Skip to main content

Preoptic and hypothalamic regulation of multi-tiered, chronologically arranged female rat sexual behavior

Abstract

As in many mammalian behaviors, sexual behavior exhibits structure. Each modular components of the structure, that are linked together over time, occur in probabilistic manner. Endocrine milieu, in particular sex hormones, define the probability to synchronize the behavior with the production of gametes. Developmental experience and environmental cues affect the hormonal milieu of the brain. This is especially true in female mammals, in which ova mature with certain intervals along with ovarian secretion of sex hormones. Estrogens secreted by mature ovarian follicles support both affiliative and executive components of female sexual behavior. In the absence of the ovarian steroids, females avoid males when possible, or antagonize and reject males when put together. Female sexual behavior is intimately linked with the estrous cycle in many species such that females are only receptive for a brief period at the estrus stage surrounding ovulation. Thus, in the rat, females strongly influence the outcome of mating encounter with a male. Affiliative or solicitatory behavior shown by females in estrus leads to the female adapting the lordosis posture, which is characterized by hindleg postural rigidity and lordotic dorsiflexion of the spine, in response to touch-pressure somatosensory stimuli on the skin of the flanks, rump-tail base, perineum region given by male partner. The posture facilitates intromission and consequently fertilization. Although dependence on estrogens is the most important feature of female rat sexual behavior, cervical probing combined with palpation of the hindquarter skin acts as a supranormal stimulus to elicit lordosis. Thus, lordosis behavior is a hub of multi-tiered, chronologically arranged set of behaviors and estrogen appear to alter excitability of neural network for lordosis.

The lordosis reflex, dorsiflexion of the vertebral column (Fig. 1), is an essential element of female rat copulatory behavior. This stereotyped behavior depends strongly on estrogen and elicited by somatosensory stimuli on rump-tail base-perineal skin [1]. Odor cues influence female rats' mate choice of male partners [2, 3], but are not indispensable for the execution of the behavior [4]; the sighting of the particular male mating with another female do not necessarily alter females’ choice [5]. Ultrasound vocalization in the rat and mouse express their arousal and emotional states [6], however, is correlated with sexual motivation rather than the lordosis behavior [7].

Fig. 1
figure 1

A. Lordosis posture in a female rat in estrus, immediately after the individual was dismounted; B. Projections from the ventromedial nucleus of the hypothalamus (VMH) or the medial preoptic area (POA). Arrows depict sites of estrogen induced excitation (B, in red) or inhibition (C, in blue). Other CG midbrain central gray, VTA ventral tegmental area. Reference number for each connection is in parentheses

Neural axis for lordosis behavior

Sites for estrogen action

An earlier study by Barfield and Chen [8] showed stereotaxic implants of crystalline estrogen (estradiol benzoate) were effective in both the preoptic area (POA) and the ventromedial hypothalamic nucleus (VMH), particularly, its ventrolateral quadrant (vlVMH), in terms of lordosis and preceding affiliative behaviors, such as ear wiggling and darting, with the most intense responses were induced by the implants in the VMH. Electrical stimulation study revealed crucial role of the VMH in supporting lordosis reflex in ovariectomized female rats treated with subthreshold dose of estrogen [9]. Progressive increase in lordosis performance followed electrical stimulation of the VMH with a delayed onset and the effect persisted after termination of current application.

In contrast to the tardive response to VMH stimulation, electrical stimulation of the midbrain central gray (CG) caused a prompt and transient facilitation of lordosis [10]. The CG is a projection target of the VMH [11, 12]. CG neurons send axons to the medullary core, a site of origin of medullospinal projection which implement lordosis by innervating spinal motoneurons. Antidromic action potentials were recorded from CG neurons in response to stimulation of the medulla.

Thus, neural axis for the estrogen-dependent behavioral activation of lordosis originates in the VMH and descends to the medulla via the CG. Stimulation of the VMH increased the frequency of successful antidromic propagation into the somatodendritic complex of certain CG cells showing excitatory nature of the VMH innervation of these CG cells [13]. Estrogen treatment of the ovariectomized rat had a similar effect on the antidromic propagation as the VMH stimulation. ERα-expressing neurons project to the CG [14]. Estrogen enhances depolarizing action of N-methyl-D-aspartate (NMDA) and other molecules on VMH neurons by inhibiting K+ currents [15].

Inhibition of lordosis by POA projection to the midbrain

Along with the vlVMH, implants of crystalline estrogen in the medial POA also enhances lordosis and other affiliative behaviors albeit a larger dose of the hormone is needed [8]. The behavioral activation is, however, not a result of neural activation, but due to inhibition of POA efferents to the midbrain [16, 17]. Electrical stimulation of the POA reduced the probability of successful propagation into the somatodendritic complex in antidromic action potentials. Stimulation of the ventromedial nucleus of the hypothalamus (VMH) increased it. Conversely, electrolytic lesion of the POA facilitated, while VMH lesion reduced antidromic spike invasion.

Excitatory VMH innervation of the CG

Electrical stimulation studies [9] suggest that estrogen in the vlVMH causes behavioral activation of lordosis while causes behavioral disinhibition in the POA. Partial subsistence of sexual behavior in female rats with vlVMH lesion [18] can be a result of estrogen action on the POA. Cholecystokinin A receptor-expressing cells in the vlVMH were the key controllers of female sexual behaviors [19]. Work in mice identified the vlVMH as the attack center. Ablation of ERα-positive vlVMH neurons in females greatly diminished sexual receptivity and in males reduced mating and aggression [20]. Yang et al. [21] showed that the activity of these vlVMH neurons in males represents aggression performed by self and others as mirror neurons. Thus, the vlVMH likely mediates multiple social behaviors and future studies will need to address how factors including social experience, hormonal state, and behavioral context influence which behaviors are generated.

In addition to classical ERα, the vlVMH contains neurons with a putative G-protein coupled membrane ER called GPR30 [22]. GPR30 binds 17β-estradiol with an affinity similar as ERα and activates both PKA and extracellular-regulated kinase signaling pathways. Administration of G-1, a selective agonist for GPR30 [23], in an estradiol-progesterone priming paradigm, increased lordosis behavior in female mice [24]. GPR30 activation phosphorylate the classical ERα, showing that crosstalk with ERα is important in the display of this and other behaviors, many of which are absent in ERα-null mice [25]. GPR30-mediated phosphorylation may be also involved in the estrogen-dependent masculinization of the POA during development [26].

Progesterone and nonsteroidal molecules

Although the lordosis behavior can be elicited solely by estrogen in experimental settings, successive action of progesterone is needed to induce a full set of female rat sexual behavior. Estrogen induces progestin receptors or receptors for progesterone in the hypothalamus and preoptic area [27]. A series of pharmacological studies, initially conducted with substances such as GABA [28], gonadotropin releasing hormone (GnRH) [29, 30] and prostaglandin E2 [31] also modulated lordosis behavior in ovariectomized, estrogen-primed rats.

Other neurotropic drugs

Neurotropic non-steroidal molecules, like reserpine, and other peptides, e.g., oxytocin, prolactin, opioids, substitutes progesterone, but not estrogen, to induce lordosis behavior. GABA may depolarize hypothalamic neurons due to high chloride content as we have shown in GnRH neurons [32]. The diversity of effective agents suggests that multiple signaling pathways may be involved in the regulation of female sexual behavior [33]. GABA action may be enhanced by a direct action of progesterone on GABA receptors [34].

The VMH is also important for the modulation of lordosis by molecules other than ovarian steroids, but, nevertheless, estrogen pretreatment is a prerequisite for these effects. Infusions of agonists or antagonists of acetylcholine [35], norepinephrine [36], and serotonin [37, 38] into the VMH have been examined with various results. Moreover, infusion of neuropeptides, such as GnRH [39], and oxytocin [40] into the VMH has been shown to alter lordosis. Oxytocin is one of non-steroidal molecules which has been shown in earlier studies to facilitate lordosis and affiliative behaviors in estrogen-primed ovariectomized rats [41, 42]. Affiliative responses in females constitute proceptive behavior toward male partners which in turn initiate copulatory interactions which include lordosis in females. Gentle tactile stimulation of rump, tail-base and perineum induces affiliative responses in female rats in estrus. Infusion of oxytocin antagonist into the medial POA increases rejection and disrupts receptivity in estrogen-primed ovariectomized rats [43]. Because transcription of both oxytocin ligand and its receptor depends on estrogen, observations that oxytocin stimulates lordosis behavior in female rats might indicate oxytocin mediates secondary action to estrogen [41]. Further mediation of the oxytocin effect by prostaglandin E2 and GnRH signaling cascade has been shown [44]. The vlVMH as well as the adjacent neuropil are extremely rich in oxytocin binding sites, in addition to the medial POA which express smaller but definite oxytocin binding [45].

Descending POA projections

A well circumscribed, homogeneous lesion of the POA produced by local infusion of ibotenic acid enhanced lordosis and precipitated reductions in the proceptivity of estrogen-treated ovariectomized female rats [46] (Fig. 2A). The dose of estrogen which induced the receptivity in the lesioned rats was much smaller than that required in the sham-operated animals (Fig. 2B). The reduced proceptivity was characterized by the loss of behavioral sensitivity to estrogen. Ibotenic acid has been described as having several practical advantages over other excitotoxins in producing homogeneous lesions over a relatively large area, leaving fibers of passage unaffected [47]. The failure of prepubertal infusion of kainic acid into the POA to alter sexual behavior in non-ovariectomized rats has been ascribed to functional compensation, because, at the time of behavioral observation several weeks after the infusion, estrous cyclicity was present in the infused animals [48]. In the male rats, ibotenic acid lesion of the POA abolishes sexual motivation [49], but promotes the lordosis [50]. The decrease in the amount of estrogen needed to induce lordosis in the lesioned animals suggests that ibotenic acid removed POA neurons that would otherwise be inhibited by estrogen. Indeed, electrical stimulation of the POA suppresses lordosis (Fig. 2C), and electrophysiological recordings have shown that estrogen mostly inhibits neuronal activity in the POA [51]. In female rats carrying Electrical stimulation of the area of the ibotenic acid lesion, which activated fibers of passage with origins in the septum or cingulate cortex [52, 53]. The female rat POA contains many neurons that are targets of estrogen [54, 55]. With the dorsal inputs to the POA removed by anterior roof cut (Fig. 2D, E) [56], electrical stimulation of the deafferented POA caused prompt interruption of lordosis (Fig. 2F) without affecting proceptive behavior. Selective disruption of the stria medularis, which carries POA inputs from the amygdala, was effective as the anterior roof cut to interrupt lordosis (Fig. 2G). The removal of POA projections to the ventral tegmental area (VTA) which continues to midbrain regions [57, 58], disinhibited the receptivity. Behavioral effects of axons of passage in the POA [59] that were spared by the excitotoxin lesion, were examined by combined ibotenic acid lesion and focal stimulation of the POA. The neurotoxic effect of ibotenic acid is expressed postsynaptically via NMDA receptor [60], that has been found in large numbers in the forebrain [61]. POA infusion of an NMDA antagonist interrupted the lordosis in fully receptive females [62]; a reduced excitatory amino acid transmission to GnRH neurons has been associated with this effect.

Fig. 2
figure 2

A. Photomicrograph of a frontal section showing ibotenic acid lesion in the medial prcoptic area (POA), immunocytochemistry against glial fibrillary acidic protein. Arrows indicate damage by the injection needles. The lesion also caused enlarged third ventricle (V3). OC, optic chiasm. B. Ibotenic acid lesion of the POA enhanced the lordosis quotient (percent occurrence of lordosis per 10 mounts) in ovariectomized rats given 1.5 μg estradiol benzoate. Significantly fewer animals with the lesion showed solicitatory behavior; the rejection rate (percent interruption of copulatory interactions by females) was high in animals with the lesion. Open bars, control; shaded bars, lesioned. A and B; Reprinted from [46] with permission from Elsevier. C. Electrical stimulation of the POA caused slow-onset, long-lasting inhibition of lordosis score (LS, an arbitrary unit with maximal dorsiflexion set at 3) in ovariectomized, estrogen-treated rats. D. A scheme of the anterior roof cut of the POA by rotation of an L-shaped knife. The cut disrupted dorsal input to the POA at its junction to the lateral septum (LS), dorsal to the anterior commissure (AC). E. The anterior roof cut of the POA in frontal sections at levels a-a' and b–b′ in D. CG, midbrain central gray; VMH, ventromedial hypotalamic nucleus; VTA, ventral tegmental area. F. Electrical stimulation of the POA with parameters same as in C. Note short time course in both onset and recovery of the inhibition. G. Bilateral disruption of the stria medularis was as effective as the frontal roof cut to cause POA-bound prompt inhibition of lordosis. CF; Reprinted from [56] with permission from Elsevier. H, Stimulation sites in the VTA; filled circles, suppression of lordosis at currents below 50 μA; filled triangles, above 50 μA; open circles minor or no effect. Numbers denote distance from interaural line. CP cerebral peduncle, IP interpeduncular nucleus, ML medial lemniscus, P pons, R red nucleus, SN substantia nigra, III oculomotor nerve. VTA Stimulation caused prompt suppression of lordosis I. Reprinted from [63] with permission from Elsevier

The VTA

VTA stimulation in freely moving, estrogen-primed ovariectomized female rats caused a rapid and strong suppression of lordosis in response to either male mounts or manual cutaneous stimuli (Fig. 2H, I) [63]. The interruption occurred in a graded manner to increased stimulus intensity, with a low threshold. After the termination of electrical stimulation, lordosis performance returned promptly to the prestimulation level. No aversive response accompanied the blockade of lordosis. Electrical stimulation specifically blocked lordosis, without disrupting the preceptive behavior.

The POA contains a separate pool of neurons that promote proceptivity in addition to those that inhibit the receptivity. In particular, the excitability of POA neurons with axons to the VTA diminishes following estrogen in a sexually dimorphic pattern, in parallel to the capability of this steroid to induce the receptivity [16, 17].

The VTA is one of the major terminal fields of estrogen-concentrating POA neurons [64]. Electrical stimulation of axons in passage in the POA, which survived ibotenic acid lesion of the area [56] differed that the stimulation interrupted lordosis without inducing any proceptivity, whereas that of the deafferented POA [56] or the VTA [63] specifically blocked lordosis while preserving the proceptive components of female rat sexual behavior. Such dissociation of receptive and proceptive components of female sexual behavior also follows lesion or stimulation of several other brain structures [65, 66]. Among others, Whitney [67] showed that the higher lordosis quotients in female rats with POA lesion diminish if the females were allowed to evade male partner. The female could control her proximity to males. It appears that the POA contains a neural substrate for the promotion of proceptivity in addition to the suppression of lordosis. Large POA lesions or those including the rostral periventricular POA [68, 69] might also diminish lordosis through the destruction of local GnRH neurons or their axons [70]. ER positive neurons in the rostral periventricular POA [71] have been implicated in motivation of female mice sexual behavior through their innervation of GnRH neurons by kisspeptin [72].

The inhibition of proceptivity and receptivity obtained by electrical stimulation of focus of ibotenic acid lesion in the POA is due to the activation of fibers of passage which survived the excitotoxin challenge. In the lesioned animals, further transection to remove the POA afferents nullified the focal stimulation effect. The transaction was similar as the dorsal deafferentation of the POA by [73] which facilitated lordosis in non-lesioned animals (Fig. 2F, G). In the non-lesioned animals, electrical stimulation of the deafferented POA, which presumably activated local efferents, augmented the behavioral suppression through elimination of inputs via stria terminalis [56]. Thus, the POA is heterogenous not only in neuronal composition but also in axons penetrating the structure. The heterogeneity of the POA might have contributed to discrepancies among the reported POA effects on female rat sexual behavior.

The septum

The inhibitory effect of the septum on the lordosis reflex has been recognized [73, 74]. The basic organization of midbrain projections out of the lateral septum [75] is remarkably like that of the POA, and a massive projection of the septum projects caudally to the midbrain to end in the CG, VTA, dorsal and median raphe, and laterodorsal tegmental nucleus [76, 77]. Some of these midbrain structures have been implicated in the inhibitory regulation of the lordosis reflex [57, 63, 78]. The lateral septum and the POA also resemble each other in that both contain ER-positive neurons [55]. The qualitatively similar behavioral consequences of the deletion of local efferent neurons in the POA and the removal of efferent of the septum [56] present a coherent picture of the roles of the POA and septum in the regulation of lordosis as far as they are currently understood. In conclusion, the POA contains neurons that tonically inhibit the lordosis.

Motivational behavior in female rats

While anestrous female rats consistently avoid males, females in estrous approach and provoke males to initiate mounting. The series of affiliative or solicitatory behavior in females includes peculiar pattern of increased locomotor activity often called hopping and darting in front of the males. Increased locomotor activity in female rats in estrus embodies enhanced sexual motivation [79,80,81]. Estrogen action on ERα in the POA is responsible for the increased locomotion [82]. Subsets of ERα-positive POA neurons other than those involved in the inhibition of lordosis enhance locomotion. Behaviors such as hopping and darting have been used as indices of sexual motivation in females, as have other patterns of behavior such as seeking proximity to a sexually active male [66, 83, 84], the estrous female exhibits a sequence of events consisting of approach toward, orientation to, and rapid run away from the male [85]. The spinal stepping mechanism that induces locomotion is regulated by the midbrain locomotor region (MLR) in the rat [86], as originally shown in the cat [87]. In the rat, the MLR has been identified in the rostral part of the pedunculopontine and cuneiform nuclei of the midbrain [88]. Estrogen is responsible for the increased locomotion in female rats in estrus, because systemic administration of estrogen to ovariectomized rats increases both open-field [89] and wheel-running [90] activities.

The medial POA has been positively identified as a site for estrogen-induced activation of wheel running [91] while the brain site for the action of estrogen on the open-field activity [89, 92] has not been singled out. This may be due to the fact that the open-field activity is confounded by multiple factors, such as fear and emotionality. The preoptic area contributes to the rostro-caudal neural axis for the locomotor synergy [93] with its heavy projections to the MLR [88, 94]. The preoptic locomotor region [95] from which stepping can be initiated by chemical [96] or electrical [95] stimulation, is in the medial portion of the lateral POA. In contrast, the locomotor activity can be consistently reduced by cholinergic activation of the periventricular POA [97, 98]. Estrogen excites axons in the MLR with origins in the lateral POA, but inhibits those from the medial POA [99]. In the light of electrical and chemical stimulation of the POA, such diametric effects of estrogen would culminate in enhanced wheel running.

Estrogen action to alter axonal excitability

The opposite effects of estrogen on the medial and lateral POA neurons are reminiscent of our other observations. Estrogen decreased antidromic activation thresholds in axons in the CG with origins in the VMH [100], but increased the thresholds in POA axons in the VTA [17]. The effects of estrogen on the axonal excitability are sex specific [101] and can be modified by neonatal endocrine treatments [17, 100].

Estrogen-induced changes in the axonal excitability occur in long time-courses, which implicate the involvement of receptor-mediated, genomic activation [102]. Both the medial and lateral POA contain neurons that express ER [54, 103]. Estrogen promoted the activity of Na+,K+-transporting adenosine triphosphatase in the medial basal hypothalamus while it suppressed the activity in the POA [104]. The effects were also sex specific [105]. This enzyme can reach axonal terminals by axonal flow and alter its excitability [106]. In rat uterus, estrogen alters expression of K+ channels [107]. Estrogen activates phosphatidyl inositol pathway which in turn can increase protein kinase C activity, Ca2+ mobilization, and arachidonic acid metabolism [108]. With whole-cell patch clamp study in the GT1-7 cells, we have shown that estrogen at physiological concentrations augments K(Ca) currents via ERβ, at least partly by increasing the transcription of BK channel genes, thereby modifying the cellular excitability. A large part of the K(Ca) currents is likely to comprise BK currents [109].

Conclusion

Sexual behavior is a multitiered set of components which occur sequentially under the influence of sex hormones, particularly estrogen in female rodents. This characteristic has been thoroughly utilized in studies to identify and analyze neural mechanism for the behavior. Thus, separate neuronal circuitries, inter alia, those that promote or suppress each of proceptive and receptive component of sexual behavior have been shown. This multiplicity contributed to confusions in the interpretation of the results of studies which employed focal brain stimulation or lesion. For example, systemic administration of subthreshold dose of estrogen was needed to obtain enhanced lordosis from local implants of estrogen crystals in the VMH [8] or electrical stimulation of the VMH [9]; large amount of estrogen reinstituted lordosis in female rats with VMH lesions [18]. The latter must be due to estrogen-induced removal of inhibitory POA effects. The failure to remove estrogen-dependent POA inhibition might have caused suppressed sexual receptivity in ovariectomized rats with estrogen implants in the VMH [110]. The series of research results presented here indicate that the study of sex differences and sexual differentiation of the brain, which has been focused on morphology, has reached a stage where physiological and molecular biological approaches, such as neuronal channel expression and neurotransmitter dynamics, are required. Further developments in animal study in the field may help enhance understanding of human sexuality.

Availability of data and materials

Extended list of the author’s original publications can be found on online at http://sakuma-physiol.jp/cv06.pdf.

References

  1. Kow LM, Montgomery MO, Pfaff DW (1979) Triggering of lordosis reflex in female rats with somatosensory stimulation: quantitative determination of stimulus parameters. J Neurophysiol 42:195–202

    Article  CAS  PubMed  Google Scholar 

  2. Brennan PA, Kendrick KM (2006) Mammalian social odours: attraction and individual recognition. Philos Trans R Soc Lond B Biol Sci 361:2061–2078

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Xiao K, Kondo Y, Sakuma Y (2004) Sex-specific effects of gonadal steroids on conspecific odor preference in the rat. Horm Behav 46:356–361

    Article  CAS  PubMed  Google Scholar 

  4. Choleris E, Clipperton-Allen AE, Phan A, Kavaliers M (2009) Neuroendocrinology of social information processing in rats and mice. Front Neuroendocrinol 30:442–459

    Article  CAS  PubMed  Google Scholar 

  5. Galef BG Jr, Lim TCW, Gilbert GS (2008) Evidence of mate choice copying in Norway rats. Rattus norvegicus Anim Behav 75:1117–1123

    Article  Google Scholar 

  6. Premoli M, Pietropaolo S, Wohr M, Simola N, Bonini SA (2023) Mouse and rat ultrasonic vocalizations in neuroscience and neuropharmacology: State of the art and future applications. Eur J Neurosci 57:2062–2096

    Article  CAS  PubMed  Google Scholar 

  7. White NR, Barfield RJ (1989) Playback of female rat ultrasonic vocalizations during sexual behavior. Physiol Behav 45:229–233

    Article  CAS  PubMed  Google Scholar 

  8. Barfield RJ, Chen JJ (1977) Activation of estrous behavior in ovariectomized rats by intracerebral implants of estradiol benzoate. Endocrinology 101:1716–1725

    Article  CAS  PubMed  Google Scholar 

  9. Pfaff DW, Sakuma Y (1979) Facilitation of the lordosis reflex of female rats from the ventromedial nucleus of the hypothalamus. JPhysiol (Lond) 288:189–202

    Article  CAS  Google Scholar 

  10. Sakuma Y, Pfaff DW (1979) Facilitation of female reproductive behavior from mesensephalic central gray in the rat. Am J Physiol 237:R278-284

    CAS  PubMed  Google Scholar 

  11. Saper CB, Swanson LW, Cowan WM (1976) The efferent connections of the ventromedial nucleus of the hypothalamus of the rat. J Comp Neurol 169:409–442

    Article  CAS  PubMed  Google Scholar 

  12. Krieger MS, Conrad LC, Pfaff DW (1979) An autoradiographic study of the efferent connections of the ventromedial nucleus of the hypothalamus. J Comp Neurol 183:785–815

    Article  CAS  PubMed  Google Scholar 

  13. Sakuma Y, Pfaff DW (1980) Excitability of female rat central gray cells with medullary projections: changes produced by hypothalamic stimulation and estrogen treatment. J Neurophysiol 44:1012–1023

    Article  CAS  PubMed  Google Scholar 

  14. Morrell JI, Pfaff DW (1982) Characterization of estrogen-concentrating hypothalamic neurons by their axonal projections. Science 217:1273–1276

    Article  CAS  PubMed  Google Scholar 

  15. Kow LM, Pataky S, Dupre C, Phan A, Martin-Alguacil N, Pfaff DW (2016) Analyses of rapid estrogen actions on rat ventromedial hypothalamic neurons. Steroids 111:100–112

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Hasegawa T, Sakuma Y (1990) Estrogen-induced suppression of female rat forebrain neurons with axons to ventral midbrain. Neurosci Lett 119:171–174

    Article  CAS  PubMed  Google Scholar 

  17. Hasegawa T, Sakuma Y (1993) Developmental effect of testosterone on estrogen sensitivity of the rat preoptic neurons with axons to the ventral tegmental area. Brain Res 611:1–6

    Article  CAS  PubMed  Google Scholar 

  18. Pfaff DW, Sakuma Y (1979) Deficit in the lordosis reflex of female rats caused by lesions in the ventromedial nucleus of the hypothalamus. J Physiol (Lond) 288:203–210

    Article  CAS  PubMed  Google Scholar 

  19. Yin L, Hashikawa K, Hashikawa Y, Osakada T, Lischinsky JE, Diaz V, Lin D (2022) VMHvllCckar cells dynamically control female sexual behaviors over the reproductive cycle. Neuron. https://doi.org/10.1016/j.neuron.2022.06.026

    Article  PubMed  PubMed Central  Google Scholar 

  20. Yang CF, Chiang MC, Gray DC, Prabhakaran M, Alvarado M, Juntti SA, Unger EK, Wells JA, Shah NM (2013) Sexually dimorphic neurons in the ventromedial hypothalamus govern mating in both sexes and aggression in males. Cell 153:896–909

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Yang T, Bayless DW, Wei Y, Landayan D, Marcelo IM, Wang Y, DeNardo LA, Luo L, Druckmann S, Shah NM (2023) Hypothalamic neurons that mirror aggression. Cell 186:1195–2121

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Prossnitz ER, Sklar LA, Oprea TI, Arterburn JB (2008) GPR30: a novel therapeutic target in estrogen-related disease. Trends Pharmacol Sci 29:116–123

    Article  CAS  PubMed  Google Scholar 

  23. Bologa CG, Revankar CM, Young SM, Edwards BS, Arterburn JB, Kiselyov AS, Parker MA, Tkachenko SE, Savchuck NP, Sklar LA, Oprea TI, Prossnitz ER (2006) Virtual and biomolecular screening converge on a selective agonist for GPR30. Nat Chem Biol 2:207–212

    Article  CAS  PubMed  Google Scholar 

  24. Anchan D, Gafur A, Sano K, Ogawa S, Vasudevan N (2014) Activation of the GPR30 receptor promotes lordosis in female mice. Neuroendocrinology 100:71–80

    Article  CAS  PubMed  Google Scholar 

  25. Hadjimarkou MM, Vasudevan N (2018) GPER1/GPR30 in the brain: crosstalk with classical estrogen receptors and implications for behavior. J Steroid Biochem Mol Biol 176:57–64

    Article  CAS  PubMed  Google Scholar 

  26. Wada-Kiyama Y, Suzuki C, Hamada T, Rai D, Kiyama R, Kaneda M, Sakuma Y (2013) Estrogen-induced cell signaling in the sexually dimorphic nucleus of the rat preoptic area: potential involvement of cofilin in actin dynamics for cell migration. Biochem Biophys Res Commun 434:287–292

    Article  CAS  PubMed  Google Scholar 

  27. MacLusky NJ, McEwen BS (1978) Oestrogen modulates progestin receptor concentrations in some rat brain regions but not in others. Nature 274:276–278

    Article  CAS  PubMed  Google Scholar 

  28. McCarthy MM, Malik KF, Feder HH (1990) Increased GABAergic transmission in medial hypothalamus facilitates lordosis but has the opposite effect in preoptic area. Brain Res 507:40–44

    Article  CAS  PubMed  Google Scholar 

  29. Pfaff DW (1973) Luteinizing hormone-releasing factor potentiates lordosis behavior in hypophysectomized ovariectomized female rats. Science 182:1148–1149

    Article  CAS  PubMed  Google Scholar 

  30. Sakuma Y, Pfaff DW (1983) Modulation of the lordosis reflex of female rats by LHRH, its antiserum and analogs in the mesencephalic central gray. Neuroendocrinology 36:218–224

    Article  CAS  PubMed  Google Scholar 

  31. Rodriguez-Sierra JF, Komisaruk BR (1977) Effects of prostaglandin E2 and indomethacin on sexual behavior in the female rat. Horm Behav 9:281–289

    Article  CAS  PubMed  Google Scholar 

  32. Watanabe M, Sakuma Y, Kato M (2009) GABAA receptors mediate excitation in adult rat GnRH neurons. Biol Reprod 81:327–332

    Article  CAS  PubMed  Google Scholar 

  33. Mani SK, Blaustein JD (2012) Neural progestin receptors and female sexual behavior. Neuroendocrinology 96:152–161

    Article  CAS  PubMed  Google Scholar 

  34. Brinton RD, Thompson RF, Foy MR, Baudry M, Wang J, Finch CE, Morgan TE, Pike CJ, Mack WJ, Stanczyk FZ, Nilsen J (2008) Progesterone receptors: form and function in brain. Front Neuroendocrinol 29:313–339

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Dohanich GP, Barr PJ, Witcher JA, Clemens LG (1984) Pharmacological and anatomical aspects of cholinergic activation of female sexual behavior. Physiol Behav 32:1021–1026

    Article  CAS  PubMed  Google Scholar 

  36. Foreman MM, Moss RL (1978) Role of hypothalamic alpha and beta adrenergic receptors in the control of lordotic behavior in the ovariectomized-estrogen primed rat. Pharmacol Biochem Behav 9:235–241

    Article  CAS  PubMed  Google Scholar 

  37. Foreman MM, Moss RL (1978) Role of hypothalamic serotonergic receptors in the control of lordosis behavior in the female rat. Horm Behav 10:97–106

    Article  CAS  PubMed  Google Scholar 

  38. Luine VN, Renner KJ, Frankfurt M, Azmitia EC (1984) Facilitated sexual behavior reversed and serotonin restored by raphe nuclei transplanted into denervated hypothalamus. Science 226:1436–1439

    Article  CAS  PubMed  Google Scholar 

  39. Foreman MM, Moss RL (1977) Effects of subcutaneous injection and intrahypothalamic infusion of releasing hormones upon lordotic response to repetitive coital stimulation. Horm Behav 8:219–234

    Article  CAS  PubMed  Google Scholar 

  40. Caldwell JD, Prange AJ Jr, Pedersen CA (1986) Oxytocin facilitates the sexual receptivity of estrogen-treated female rats. Neuropeptides 7:175–189

    Article  CAS  PubMed  Google Scholar 

  41. Arletti R, Bertolini A (1985) Oxytocin stimulates lordosis behavior in female rats. Neuropeptides 6:247–253

    Article  CAS  PubMed  Google Scholar 

  42. Pedersen C (2002) Oxytocin maintains as well as initiates female sexual behavior: effects of a highly selective oxytocin antagonist. Horm Behav 41:170–177

    Article  CAS  PubMed  Google Scholar 

  43. Caldwell JD, Johns JM, Faggin BM, Senger MA, Pedersen CA (1994) Infusion of an oxytocin antagonist into the medial preoptic area prior to progesterone inhibits sexual receptivity and increases rejection in female rats. Horm Behav 28:288–302

    Article  CAS  PubMed  Google Scholar 

  44. Dominguez-Ordonez R, Garcia-Juarez M, Tapia-Hernandez S, Luna-Hernandez A, Galindo-Madrid ME, Tecamachaltzi-Silvaran MB, Hoffman KL, Pfaus JG, Gonzalez-Flores O (2021) Oxytocin induces lordosis behavior in female rats through the prostaglandin E2/GnRH signaling system. Horm Behav. https://doi.org/10.1016/j.yhbeh.2021.105081

    Article  PubMed  Google Scholar 

  45. Barberis C, Tribollet E (1996) Vasopressin and oxytocin receptors in the central nervous system. Crit Rev Neurobiol 10:119–154

    Article  CAS  PubMed  Google Scholar 

  46. Hoshina Y, Takeo T, Nakano K, Sato T, Sakuma Y (1994) Axon-sparing lesion of the preoptic area enhances receptivity and diminishes proceptivity among components of female rat sexual behavior. Behav Brain Res 61:197–204

    Article  CAS  PubMed  Google Scholar 

  47. Kohler C, Schwarcz R (1983) Comparison of ibotenate and kainate neurotoxicity in rat brain: a histological study. Neuroscience 8:819–835

    Article  CAS  PubMed  Google Scholar 

  48. Clough RW, Rodriguez Sierra JF, Hoppe LB (1986) Kainic acid lesioning of the preoptic area alters positive feedback of estrogen in prepubertal female rats. Biol Reprod 35:1269–1276

    Article  CAS  PubMed  Google Scholar 

  49. Hansen S, Kohler C, Goldstein M, Steinbusch HV (1982) Effects of ibotenic acid-induced neuronal degeneration in the medial preoptic area and the lateral hypothalamic area on sexual behavior in the male rat. Brain Res 239:213–232

    Article  CAS  PubMed  Google Scholar 

  50. Olster DH (1993) Ibotenic acid-induced lesions of the medial preoptic area/anterior hypothalamus enhance the display of progesterone-facilitated lordosis in male rats. Brain Res 626:99–105

    Article  CAS  PubMed  Google Scholar 

  51. Yagi K (1973) Changes in firing rates of single preoptic and hypothalamic units following an intravenous administration of estrogen in the castrated female rat. Brain Res 53:343–352

    Article  CAS  PubMed  Google Scholar 

  52. Yamanouchi K, Arai Y (1977) Possible inhibitory role of the dorsal inputs to the preoptic area and hypothalamus in regulating female sexual behavior in the female rat. Brain Res 127:296–301

    Article  CAS  PubMed  Google Scholar 

  53. Lisk RD, MacGregor L (1982) Subproestrus estrogen levels facilitate lordosis following septal or cingulate lesions. Neuroendocrinology 35:313–320

    Article  CAS  PubMed  Google Scholar 

  54. Corodimas KP, Morrell JI (1990) Estradiol-concentrating forebrain and midbrain neurons project directly to the medulla. J Comp Neurol 291:609–620

    Article  CAS  PubMed  Google Scholar 

  55. DonCarlos LL, Monroy E, Morrell JI (1991) Distribution of estrogen receptor-immunoreactive cells in the forebrain of the female guinea pig. J Comp Neurol 305:591–612

    Article  CAS  PubMed  Google Scholar 

  56. Takeo T, Chiba Y, Sakuma Y (1993) Suppression of the lordosis reflex of female rats by efferents of the medial preoptic area. Physiol Behav 53:831–838

    Article  CAS  PubMed  Google Scholar 

  57. Arendash GW, Gorski RA (1983) Suppression of lordotic responsiveness in the female rat during mesencephalic electrical stimulation. Pharmacol Biochem Behav 19:351–357

    Article  CAS  PubMed  Google Scholar 

  58. Sakamoto Y, Suga S, Sakuma Y (1993) Estrogen-sensitive neurons in the female rat ventral tegmental area: a dual route for the hormone action. J Neurophysiol 70:1469–1475

    Article  CAS  PubMed  Google Scholar 

  59. Schwarcz R, Fuxe K, Agnati LF, Hokfelt T, Coyle JT (1979) Rotational behaviour in rats with unilateral striatal kainic acid lesions: a behavioural model for studies on intact dopamine receptors. Brain Res 170:485–495

    Article  CAS  PubMed  Google Scholar 

  60. Watkins JC, Olverman HJ (1987) Agonists and antagonists for excitatory amino acid receptors. Trends Neurosci 10:265–272

    Article  CAS  Google Scholar 

  61. Monaghan DT, Cotman CW (1986) Identification and properties of N-methyl-D-aspartate receptors in rat brain synaptic plasma membranes. Proc Natl Acad Sci USA 83:7532–7536

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. McCarthy MM, Curran GH, Feder HH (1991) Excitatory amino acid modulation of lordosis in the rat. Neurosci Lett 126:94–97

    Article  CAS  PubMed  Google Scholar 

  63. Hasegawa T, Takeo T, Akitsu H, Hoshina Y, Sakuma Y (1991) Interruption of the lordosis reflex of female rats by ventral midbrain stimulation. Physiol Behav 50:1033–1038

    Article  CAS  PubMed  Google Scholar 

  64. Fahrbach SE, Morrell JI, Pfaff DW (1986) Effect of varying the duration of pre-test cage habituation on oxytocin induction of short-latency maternal behavior. Physiol Behav 37:135–139

    Article  CAS  PubMed  Google Scholar 

  65. Dixson AF, Hastings MH (1992) Effects of ibotenic acid-induced neuronal degeneration in the hypothalamus upon proceptivity and sexual receptivity in the female marmoset (Callithrix jacchus). J Neuroendocrinol 4:719–726

    Article  CAS  PubMed  Google Scholar 

  66. Pfeifle JK, Edwards DA (1983) Midbrain lesions eliminate sexual receptivity but spare sexual motivation in female rats. Physiol Behav 31:385–589

    Article  CAS  PubMed  Google Scholar 

  67. Whitney JF (1986) Effect of medial preoptic lesions on sexual behavior of female rats is determined by test situation. Behav Neurosci 100:230–235

    Article  CAS  PubMed  Google Scholar 

  68. Leedy MG (1984) Effects of small medial preoptic lesions on estrous cycles and receptivity in female rats. Psychoneuroendocrinology 9:189–196

    Article  CAS  PubMed  Google Scholar 

  69. Popolow HB, King JC, Gerall AA (1981) Rostral medial preoptic area lesions’ influence on female estrous processes and LHRH distribution. Physiol Behav 27:855–861

    Article  CAS  PubMed  Google Scholar 

  70. Witkin JW, Paden CM, Silverman AJ (1982) The luteinizing hormone-releasing hormone (LHRH) systems in the rat brain. Neuroendocrinology 35:429–438

    Article  CAS  PubMed  Google Scholar 

  71. Orikasa C, Kondo Y, Hayashi S, McEwen BS, Sakuma Y (2002) Sexually dimorphic expression of estrogen receptor b in the anteroventral periventricular nucleus of the rat preoptic area: Implication in luteinizing hormone surge. Proc Natl Acad Sci USA 99:3306–3311

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Hellier V, Brock O, Candlish M, Desroziers E, Aoki M, Mayer C, Piet R, Herbison A, Colledge WH, Prevot V, Boehm U, Bakker J (2018) Female sexual behavior in mice is controlled by kisspeptin neurons. Nat Commun. https://doi.org/10.1038/s41467-017-02797-2

    Article  PubMed  PubMed Central  Google Scholar 

  73. Yamanouchi K, Arai Y (1990) The septum as origin of a lordosis-inhibiting influence in female rats: effect of neural transection. Physiol Behav 48:351–355

    Article  CAS  PubMed  Google Scholar 

  74. Nance DM, Shryne J, Gorski RA (1975) Effects of septal lesions on behavioral sensitivity of female rats to gonadal hormones. Horm Behav 6:59–64

    Article  CAS  PubMed  Google Scholar 

  75. Swanson LW, Cowan WM (1979) The connections of the septal region in the rat. J Comp Neurol 186:621–655

    Article  CAS  PubMed  Google Scholar 

  76. Conrad LC, Pfaff DW (1976) Autoradiographic tracing of nucleus accumbens efferents in the rat. Brain Res 113:589–596

    Article  CAS  PubMed  Google Scholar 

  77. Swanson LW (1976) An autoradiographic study of the efferent connections of the preoptic region in the rat. J Comp Neurol 167:227–256

    Article  CAS  PubMed  Google Scholar 

  78. Kakeyama M, Yamanouchi K (1994) Two types of lordosis-inhibiting systems in male rats: dorsal raphe nucleus lesions and septal cuts. Physiol Behav 56:189–192

    Article  CAS  PubMed  Google Scholar 

  79. Quadagno DM, Shryne J, Anderson C, Gorski RA (1972) Influence of gonadal hormones on social, sexual, emergence, and open field behaviour in the rat (Rattus norvegicus). Anim Behav 20:732–740

    Article  CAS  PubMed  Google Scholar 

  80. Edwards DA, Einhorn LC (1986) Preoptic and midbrain control of sexual motivation. Physiol Behav 37:329–335

    Article  CAS  PubMed  Google Scholar 

  81. Rivas FJ, Mir D (1990) Effects of nucleus accumbens lesion on female rat sexual receptivity and proceptivity in a partner preference paradigm. Behav Brain Res 41:239–249

    Article  CAS  PubMed  Google Scholar 

  82. Spiteri T, Ogawa S, Musatov S, Pfaff DW, Agmo A (2012) The role of the estrogen receptor alpha in the medial preoptic area in sexual incentive motivation, proceptivity and receptivity, anxiety, and wheel running in female rats. Behav Brain Res 230:11–20

    Article  CAS  PubMed  Google Scholar 

  83. Clark AS, Pfeifle JK, Edwards DA (1981) Ventromedial hypothalamic damage and sexual proceptivity in female rats. Physiol Behav 27:597–602

    Article  CAS  PubMed  Google Scholar 

  84. Erskine MS (1989) Solicitation behavior in the estrous female rat: a review. Horm Behav 23:473–502

    Article  CAS  PubMed  Google Scholar 

  85. McClintock MK, Adler NT (1978) The role of the female during copulation in wild and domestic Norway rats (Rattus norvegicus). Behaviour 67:67–96

    Article  Google Scholar 

  86. Garcia-Rill E (1986) The basal ganglia and the locomotor regions. Brain Res 396:47–63

    Article  CAS  PubMed  Google Scholar 

  87. Shik ML, Orlovsky GN (1976) Neurophysiology of locomotor automatism. Physiol Rev 56:465–501

    Article  CAS  PubMed  Google Scholar 

  88. Swanson LW, Mogenson GJ, Gerfen CR, Robinson P (1984) Evidence for a projection from the lateral preoptic area and substantia innominata to the “mesencephalic locomotor region” in the rat. Brain Res 295:161–178

    Article  CAS  PubMed  Google Scholar 

  89. Blizard DA, Lippman HR, Chen JJ (1975) Sex differences in open-field behvior in the rat: the inductive and activational role of gonadal hormones. Physiol Behav 14:601–608

    Article  CAS  PubMed  Google Scholar 

  90. Blizard DA (1983) Sex differences in running-wheel behaviour in the rat: the inductive and activational effects of gonadal hormones. Anim Behav 31:378–384

    Article  CAS  Google Scholar 

  91. Fahrbach SE, Meisel RL, Pfaff DW (1985) Preoptic implants of estradiol increase wheel running but not the open field activity of female rats. Physiol Behav 35:985–992

    Article  CAS  PubMed  Google Scholar 

  92. Morgan MA, Pfaff DW (2001) Effects of estrogen on activity and fear-related behaviors in mice. Horm Behav 40:472–482

    Article  CAS  PubMed  Google Scholar 

  93. Mori S, Matsuyama K, Kohyama J, Kobayashi Y, Takakusaki K (1992) Neuronal constituents of postural and locomotor control systems and their interactions in cats. Brain Dev 14(Suppl):S109–S120

    PubMed  Google Scholar 

  94. Swanson LW, Mogenson GJ, Simerly RB, Wu M (1987) Anatomical and electrophysiological evidence for a projection from the medial preoptic area to the “mesencephalic and subthalamic locomotor regions” in the rat. Brain Res 405:108–122

    Article  CAS  PubMed  Google Scholar 

  95. Sinnamon HM (1992) Microstimulation mapping of the basal forebrain in the anesthetized rat: The “preoptic locomotor region.” Neuroscience 50:197–207

    Article  CAS  PubMed  Google Scholar 

  96. Sinnamon HM (1987) Glutamate and picrotoxin injections into the preoptic basal forebrain initiate locomotion in the anesthetized rat. Brain Res 400:270–277

    Article  CAS  PubMed  Google Scholar 

  97. Brudzynski SM, Eckersdorf B (1984) Inhibition of locomotor activity during cholinergically-induced emotional-aversive response in the cat. Behav Brain Res 14:247–253

    Article  CAS  PubMed  Google Scholar 

  98. Brudzynski SM, Mogenson GJ (1986) Decrease of locomotor activity by injections of carbachol into the anterior hypothalamic/preoptic area of the rat. Brain Res 376:38–46

    Article  CAS  PubMed  Google Scholar 

  99. Takeo T, Sakuma Y (1995) Diametrically opposite effects of estrogen on the excitability of female rat medial and lateral preoptic neurons with axons to the midbrain locomotor region. Neurosci Res 22:73–80

    Article  CAS  PubMed  Google Scholar 

  100. Sakuma Y (1984) Influences of neonatal gonadectomy or androgen exposure on the sexual differentiation of the rat ventromedial hypothalamus. J Physiol 349:273–286

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Suga S, Sakuma Y (1994) Dihydrotestosterone-sensitive neurons in the male rat ventromedial hypothalamus. Brain Res Bull 33:205–210

    Article  CAS  PubMed  Google Scholar 

  102. McEwen BS (1991) Non-genomic and genomic effects of steroids on neural activity. Trends Pharmacol Sci 12:141–147

    Article  CAS  PubMed  Google Scholar 

  103. Simerly RB, Chang C, Muramatsu M, Swanson LW (1990) Distribution of androgen and estrogen receptor mRNA-containing cells in the rat brain: an in situ hybridization study. J Comp Neurol 294:76–95

    Article  CAS  PubMed  Google Scholar 

  104. Rodriguez del Castillo A, Battaner E, Guerra M, Alonso T, Mas M (1987) Regional changes of brain Na+, K+-transporting adenosine triphosphatase related to ovarian function. Brain Re 416:113–118

    Article  CAS  Google Scholar 

  105. Guerra M, Rodriguez Del Castillo A, Battaner E, Mas M (1987) Androgens stimulate preoptic area Na+, K+-ATPase activity in male rats. Neurosci Lett 78:97–100

    Article  CAS  PubMed  Google Scholar 

  106. Stahl WL (1986) The Na, K-ATPase of nervous tissue. Neurochem Int 8:449–476

    Article  CAS  PubMed  Google Scholar 

  107. Pragnell M, Snay KJ, Trimmer JS, Maclusky NJ, Naftolin F, Kaczmarek LK, Boyle MB (1990) Estrogen induction of a small, putative K+ channel messenger RNA in rat uterus. Neuron 4:807–812

    Article  CAS  PubMed  Google Scholar 

  108. Mobbs CV, Kaplitt M, Kow LM, Pfaff DW (1991) PLC-α: A common mediator of the action of estrogen and other hormones? Mol Cell Endocrinol 80:C187–C191

    Article  CAS  PubMed  Google Scholar 

  109. Nishimura I, Ui-Tei K, Saigo K, Ishii H, Sakuma Y, Kato M (2008) 17β-Estradiol at physiological concentrations augments Ca2+-activated K+ currents via estrogen receptor β in the gonadotropin-releasing hormone neuronal cell line GT1-7. Endocrinology 149:774–782

    Article  CAS  PubMed  Google Scholar 

  110. Bast JD, Hunts C, Renner KJ, Morris RK, Quadagno DM (1987) Lesions in the preoptic area suppressed sexual receptivity in ovariectomized rats with estrogen implants in the ventromedial hypothalamus. Brain Res Bul 18:153–158

    Article  CAS  Google Scholar 

  111. Kato A, Sakuma Y (2000) Neuronal activity in female rat preoptic area associated with sexually motivated behavior. Brain Res 862:90–102

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Author's original studies cited in this review were supported by: JSPS Grants-in-aid for Scientific Research; MEXT Grants for Scientific Research on Priority Areas; Strategic Grants to Promote Research at Private Universities, MEXT; Match-funding by Nippon Medical School Foundation. The author thanks the regents of the Physiological Society of Japan for providing him this opportunity to compile his works on the occasion to commemorate 100th anniversary of the Society

Funding

JSPS Grants-in-aid for Scientific Research (18390070; 18659061; 22390043 &c); MEXT Grants for Scientific Research on Priority Area “Mechanism of Sexual Differentiation” (1686210); Strategic Grant to Promote Research at Private Universities, MEXT (S0801035); Matching funds by Nippon Medical School Foundation &c Details are provided at https://bit.ly/2Odi9bs

Author information

Authors and Affiliations

Authors

Contributions

Author's original works referred to in this review were designed, carried out, analyzed and wrote by him.

Corresponding author

Correspondence to Yasuo Sakuma.

Ethics declarations

Ethics approval and consent to participate

The author's original works quoted in this review were carried out under approvals by institutional animal research ethics committees of universities at Gunma, Niigata, Hirosaki, and Nippon Medical School.

Consent for publication

The author is the sole agent of this manuscript.

Competing interests

The author has no competing interests to disclose.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sakuma, Y. Preoptic and hypothalamic regulation of multi-tiered, chronologically arranged female rat sexual behavior. J Physiol Sci 73, 35 (2023). https://doi.org/10.1186/s12576-023-00890-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12576-023-00890-4