1bR gene expression may thus be a marker of the corticotroph phenotype and can be used to help shed light on the pathophysiological mechanism of ectopic ACTH syndrome (30, 35).

V2 Receptor

V2R is located on vascular smooth muscle cells, vascular endothelium, and the collecting ducts of the renal medulla (20). The hydro-osmotic or antidiuretic effect of VP occurs via activation of V2R (36, 37). VP adjusts water homeostasis regulation of the fast shuttling of aquaporin 2 to the cell surface and stimulates the synthesis of mRNA encoding this protein (38, 39).

Phillips et al. (28) evaluated V1Rs and V2R binding sites in vitro using selective radioligands and demonstrated that there was no vasoconstrictor activity of V2R in the endothelium, liver, brain, spinal cord, sympathetic ganglia, heart, or vascular smooth muscle. In this study, specific binding was only identified in the kidney, which is consistent with the known distribution of antidiuretic V2Rs on renal collecting tubules (28).

Other Receptors

Both VP and OT can bind with OTR but not with the same affinity. OTRs are coupled to Gq/11 class binding proteins, which stimulate phospholipase C activity (34).

P2 purinergic receptors also belong to the seven-transmembrane domain GPCR superfamily. Its role was confirmed in cardiac endothelium because VP exhibited effects through activation of P2Rs (40).

Vasopressin and Synaptic Plasticity

Vasotocin/vasopressin neuropeptides affect several sex-typical and species-specific behaviors and produce an integrational neural substrate for the dynamic regulation of these behaviors via endocrine and sensory stimuli. Different types of social behaviors are influenced by VT and VP, influencing the objectives of many neuroanatomical studies of VT/VP distribution and central nervous system (CNS) targets (4, 41, 42). There is a sexually dimorphic vasopressinergic extrahypothalamic network that plays a key role in the modulation of behavior by VP (43), and sex differences in VP distribution were first demonstrated in rats (44).

In vertebrates, behaviors such as reproductive, smell recognition, social communication, pair-bonding, parental care, and aggressive behavior are also modulated by VT/VP (41, 45, 46). In different species of vertebrates, both central and peripheral VT/VP administration stimulate spawning behavior, phonotaxis, sexual receptivity, lordosis, courtship, and mating (4, 41, 47).

In the rat brain, V1aR is believed to play the predominant role in regulating behavior. V1aR is found in neural networks involved in responses related to social behavior and exhibits considerable plasticity (45, 46, 48). V1aR is expressed in olfactory bulb, hippocampal dentate gyrus, cerebellum, septal nuclei, accumbens nucleus, arcuate nuclei, suprachiasmatic nuclei, and periventricular nuclei and the lateral hypothalamic area, parvocellular paraventricular and anteroventral nucleus of the thalamus, circumventricular organs including the pineal, and the subfornical organ (49, 50). Further, V1bR also participates in the neural regulation of social behaviors (51), but has received much less attention due to a lack of specific drugs (52, 53). V1bR transcripts and immunoreactive cell bodies are localized to the cerebellum, cerebral cortex, hippocampus, olfactory bulb (including in the area periglomerular), PVN, piriform cortex layer II, red nucleus, septum, and suprachiasmatic nucleus (5461).

Vasopressinergic as well as oxytocinergic systems can be modulated by circulating gonadal steroids and together are involved in behavioral regulation (4, 41, 62, 63). Sex hormones affect the quantity and localization of the VT/VP receptors in the brain (64). Previous studies (64, 65) have shown that behavioral responses, such as courtship behavior in newts, aggression in hamsters, and parental behaviors in voles (66), depend on VT/VP and the presence of gonadal steroids.

Role of Vasopressin in Parental and Sexual Behavior

The ability of an animal to recognize intra- and interspecific individuals is essential for all complex relationships. Several studies (5, 46, 63, 6771) have shown that VP participates in the modulation of non-social and social behaviors.

In this sense, vasopressinergic neurons play an important role in coding information from social contact (47). In rodents, social information is primarily mediated by the exchange of olfactory information, and there is evidence that VP signaling is important in brain areas where olfactory information is processed. Wacker et al. (51) described populations of vasopressinergic neurons in the main and accessory olfactory bulbs and anterior olfactory nucleus that are involved in processing social odor cues. Pharmacological studies have shown that VP administration improves social recognition in both sexes (58, 63). If applied bilaterally in the olfactory bulbs, extending the memory retention interval for the recognition of male rat odor is extended (72).

In rodents, intracerebroventricular microinjections of VP agonists facilitate social memory and can significantly extend social recognition and memory consolidation for as much as 120 min (73, 74). V1aR antagonists produce marked effects on learning, memory, and social behaviors (59, 63). Intracerebroventricular and intraseptal microinjection of V1aR antagonists block social recognition, and VP microinjection can rescue deficits in social recognition in Brattleboro rats that lack VP (7477).

Bielsky et al. (45) demonstrated that V1aR knockout mice (V1aRKO) display an enormous deficit in social recognition, providing strong evidence that V1aR is essential for this action. Similarly, V1bR knockout mice (V1bRKO) also presented deficits in the social memory test (52, 53). However, the increased V1aR expression in the lateral septum (LS) facilitated social behavior (78).

Among other social behaviors, the parental behavior exerts an essential role in the behavioral development of offspring (79). Maternal care is best studied; however, pup-directed positive behaviors, such as retrieval and kyphosis (huddling), are exhibited by no-parturient animals and are characteristics of socially monogamous or cooperative breeding species. Male parental care has been primarily studied in relation to the vasopressinergic system, which is sexually dimorphic and androgen-dependent because testosterone promotes VP synthesis (79).

Paternal care is highly demonstrated by prairie vole males (80, 81). Lim and Young (82) demonstrated that the infusion of VP (0.1 ng) directly into the LS of these animals can enhance licking/grooming (LG) behavior toward pups and that this behavior is blocked by the use of a V1aR antagonist (80). The father LG behaviors and retrieving the pups exert a pivotal role on the development of the VP systems and aggressive behavior of their adult offspring (79). In adult male rats, LG behaviors increase the levels of V1aR binding within the amygdale nucleus (46).

Besides the parental behavior, the VP system is also an important mediator of maternal behavior. In the peripartum and lactation periods, there is an increase in the expression of mRNA, receptors, and density/binding of VP in the maternal brain. This increase contributes to the adaptations that develop in the female maternal care. Previous studies (8385) have already suggested a role for VP facilitating maternal care, but more recent studies showed substantial evidence of this neuropeptide facilitating maternal behavior [for review, see Bosch and Newman (86)]. Furthermore, in females, V1aR density was significantly correlated with postpartum LG of the offspring (87).

Most studies are based on infusion of VP and V1aR antagonists in specific areas of the CNS participating in the neural circuitry of maternal behavior and thus demonstrating the modulation of VP in this behavior [for review, see Ref. (86, 8891)].

Moreover, previous studies (92, 93) with HAB (high anxiety-related behavior) and LAB (low anxiety-related behavior) dams also reinforce that central VP modulates the maternal behavior because the blocking of V1aR by repeated acute intracerebroventricular administration of a selective antagonist promoved decrease arched back nursing and the time the dam spent with the pups in HAB dams.

In conjunction with these findings, the involvement of V1bR in the modulation of maternal behavior was demonstrated in V1bRKO mice (94), which showed that lactating females who received the V1bR antagonist in the lateral ventricle decreased the nursing and interaction with their pups.

In addition to maternal care, lactating rats exhibit aggressive behavior that is observed during the first two postpartum weeks and which aims to protect the offspring against a potentially dangerous intruder (95, 96). The role of VP in aggression has received attention, and previous studies (86, 88, 89, 91, 97100) provide substantial evidence for VP promoting maternal aggressive behavior. Neuroanatomical studies with multiparous rats revealed significant increases in V1aR mRNA expression in the amygdala, SON, and LS in females on the fifth postpartum day when compared with primiparous rats (101, 102).

Bosch and Neumann (90) demonstrated that microinjection of a selective V1aR antagonist bilaterally into the BNST reduced maternal aggression (91) and the VP within the central nucleus of the amygdala (CeA) was positively correlated with the increased offensive behavior (90). Furthermore, other study by Bosch and Neumann (92) reported that in HAB rats VP promotes maternal aggression, and Lonstein and Gammie (103) showed the increased expression of the VP gene in the PVN. On the other hand, studies with Sprague-Dawley lactating rats showed different results, as the intracerebroventricular infusions of VP reduced maternal aggression, while treatments with an V1aR antagonist increased maternal aggression during early lactation (88).

Vasopressin-deficient Brattleboro rats exhibit reduced aggressive maternal behavior and reduced attacks in males without sexual experience against intruders (99). In other study on female pregnancy (51, 104), increasing V1aR levels in the PVN, CeA, and LS were positively correlated with aggressive behavior. The fluctuations observed in the OTR and V1aR in important areas of the CNS appear to regulate maternal aggression during the peripartum period. In a pharmacological study, microinjection of a selective V1aR antagonist bilaterally into the BNST reduced maternal aggression behavior but did not alter maternal care (105).

Furthermore, VP may be a new target for studies on treatments involving V1aR antagonists or synthetic VP to promote maternal care or suppress aggression in lactating females exposed to chronic stress-associated disorders (98). Using a model of VP-deficient mothers, Fodor et al. (105) demonstrated that these rats have decreased LG behavior and act less depressive. Thus, they suggest that VP antagonists could be an option for future studies on postpartum depression; however, the possible side effects of maternal neglect require further investigation (90, 105).

Some findings suggest that V1bR might also be involved in the modulation of aggressive behavior in both females and males. V1bRKO lactating mice showed an increased latency and decreased number of attacks against intruders when compared with wild-type mothers (97). The defensive behavior was studied in V1bRKO male mice and increased social behavioral responses were observed (106). In support, V1bRKO male mice also had impaired attack behavior toward a conspecific (97).

The role of this nonapeptide in sexual behavior has also been described in the past decades, and previous studies showed that VT/VP modulates specific types of vocalization in rats and squirrel monkeys (107). VT/VP central administration enhances pair-bonding and smell-recognition behaviors as a key feature for the onset of sexual behavior (66, 107).

Since the 1980s, studies have described the importance of vasopressinergic projections in both sexes in relation to sexual behavior. In this context, Meyerson et al. (104) administered an antagonist of VPRs into the lateral ventricles of female Sprague-Dawley rats in the neonatal period; this treatment induced a persistent increase in VP content and facilitated female sexual behavior. This study provided functional and immunocytochemical evidence of the importance of VP in female sexual behavior. In the same decade, Södersten et al. (108) reported that intracerebroventricular injections of VP inhibit sexual behavior in receptive female Wistar rats. These findings were reinforced by Pedersen and Boccia (109), which suggested that VP influences ovarian steroid activation of female sexual behavior via interactions with OT.

In males, Bohus (110) observed that VP agonists reversed the decrease of sexual behavior after castration, but injections of VP into LS of males did not dramatically alter sexual behavior (111, 112). In the 1990s, the role of VP in the modulation of male sexual behavior was hypothesized to be due to the refractory period in rats and consequently be an inhibitor of sexual behavior in males (112). In 1993, Winslow et al. (66) reported that VP is necessary to partner preference formation in monogamous prairie vole. However, these studies commonly focused on differences in neurotransmitter systems in brain structure between sexes instead of the role of the neurotransmitter on sexual behavior. Knowledge about differences in cell density, neurotransmitter content, receptors distribution, and vasopressinergic projections in rats and voles between sexes does not necessarily cause sex differences in sexual behavior (113).

Knockout animals were also used to investigate the role of VP and its receptors in sexual behavior. V1bRKO mice exhibited deficits in social behaviors that require olfactory function, including aggression and social recognition, but these animals had normal sexual behavior (52). The BNST is a sexually dimorphic structure that can be involved in the control of male sexual behavior because males have more VP neurons and denser projections from this area and in the medial amygdaloid nucleus than females (114). However, studies about VP innervation have focused on female sexual behavior. VP innervation from the LS inhibits sexual behavior in females; thus, hypothetically, the higher levels of VP in males are correlated with less lordosis behavior (115).

In humans, VP has been reported as a selective enhancer of recognition of sexual cues in a behavioral task administered to males (116). Additionally, Argiolas and Melis (117) conducted an elegant review about the control of sexual behavior by neuropeptides in the species studied thus far, including rats, mice, monkeys, and humans (117), and describe VP as an ineffective neuropeptide on copulatory behavior in males but as an inhibitor of lordosis in female sexual behavior.

The role of VP in sexual behavior remains unclear. Controversial results can be explained by the different methods applied and the interactions of this peptide with others, which should always be considered.

In conclusion, the lineage for VP or VP-like neuropeptides is evolutionarily stable and present in all vertebrate species. The neural distribution of VP neurons and its projections and also the distribution of VPRs are widely investigated for more than 30 years. These extensive data collected shows that male and female have important differences in the VP distribution across the CNS. The role of VP in osmotic and cardiovascular homeostasis is well established in the literature; however, the neuropeptides VT/VP are also critically involved in the modulation of social behaviors in many species. In mammals, the parental and sexual behaviors are mediated by VT/VP system and mainly by V1aR in specific brain circuits that provide important information for the onset and control of these behaviors in normal and pathological conditions. Future investigations must be conducted to further elucidate the involvement of V1bR modulating social behaviors.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

1. Hoyle CH. Neuropeptide families and their receptors: evolutionary perspectives. Brain Res (1999) 848(1–2):1–25. doi: 10.1016/S0006-8993(99)01975-7

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Acher R, Chauvet J, Chauvet MT. Man and the chimaera. Selective versus neutral oxytocin evolution. Adv Exp Med Biol (1995) 395:615–27.

PubMed Abstract | Google Scholar

3. Gilligan P, Brenner S, Venkatesh B. Neurone-specific expression and regulation of the puffer fish isotocin and vasotocin genes in transgenic mice. J Neuroendocrinol (2003) 15:1027–36. doi:10.1046/j.1365-2826.2003.01090.x

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Rose JD, Moore FL. Behavioral neuroendocrinology of vasotocin and vasopressin and the sensorimotor processing hypothesis. Front Neuroendocrinol (2002) 23:317–41. doi:10.1016/S0091-3022(02)00004-3

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Mavani GP, Devita MV, Michelis MF. A review of the nonpressor and nonantidiuretic actions of the hormone vasopressin. Front Med (2015) 2:19. doi:10.3389/fmed.2015.00019

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Van Kesteren RE, Tensen CP, Smit AB, Van Minnen J, Kolakowski LF, Meyerhof W. Co-evolution of ligand–receptor pairs in the vasopressin/oxytocin superfamily of bioactive peptides. J Biol Chem (1996) 271:3619–26. doi:10.1074/jbc.271.7.3619

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Warne JM. Cloning and characterization of an arginine vasotocin receptor from the euryhaline flounder Platichthys flesus. Gen Comp Endocrinol (2001) 122:312–9. doi:10.1006/gcen.2001.7644

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Robertson GL. The regulation of vasopressin function in health and disease. Recent Prog Horm Res (1976) 33:333–85.

PubMed Abstract | Google Scholar

9. Ludwig M. Dendritic release of vasopressin and oxytocin. J Neuroendocrinol (1998) 10(12):881–95. doi:10.1046/j.1365-2826.1998.00279.x

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Oliver H, Schaefer EA. On the physiological action of extracts of the pituitary body and certain other glandular organs. J Physiol (1895) 18:277–9. doi:10.1113/jphysiol.1895.sp000565

CrossRef Full Text | Google Scholar

11. von den Velden R. The renal effects of hypophyseal extract in humans [in German]. Berl Klin Wochenscgr (1913) 50:2083–6.

Google Scholar

12. Turner RA, Pierce JG, Du Vigneaud V. The purification and the amino acid content of vasopressin preparation. J Biol Chem (1951) 191:21–8.

Google Scholar

13. Rao VV, Löffler C, Battey J, Hansmann I. The human gene for oxytocin-­neurophysin I (OXT) is physically mapped to chromosome 20p13 by in situ hybridization. Cytogenet Cell Genet (1992) 61:271–3. doi:10.1159/000133420

PubMed Abstract | CrossRef Full Text | Google Scholar

14. 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) 260:10236–41.

PubMed Abstract | Google Scholar

15. Caldwell HK, Lee HJ, Macbeth AH, Young WS III. Vasopressin: behavioral roles of an “original” neuropeptide. Prog Neurobiol (2008) 84:1–24. doi:10.1016/j.pneurobio.2007.10.007

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Holmes CL, Patel BM, Russell JA, Walley KR. Physiology of vasopressin relevant to management of septic shock. Chest (2001) 120(3):989–1002. doi:10.1378/chest.120.3.989

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Barsegyan AR, Atsak P, Hornberger WB, Jacobson PB, van Gaalen MM, Roozendaal B. The vasopressin 1b receptor antagonist A-988315 blocks stress effects on the retrieval of object-recognition memory. Neuropsychopharmacology (2015) 40(8):1979–89. doi:10.1038/npp.2015.48

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Pasquali R, Gagliardi L, Vicennati V, Gambineri A, Colitta D, Ceroni L, et al. ACTH and cortisol response to combined corticotropin releasing hormone-arginine vasopressin stimulation in obese males and its relationship to body weight, fat distribution and parameters of the metabolic syndrome. Int J Obes Relat Metab Disord (1999) 23(4):419–24. doi:10.1038/sj.ijo.0800838

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Holmes CL, Landry DW, Granton JT. Science review: vasopressin and the cardiovascular system part-I – receptor physiology. Crit Care (2003) 7:427–34. doi:10.1186/cc2338

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Tanındı A, Töre HF. Hiponatremi tedavisinde Vaptan kullanımı use of “Vaptans” in treatment of hyponatremia. Turk Kardiyol Dern Ars (2015) 43(3):292–301. doi:10.5543/tkda.2015.71508

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Griebel G, Holsboer F. Neuropeptide receptor ligands as drugs for psychiatric diseases: the end of the beginning? Nat Rev Drug Discov (2012) 11(6):462–78. doi:10.1038/nrd3702

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Wu N, Siyuan Shang S, Su Y. The arginine vasopressin V1b receptor gene and prosociality: mediation role of emotional empathy. PsyCh Journal (2015) 4(3):160–5. doi:10.1002/pchj.102

CrossRef Full Text | Google Scholar

23. Koshimizu TA, Nakamura K, Egashira N, Hiroyama M, Nonoguchi H, Tanoue A. Vasopressin V1a and V1b receptors: from molecules to physiological systems. Physiol Rev (2012) 92:1813–64. doi:10.1152/physrev.00035.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Neves SR, Ram PT, Iyengar R. G protein pathways. Science (2002) 296(5573):1636–9. doi:10.1126/science.1071550

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Berrada K, Plesnicher CL, Luo X, Thibonnier M. Dynamic interaction of human vasopressin/oxytocin receptor subtypes with G protein-coupled receptor kinases and protein kinase C after agonist stimulation. J Biol Chem (2000) 275(35):27229–37. doi:10.1074/jbc.M002288200

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Ross EM, Gilma AG. Biochemical properties of hormone sensitive adenylate cyclase. Ann Rev Biochem (1980) 49:533–64. doi:10.1146/annurev.bi.49.070180.002533

CrossRef Full Text | Google Scholar

27. Greenberg A, Verbalis JG. Vasopressin receptor antagonists. Kidney Int (2006) 69:2124–30. doi:10.1038/sj.ki.5000432

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Phillips PA, Abrahams JM, Kelly JM, Mooser V, Trinder D, Johnston CI. Localization of vasopressin binding sites in rat tissues using specific V1 and V2 selective ligands. Endocrinology (1990) 126:1478–84. doi:10.1210/endo-126-3-1478

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Thibonnier M, Graves MK, Wagner MS, Auzan C, Clauser E, Willard HF. Structure, sequence, expression, and chromosomal localization of the human V1a vasopressin receptor gene. Genomics (1996) 31(3):32734. doi:10.1006/geno.1996.0055

PubMed Abstract | CrossRef Full Text | Google Scholar

30. de Keyzer Y, Lenne F, Auzan C, Jégou S, René P, Vaudry H, et al. The pituitary V3 vasopressin receptor and the corticotroph phenotype in ectopic ACTH syndrome. J Clin Invest (1996) 97(5):1311–8. doi:10.1172/JCI118547

PubMed Abstract | CrossRef Full Text | Google Scholar

31. de Keyzer Y, Auzan C, Lenne F, Beldjord C, Thibonnier M, Bertagna X, et al. Cloning and characterization of the human V3 pituitary vasopressin receptor. FEBS Lett (1994) 356:215–20. doi:10.1016/0014-5793(94)01268-7

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Volpi S, Rabadan-Diehl C, Aguilera G. Vasopressinergic regulation of the hypothalamic pituitary adrenal axis and stress adaptation. Stress (2004) 7:75–83. doi:10.1080/10253890410001733535

PubMed Abstract | CrossRef Full Text | Google Scholar

33. El-Werfali W, Toomasian C, Maliszewska-Scislo M, Li C, Rossi NF. Haemodynamic and renal sympathetic responses to V1b vasopressin receptor activation within the paraventricular nucleus. Exp Physiol (2015) 100(5):553–65. doi:10.1113/expphysiol.2014.084426

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Peter J, Burbach H, Adan RA, Lolait SJ, van Leeuwen FW, Mezey E, et al. Molecular neurobiology and pharmacology of the vasopressin/oxytocin receptor family. Cell Mol Neurobiol (1995) 15:573–95. doi:10.1007/BF02071318

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Ci H, Wu N, Su Y. Clock gene modulates roles of OXTR and AVPR1b genes in prosociality. PLoS One (2014) 9(10):e109086. doi:10.1371/journal.pone.0109086

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Choe KY, Bourque CW. The osmotic control of vasopressin-releasing neurons. In: Armstrong WE, Tasker JG, editors. Neurophysiology of Neuroendocrine Neurons. Chichester, UK: John Wiley & Sons, Ltd (2014). doi:10.1002/9781118606803.ch4

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Birnbaumer M. Vasopressin receptors. Trends Endocrinol Metab (2000) 11:406–10. doi:10.1016/S1043-2760(00)00304-0

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Harris HW Jr, Zeidel ML, Jo I, Hammond TG. Characterization of purified endosomes containing the antidiuretic hormone-sensitive water channel from rat renal papilla. J Biol Chem (1994) 269:11993–2000.

PubMed Abstract | Google Scholar

39. Knepper MA, Inoue T. Regulation of aquaporin-2 water channel trafficking by vasopressin. Curr Opin Cell Biol (1997) 9:560–4. doi:10.1016/S0955-0674(97)80034-8

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Zenteno-Savin T, Sada-Ovalle I, Ceballos G, Rubio R. Effects of arginine vasopressin in the heart are mediated by specific intravascular endothelial receptors. Eur J Pharmacol (2000) 410(1):15–23. doi:10.1016/S0014-2999(00)00853-0

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Goodson JL, Bass AH. Social behavior functions and related anatomical characteristics of vasotocin/vasopressin systems in vertebrates. Brain Res Brain Res Rev (2001) 35:246–65. doi:10.1016/S0165-0173(01)00043-1

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Holley A, Bellevue S, Vosberg D, Wenzel K, Roorda S Jr, Pfaus G. The role of oxytocin and vasopressin in conditioned mate guarding behavior in the female rat. Physiol Behav (2015) 144:7–14. doi:10.1016/j.physbeh.2015.02.039

PubMed Abstract | CrossRef Full Text | Google Scholar

43. de Vries GJ, Miller MA. Anatomy and function of extrahypothalamic vasopressin systems in the brain. Prog Brain Res (1998) 119:3–20. doi:10.1016/S0079-6123(08)61558-7

PubMed Abstract | CrossRef Full Text | Google Scholar

44. de Vries GJ, Buijs RM, Swaab DF. Ontogeny of the vasopressinergic neurons of the suprachiasmatic nucleus and their extrahypothalamic projections in the rat brain – presence of a sex difference in the lateral septum. Brain Res (1981) 218(1–2):67–78. doi:10.1016/0006-8993(81)90989-6

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Bielsky I, Bao-Hu S, Szegda K, Westphal H, Young L. Profound impairment in social recognition and reduction in anxiety-like behavior in vasopressin V1a receptor knockout mice. Neuropsychopharmacology (2004) 29:483–93. doi:10.1038/sj.npp.1300360

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Albers HE. Species, sex and individual differences in the vasotocin/vasopressin system: relationship to neurochemical signaling in the social behavior neural network. Front Neuroendocrinol (2012) 35:49–71. doi:10.1016/j.yfrne.2014.07.001

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Ludwig M. Vasopressin and the olfactory system. Endocrine Abstr (2011). Available from: http://www.endocrine-abstracts.org/ea/0025/ea0025s7.1.htm

Google Scholar

48. Ferguson JN, Young LJ, Insel TR. The neuroendocrine basis of social recognition. Front Neuroendocrinol (2002) 23(2):200–24. doi:10.1006/frne.2002.0229

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Ostrowski NL, Lolait SJ, Young WS III. Cellular localization of vasopressin V1a receptor messenger ribonucleic acid in adult male rat brain, pineal, and brain vasculature. Endocrinology (1994) 135(4):1511–28. doi:10.1210/endo.135.4.7925112

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Ostrowski NL. Oxytocin receptor mRNA expression in rat brain: implications for behavioral integration and reproductive success. Psychoneuroendocrinology (1998) 23(8):989–1004. doi:10.1016/S0306-4530(98)00070-5

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Wacker DW, Engelmann M, Tobin VA, Meddle SL, Ludwig M. Vasopressin and social odor processing in the olfactory bulb and anterior olfactory nucleus. Ann N Y Acad Sci (2011) 1220:106–16. doi:10.1111/j.1749-6632.2010.05885.x

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Wersinger SR, Ginns EI, O’Carroll AM, Lolait SJ, Young WS III. Vasopressin V1b receptor knockout reduces aggressive behavior in male mice. Mol Psychiatry (2002) 7(9):975–84. doi:10.1038/sj.mp.4001195

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Wersinger SR, Kelliherb KR, Zufallb F, Lolaitc SJ, O’Carrollc AM, Young WS III. Social motivation is reduced in vasopressin 1b receptor null mice despite normal performance in an olfactory discrimination task. Horm Behav (2004) 46(5):638–45. doi:10.1016/j.yhbeh.2004.07.004

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Lolait SJ, O’Carroll AM, Mahan LC, Felder CC, Button DC, Young WS III. Extrapituitary expression of the rat V1b vasopressin receptor gene. Proc Natl Acad Sci U S A (1995) 92:6783–7. doi:10.1073/pnas.92.15.6783

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Saito M, Sugimoto T, Tahara A, Kawashima H. Molecular cloning and characterization of rat V1b vasopressin receptor: evidence for its expression in extra-pituitary tissues. Biochem Biophys Res Commun (1995) 212:751–7. doi:10.1006/bbrc.1995.2033

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Barberis C, Tribollet E. Vasopressin and oxytocin receptors in the central nervous system. Crit Rev Neurobiol (1996) 10:119–54. doi:10.1615/CritRevNeurobiol.v10.i1.60

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Vaccari C, Lolait SJ, Ostrowski NL. Comparative distribution of vasopressin V1b and oxytocin receptor messenger ribonucleic acids in brain. Endocrinology (1998) 139:5015–33. doi:10.1210/endo.139.12.6382

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Stemmelin J, Lukovic L, Salome N, Griebel G. Evidence that the lateral septum is involved in the antidepressant-like effects of the vasopressin V1b receptor antagonist, SSR149415. Neuropsychopharmacology (2005) 30:35–42. doi:10.1038/sj.npp.1300562

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Hernando F, Schoots O, Lolait SJ, Burbach JPH. Immunohistochemical localization of the vasopressin V1b receptor in the rat brain and pituitary gland: anatomical support for its involvement in the central effects of vasopressin. Endocrinology (2001) 142:1659–68. doi:10.1210/endo.142.4.8067

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Tobin VA, Hashimoto H, Wacker DW, Takayanagi Y, Langnaese K, Caquineau C, et al. An intrinsic vasopressin system in the olfactory bulb is involved in social recognition. Nature (2010) 464:413–7. doi:10.1038/nature08826

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Stevenson EL, Caldwell HK. The vasopressin 1b receptor and the neural regulation of social behavior. Horm Behav (2012) 61:277–82. doi:10.1016/j.yhbeh.2011.11.009

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Grossmann R, Jurkevich A, Kohler A. Sex dimorphism in the avian arginine vasotocin system with special emphasis to the bed nucleus of the stria terminalis. Comp Biochem Physiol A MolIntegr Physiol (2002) 131:833–7. doi:10.1016/S1095-6433(02)00021-1

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Clipperton-Allen AE, Lee AW, Reyes A, Devidze N, Phan A, Pfaff DW, et al. Oxytocin, vasopressin and estrogen receptor gene expression in relation to social recognition in female mice. Physiol Behav (2012) 105(4):915–24. doi:10.1016/j.physbeh.2011.10.025

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Delville Y, Mansour KM, Ferris CF. Testosterone facilitates aggression by modulating vasopressin receptors in the hypothalamus. Physiol Behav (1996) 60:25–9. doi:10.1016/0031-9384(95)02246-5

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Wang Z, De Vries GJ. Testosterone effects on paternal behavior and vasopressin immunoreactive projections in prairie voles (Microtus ochrogaster). Brain Res (1993) 631:156–60. doi:10.1016/0006-8993(93)91203-5

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Winslow JT, Hastings N, Carter CS, Harbaugh CR, Insel TR. A role for central vasopressin in pair bonding in monogamous prairie voles. Nature (1993) 365:545–8. doi:10.1038/365545a0

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Donaldson ZR, Young LJ. Oxytocin, vasopressin, and the neurogenetics of sociality. Science (2008) 322:900–4. doi:10.1126/science.1158668

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Veenema AH, Neumann ID. Central vasopressin and oxytocin release: regulation of complex social behaviours. Prog Brain Res (2008) 170:261–76. doi:10.1016/S0079-6123(08)00422-6

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Veenema AH, Bredewold R, De Vries GJ. Sex-specific modulation of juvenile social play by vasopressin. Psychoneuroendocrinology (2013) 38:2554–61. doi:10.1016/j.psyneuen.2013.06.002

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Goodson JL, Thompson RR. Nonapeptide mechanisms of social cognition, behavior and species-specific social systems. Curr Opin Neurobiol (2010) 20:784–94. doi:10.1016/j.conb.2010.08.020

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Bredewold R, Smith CJW, Dumais MK, Alexa H, Veenema AH. Sex-specific modulation of juvenile social play behavior by vasopressin and oxytocin depends on social context. Front Behav Neurosci (2014) 8:216. doi:10.3389/fnbeh.2014.00216

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Dantzer R, Bluthe RM, Koob GF, Le Moal M. Modulation of social memory in male rats by neurohypophyseal peptides. Psychopharmacology (1987) 91:363–8. doi:10.1007/BF00518192

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Le Moal M, Dantzer R, Michaud B, Koob GF. Centrally injected arginine vasopressin (AVP) facilitates social memory in rats. Neurosci Lett (1987) 77(3):353–9. doi:10.1016/0304-3940(87)90527-1

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Engelmann M, Landgraf R. Microdialysis administration of vasopressin into the septum improves social recognition in Brattleboro rats. Physiol Behav (1994) 55:145–9. doi:10.1016/0031-9384(94)90022-1

PubMed Abstract | CrossRef Full Text | Google Scholar

75. van Wimersma Greidanus TB, Maigret C. The role of limbic vasopressin and oxytocin in social recognition. Brain Res (1996) 713:153–9. doi:10.1016/0006-8993(95)01505-1

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Everts HGJ, Koolhaas JM. Differential modulation of lateral septal vasopressin receptor blockade in spatial-learning, social recognition, and anxiety-related behaviors in rats. Behav Brain Res (1999) 99:7–16. doi:10.1016/S0166-4328(98)00004-7

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Landgraf R, Frank E, Aldag JM, Neumann ID, Ren X, Terwilliger EF, et al. Viral vector-mediated gene transfer of the vole V1a vasopressin receptor in the rat septum: improved social discrimination and active social behavior. Eur J Neurosci (2003) 18:403–11. doi:10.1046/j.1460-9568.2003.02750.x

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Frazier CR, Trainor BC, Cravens CJ, Whitney TK, Marler CA. Paternal behavior influences development of aggression and vasopressin expression in male California mouse offspring. Horm Behav (2006) 50:699–707. doi:10.1016/j.yhbeh.2006.06.035

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Bales KL, Kim AJ, Lewis-Reese AD, Sue Carter C. Both oxytocin and vasopressin may influence alloparental behavior in male prairie voles. Horm Behav (2004) 45:354–61. doi:10.1016/j.yhbeh.2004.01.004

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Wang Z, Ferris CF, De Vries GJ. Role of septal vasopressin innervation in paternal behavior in prairie voles (Microtus ochrogaster). Proc Natl Acad Sci U S A. (1994) 91:400–4. doi:10.1073/pnas.91.1.400

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Wang ZX, Liu Y, Young LJ, Insel TR. Hypothalamic vasopressin gene expression increases in both males and females postpartum in a biparental rodent. J Neuroendocrinol (2000) 12:111–20. doi:10.1046/j.1365-2826.2000.00435.x

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Lim MM, Young LJ. Neuropeptidergic regulation of affiliative behavior and social bonding in animals. Horm Behav (2006) 50:506–17. doi:10.1016/j.yhbeh.2006.06.028

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Pedersen CA, Ascher JA, Monroe YL, Prange AJ Jr. Oxytocin induces maternal behavior in virgin female rats. Science (1982) 216:648–50. doi:10.1126/science.7071605

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Pedersen CA, Caldwell JD, Johnson MF, Fort SA, Prange AJ Jr. Oxytocin antiserum delays onset of ovarian steroid-induced maternal behavior. Neuropeptides (1985) 6:175–82. doi:10.1016/0143-4179(85)90108-8

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Pedersen CA, Caldwell JD, Walker C, Ayers G, Mason GA. Oxytocin activates the postpartum onset of rat maternal behavior in the ventral tegmental and medial preoptic areas. Behav Neurosci (1994) 108:1163–71. doi:10.1037/0735-7044.108.6.1163

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Bosch OJ, Neumann ID. Both oxytocin and vasopressin are mediators of maternal care and aggression in rodents: from central release to sites of action. Horm Behav (2012) 61:293–303. doi:10.1016/j.yhbeh.2011.11.002

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Curley JP, Jensen CL, Franks B, Champagne FA. Variation in maternal and anxiety-like behavior associated with discrete patterns of oxytocin and vasopressin 1a receptor density in the lateral septum. Horm Behav (2012) 61(3):454–61. doi:10.1016/j.yhbeh.2012.01.013

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Nephew BC, Bridges RS. Central actions of arginine vasopressin and a V1a receptor antagonist on maternal aggression, maternal behavior, and grooming in lactating rats. Pharmacol Biochem Behavx. (2008) 91:77–83. doi:10.1016/j.pbb.2008.06.013

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Nephew BC, Byrnes EM, Bridges RS. Vasopressin mediates enhanced offspring protection in multiparous rats. Neuropharmacology (2010) 58:102–6. doi:10.1016/j.neuropharm.2009.06.032

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Bosch OJ, Neumann ID. Vasopressin released within the central amygdala promotes maternal aggression. Eur J Neurosci (2010) 31:883–91. doi:10.1111/j.1460-9568.2010.07115.x

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Bosch OJ, Pfortsch J, Beiderbeck DI, Landgraf R, Neumann ID. Maternal behavior is associated with vasopressin release in the medial preoptic area and bed nucleus of the stria terminalis in the rat. J Neuroendocrinol (2010) 22:420–9. doi:10.1111/j.1365-2826.2010.01984.x

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Bosch OJ, Neumann ID. Brain vasopressin is an important regulator of maternal behavior independent of dams’ trait anxiety. Proc Natl Acad Sci U S A (2008) 105(44):17139–44. doi:10.1073/pnas.0807412105

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Bosch OJ. Maternal nurturing is dependent on her innate anxiety: the behavioral roles of brain oxytocin and vasopressin. Horm Behav (2011) 59:202–12. doi:10.1016/j.yhbeh.2010.11.012

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Bayerl DS, Klampfl SM, Bosch OJ. Central V1b receptor antagonism in lactating rats: impairment of maternal care but not of maternal aggression. J Neuroendocrinol (2014) 26(12):918–26. doi:10.1111/jne.12226

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Koolhaas JM, Moor E, Hiemstra Y, Bohus B. The testosterone-dependent vasopressinergic neurons in the medial amygdala and lateral septum: involvement in social behaviour of male rats. In: Jard S, Jamison R, editors. Vasopressin. Vol. 208. Netherlands: Colloque INSERM/John Libbey Eurotext. (1991). p. 213–9.

Google Scholar

96. Giovenardi M, Padoin MJ, Cadore LP, Lucion AB. Hypothalamic paraventricular nucleus modulates maternal aggression in rats: effects of ibotenic acid lesion and oxytocin antisense. Physiol Behav (1998) 63:351–9. doi:10.1016/S0031-9384(97)00434-4

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Bosch OJ. Maternal aggression in rodents: brain oxytocin and vasopressin mediate pup defence. Phil Trans R Soc B (2013) 368:20130085. doi:10.1098/rstb.2013.0085

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Coverdill AJ, McCarthy M, Bridges RS, Nephew BC. Effects of chronic central arginine vasopressin (AVP) on maternal behavior in chronically stressed rat dams. Brain Sci (2012) 2:589–604. doi:10.3390/brainsci2040589

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Fodor A, Barsvari B, Aliczki M, Balogh Z, Zelena D, Goldberg SR, et al. The effects of vasopressin deficiency on aggression and impulsiveness in male and female rats. Psychoneuroendocrinology (2014) 47:141–50. doi:10.1016/j.psyneuen.2014.05.010

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Caughey SD, Klampfl SM, Bishop VR, Pfortsch J, Neumann ID, Bosch OJ, et al. Changes in the intensity of maternal aggression and central oxytocin and vasopressin V1a receptors across the peripartum period in the rat. J Neuroendocrinol (2011) 23:1113–24. doi:10.1111/j.1365-2826.2011.02224.x

PubMed Abstract | CrossRef Full Text | Google Scholar

101. Byrnes EM, Nephew BC, Bridges RS. Neuroendocrine and behavioral adaptations following reproductive experience in the female rat. In: Bridges R, editor. Neurobiology of the Parental Mind. San Francisco: Elsevier (2008).

Google Scholar

102. Nephew BC, Bridges RS. Central vasopressin V1a and oxytocin receptor levels in primiparous and multiparous lactating rats. World Congress Neurohypophysial Hormones. Regensburg (2007).

Google Scholar

103. Lonstein JS, Gammie SC. Sensory, hormonal, and neural control of maternal aggression in laboratory rodents. Neurosci Biobehav Rev (2002) 26:869–88. doi:10.1016/S0149-7634(02)00087-8

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Meyerson BJ, Höglund U, Johansson C, Blomqvist A, Ericson H. Neonatal vasopressin antagonist treatment facilitates adult copulatory behaviour infemale rats and increases hypothalamic vasopressin content. Brain Res (1988) 473:344–51. doi:10.1016/0006-8993(88)90864-5

CrossRef Full Text | Google Scholar

105. Fodor A, Klausz B, Pintér O, Daviu N, Rabasa C, Rotllant D, et al. Maternal neglect with reduced depressive-like behavior and blunted c-fos activation in Brattleboro mothers, the role of central vasopressin. Horm Behav (2012) 62:539–51. doi:10.1016/j.yhbeh.2012.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Wersinger SR, Caldwell HK, Christiansen M, Young WS. Disruption of the vasopressin 1b receptor gene impairs the attack component of aggressive behavior in mice. Genes Brain Behav (2007) 6:653–60. doi:10.1111/j.1601-183X.2006.00294.x

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Winslow J, Insel TR. Vasopressin modulates male squirrel monkeys behavior during social separation. Eur J Pharmacol (1991) 200:95–101. doi:10.1016/0014-2999(91)90671-C

PubMed Abstract | CrossRef Full Text | Google Scholar

108. Södersten P, De Vries GJ, Buijs RM, Melin P. A daily rhythm in behavioural vasopressin sensitivity and brain vasopressin concentrations. Neurosci Lett (1985) 58:37–41. doi:10.1016/0304-3940(85)90325-8

PubMed Abstract | CrossRef Full Text | Google Scholar

109. Pedersen CA, Boccia L. Vasopressin interactions with oxytocin in the control of female sexual behavior. Neuroscience (2006) 139:843–51. doi:10.1016/j.neuroscience.2006.01.002

PubMed Abstract | CrossRef Full Text | Google Scholar

110. Bohus B. Post castration masculine behavior in the rats: the role of hypothalamo-hypophyseal peptides. Exp Brain Res (1977) 28:8.

Google Scholar

111. Koolhaas JM, van den Brink THC, Boorsma F. Medial amygdala and aggressive behavior: interaction between testosterone and vasopressin. Aggress Behav (1990) 16:223–9.

PubMed Abstract | Google Scholar

112. Moltz H. E-series prostaglandins and arginine vasopressin in the modulation of male sexual behavior. Neurosci Biobehav Rev (1990) 14:109–15. doi:10.1016/S0149-7634(05)80166-6

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Garcia-Villalon AL, Garcia JL, Fernandez N, Monge L, Gomez B, Dieguez G. Regional differences in the arterial response to vasopressin: role of endothelial nitric oxide. Br J Pharmacol (1996) 118:1848–54. doi:10.1111/j.1476-5381.1996.tb15613.x

PubMed Abstract | CrossRef Full Text | Google Scholar

114. de Vries GJ, Södersten P. Sex differences in the brain: the relation between structure and function. Horm Behav (2009) 55:589–96. doi:10.1016/j.yhbeh.2009.03.012

PubMed Abstract | CrossRef Full Text | Google Scholar

115. de Vries GJ, Panzica GC. Sexual differentiation of central vasopressin and vasotocin systems in vertebrates: different mechanisms, similar endpoints. Neuroscience (2006) 138:947–55. doi:10.1016/j.neuroscience.2005.07.050

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Guastella AJ, Kenyon AR, Unkelbach C, Alvares GA, Hickie IB. Arginine Vasopressin selectively enhances recognition of sexual cues in male humans. Psychoneuroendocrinology (2011) 36:294–7. doi:10.1016/j.psyneuen.2010.07.023

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Argiolas A, Melis MR. Neuropeptides and central control of sexual behaviour from the past to the present: a review. Prog Neurobiol (2013) 108:80–107. doi:10.1016/j.pneurobio.2013.06.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: evolutionary lineage VP, vasopressin-like, V1a receptor, V1b receptor, social behaviors

Citation: Zimmermann-Peruzatto JM, Lazzari VM, de Moura AC, Almeida S and Giovenardi M (2015) Examining the role of vasopressin in the modulation of parental and sexual behaviors. Front. Psychiatry 6:130. doi: 10.3389/fpsyt.2015.00130

Received: 21 March 2015; Accepted: 04 September 2015;
Published: 22 September 2015

Edited by:

Rochelle S. Cohen, University of Illinois at Chicago, USA

Reviewed by:

Elliott Albers, Georgia State University, USA
Oliver J. Bosch, University of Regensburg, Germany

Copyright: © 2015 Zimmermann-Peruzatto, Lazzari, de Moura, Almeida and Giovenardi. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Márcia Giovenardi, Sarmento Leite Street 245/308, Division of Physiology, Federal University of Health Science, Porto Alegre, Rio Grande do Sul 90050-170, Brazil, giovenardi.marcia@gmail.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.