Next Article in Journal
Degradation Mechanism of 2,4-Dichlorophenol by Fungi Isolated from Marine Invertebrates
Next Article in Special Issue
Progesterone Attenuates Stress-Induced NLRP3 Inflammasome Activation and Enhances Autophagy Following Ischemic Brain Injury
Previous Article in Journal
Coloured Rice Phenolic Extracts Increase Expression of Genes Associated with Insulin Secretion in Rat Pancreatic Insulinoma β-cells
Previous Article in Special Issue
Microglial and Astrocytic Function in Physiological and Pathological Conditions: Estrogenic Modulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

De Novo Synthesized Estradiol: A Role in Modulating the Cerebellar Function

by
Cristina V. Dieni
1,*,
Samuele Contemori
2,
Andrea Biscarini
3 and
Roberto Panichi
3,*
1
Department of Ophthalmology, University of Alabama at Birmingham, Birmingham, AL 35294, USA
2
Centre for Sensorimotor Performance, School of Human Movement and Nutrition Sciences, The University of Queensland, Brisbane 4072, Australia
3
Department of Experimental Medicine, Section of Physiology and Biochemistry, University of Perugia, 06129 Perugia, Italy
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(9), 3316; https://doi.org/10.3390/ijms21093316
Submission received: 23 March 2020 / Revised: 26 April 2020 / Accepted: 5 May 2020 / Published: 7 May 2020
(This article belongs to the Special Issue Steroids and the Nervous System)

Abstract

:
The estrogen estradiol is a potent neuroactive steroid that may regulate brain structure and function. Although the effects of estradiol have been historically associated with gonadal secretion, the discovery that this steroid may be synthesized within the brain has expanded this traditional concept. Indeed, it is accepted that de novo synthesized estradiol in the nervous system (nE2) may modulate several aspects of neuronal physiology, including synaptic transmission and plasticity, thereby influencing a variety of behaviors. These modulations may be on a time scale of minutes via non-classical and often membrane-initiated mechanisms or hours and days by classical actions on gene transcription. Besides the high level, recent investigations in the cerebellum indicate that even a low aromatase expression can be related to the fast nE2 effect on brain functioning. These pieces of evidence point to the importance of an on-demand and localized nE2 synthesis to rapidly contribute to regulating the synaptic transmission. This review is geared at exploring a new scenario for the impact of estradiol on brain processes as it emerges from the nE2 action on cerebellar neurotransmission and cerebellum-dependent learning.

1. Introduction

Estrogens are part of the neuroactive steroid family that may regulate the structure and function of neural networks via multiple modes and time courses. The most potent estrogen in influencing brain functions is the 17 beta-estradiol that exerts its effect via both classical long term actions on genomic mechanisms and rapid non-classical effects [1,2,3,4,5,6,7]. The estradiol impacts on neural physiology depend on its bioavailability in a brain structure, and it has been shown that, in addition to peripheral synthesis such as in the gonads, the estradiol can be locally produced in the nervous system [8]. The de novo synthesized 17 beta-estradiol in nervous tissues, defined as neurosteroid (nE2), has the same structure and mechanisms of synthesis than the gonadal derived estrogen (Figure 1).
The production of nE2 requires an aromatase-dependent conversion of testosterone, which may be either of the peripheral origins or locally synthesized from the precursor cholesterol [9,10,11,12]. Thus, estradiol is no longer merely considered a hormone produced by the ovaries and only related to the control of female sexual maturation and reproduction. Instead, it is now known to have multiple homeostatic roles, and in the nervous system, estradiol controls a variety of processes in males as well as females [2,13,14,15,16]. Via the “classical” genomic mode of action, estradiol interacts with intracellular receptors to influence transcriptional pathways and regulate DNA transcription within minutes (Figure 2) [17].
Two typical estrogen receptor isoforms (ERs: ERα, ERβ) are known to participate in classical influence. However, to produce detectable functional effects on the cellular level or the entire organism, the “classical” mode of action requires post-transcriptional events that take hours to days to occur [7,18,19]. In the brain, faster non-classical and often membrane-initiated mechanisms are observable within a few seconds (for some electrophysiological effects) to minutes after estradiol activated signaling involving intracellular cascades that may include several protein kinases and result in cytoplasmatic fluctuations of calcium concentration (Figure 2) [19,20,21,22,23,24]. Some of these intracellular signalings could also affect the gene transcription through the so-called “indirect genetic effect” [25]. However, the essential point considered here is that estradiol may rapidly modulate pathways associated with the regulation of neurotransmission and neuronal activity such as those related to cytoplasmatic protein kinases cascades and calcium fluctuation (Figure 2) [23,26].
Rapid membrane-initiated effects involve estrogen G-protein-coupled receptors (GPCRs: GPER-1 and Gαq-mER), classical estrogen receptors (ERα, ERβ), and probably other putative receptors (e.g., ER-X) [27,28,29,30,31,32,33,34,35]. It is noteworthy to mention that non-classical estrogen actions on neural circuits are not limited to those that occur via the activation of membrane-associated estradiol pathways, but also include mechanisms that can depend on the estrogen binding of cytoplasmatic receptors with the consequent activation of various intracellular signaling cascades (Figure 2) [36].
It has been established that nE2 may be a potent regulator of the neuronal functions when produced in large amounts [4,37]. However, how generalized nE2 synthesis is within the nervous system and whether some of the nE2 effects can be even mediated by low estradiol production are issues of rising interest in neurophysiology and neuroendocrinology.
Recent evidence showed that besides the high level of synthesis, even alleged minute and localized nE2 production might contribute to rapidly modulating aspects of neuronal physiology like the synaptic transmission and plasticity (Figure 2).
Experiments conducted in the brainstem and cerebellum indicate that despite the supposed low synthesis in the experimental models used, nE2 may contribute to rapidly modulating neurotransmission via dynamic change of its availability at a synaptic level [38,39,40,41,42]. It has already been demonstrated that modifications in local estradiol synthesis may be achieved by changes in aromatase activity. Although possible changes in aromatase activity result from transcriptional control that alters the enzyme concentration, these changes generally occur slowly in a time scale of hours or days. Alternatively, much more rapid changes in the rate of nE2 production may result from post-translational events that potentiate/inhibit the enzyme activity (Figure 3) [38].
Post-translational modifications are common mechanisms involved in the control of protein activity in the brain, like in the case of the regulation of neurotransmitter activity in neurons [43].
The evidence introduced above provides new insight into the effectiveness of nE2 in regulating neuronal activity and indicating that estradiol may play an essential role in the modulation of network functioning, even in brain structures in which, so far, its synthesis has been assumed to be too low to be effective. Here, we review the emerging view from this scenario that expands the traditional significance of nE2 production into a new physiological role and includes recently suggested mechanisms involved in the rapid control of the neural transmission, with particular regard to the synaptic transmission in the cerebellum.

2. The Effectiveness of the De Novo Synthesized Estradiol in Rapidly Influencing the Neuronal Functioning

It has long been assumed that the impact of nE2 on the regulation of neural functions should be restricted at those brain areas in which the estradiol synthesizing enzyme aromatase is highly expressed and, of course, estrogen receptors are present. According to this assumption, in such areas, nE2 may rapidly reach high concentrations [4,44,45], bind membrane ERs, or activate other non-classical mechanisms to initiate intracellular signaling and acutely influence the neuronal physiology [13,21,33,36,46]. Rapid activation of membrane estrogen pathways by nE2 may be facilitated by the co-localization of aromatase and estrogen receptors in the plasma membrane or their localization at the synaptic level in neurons of the central nervous system [47,48,49]. Non-classical influences include, but are not limited to, the modulation of synaptic activity, synaptogenesis, and spinogenesis (Figure 2) [23,50,51]. Rapid functional regulation of synaptic activity by estradiol has been found, so far, for the glutamatergic, GABAergic, cholinergic, and dopaminergic systems [27,41,42,52,53,54,55,56,57]. Of course, also classical genomic delayed effects influencing neuronal structure and function can be mediated by nE2 and interact with the non-classical mode of actions [2,13,19]. nE2 may, therefore, regulate over different timescale phenomena such as synaptic transmission, spine density, synaptic connectivity, synaptogenesis, and neurogenesis, that mediate behavioral changes (Figure 2) [13,20,21,27,58,59].
In many vertebrates, brain areas with the highest aromatase expression include hypothalamic preoptic nuclei, the bed nucleus of the stria terminalis, and medial amygdala, typically involved in the control of stress responses, sexual and social behaviors [2,60,61,62] and other structures that, like in the case of the hippocampus and inferior olive, are related to sensorimotor aspects of information processing and learning phenomena [44,63,64,65,66]. Consistently, besides its role in the regulation of reproduction and sexual functions [1,2,67,68,69], nE2 may influence a variety of other processes and behaviors from the mood to learning and memory formation and have a neuroprotective role as well [14,27,50,52,70,71,72,73,74,75,76,77,78].
However, despite a significant interspecific variability, low aromatase expression may be present in other brain areas such as temporal, occipital, and cingulate cortex, putamen, brainstem, and cerebellum of adult animals including rodents and humans [66,79,80,81,82].
Rodents are the most used model to study brain functions. For instance, in the cerebellum of rodents, aromatase is highly expressed only during neonatal life while it is maintained at a constant low level during peripubertal (the period between day 30 and day 45 after birth) and adulthood (post-pubertal) [66,83,84]. Thus, while the role of nE2 in regulating cerebellar development has been investigated [85,86], its effects upon the functioning of the cerebellum in adults (in the following text, the term “adult” refers to a period comprehensive the peri and post-pubertal period) have long been considered irrelevant or ignored.
Nevertheless, the age-related maintenance of a low level of aromatase expression in the cerebellum seems to have a cardinal physiological significance in the contribution of controlling the cerebellar function. Recent reports showed that nE2 synthesis rapidly influences the glutamatergic transmission between the parallel fibers and Purkinje cells in the cerebellar cortex and acutely impact cerebellar dependent behaviors in adult rodents [39,40,74,87]. Furthermore, the high level of estrogen receptors found in the cerebellum of adults is supportive of cerebellar responsiveness even to a minute and localized nE2 synthesis [88,89,90]. Table 1 summarizes the main findings showing the aromatase and estrogen receptors distribution into the cerebellar cortex in adult as well as neonatal rodents.
The above-reported findings are consistent with previous discoveries indicating an acute modulation of nE2 on glutamatergic synapses in vestibular nuclei, which are assimilable to extra cerebellar nuclei and have, like the cerebellum, a low level of aromatase expression [41,42,66,91,92,93].
Thus, even at low expression, aromatase may play a functional role in allowing the brain to exert dynamic control over local neuronal functions via nE2 synthesis, as may be the case of synaptic transmission and plasticity in the cerebellar circuitry.

3. Significance of the nE2 Influence upon Cerebellar Functioning

A relatively simple model that offers the unique opportunity to study neuronal aspects responsible for cerebellar-dependent learning is provided by the adaptation of the vestibulo-ocular reflex (VOR), a gaze stabilizing reflex which makes possible clear vision during head movements. VOR adaptation is mediated by combining visuo-vestibular stimulation, and it is an essential process to maintain calibrate and effective the reflex [106,107,108,109]. Furthermore, adaptive increases (gain up) and decreases (gain down) in the magnitude of eye movements in VOR can be experimentally induced by coupling specific visual-vestibular stimuli [74,110,111]. Most important, the VOR model provides the advantage to examine a motor behavior that strictly depends on a well-known circuitry in which the nE2 production can be deficient but localized over the synaptic system essential for the cerebellar learning [62,112,113,114,115,116]. In other words, the VOR model is an ideal arena to study the efficacy of nE2 in modulating neurotransmission and related behavior.
The VOR is mediated by a direct neural pathway through the vestibular nuclei in the brainstem and an additional more complex loop within the flocculus and paraflocculus of the cerebellum. The loop includes glutamatergic axons from granular cells, known as parallel fibers, that synapse onto Purkinje cells (PCs) in the cerebellar cortex, and inhibitory GABAergic projections from PCs to the direct vestibular pathway [113,117,118,119,120] (Figure 4).
Therefore, increases in the efficacy of inhibitory projections from PCs would result in a corresponding decrease in the amplitude of the downstream signal transmitted through the direct pathway controlling the VOR. In contrast, reductions in inhibitory projections would result in a corresponding increase in the downstream signal (Figure 4). Consistently, experimental evidence indicates that the leading candidate mechanisms associated with memory encoding in VOR adaptation are the long-term plasticity (potentiation and depression) at the parallel fiber-PC synapses (Figure 4) [112,121,122,123]. Other synaptic plasticity sites that may play a late role in VOR memory consolidation could be present in the direct pathway in vestibular nuclei (Figure 4) [117,118,124,125]. Notably, as mentioned above, it has recently been shown that in adult rodents, nE2, as well as a high plasma concentration of estrogen, may similarly modulate the neuronal transmission at specific synapses into the VOR network as it is the case of the synapse between parallel fibers and PCs [39,40,41,42,87,91,97,126]. Moreover, the evidence that nE2 synthesis may also rapidly regulate the encoding of VOR adaptation and some other motor behavior controlled by the cerebellum is consistent with a nE2 modulation upon the cerebellar network [39,40,74,122].
The finding that cerebellar neural pathways are responsive to nE2 in adult rodents has relevance for the potential impact that the de novo estradiol syntheses may exert on the control of multiple aspects of cerebellar functioning after the neonatal period. Indeed, considering the homogeneity of structures and synaptic organization throughout different cerebellar regions that contribute to modulate various sensory-motor aspects related to movement and even non-motor phenomena [120,127,128,129], common nE2 modulatory mechanisms could be shared across the cerebellum to influence a variety of processes and functions. For instance, the effects of estradiol observed on the regulation of cerebellar neurotransmission and related to the modulation of learning in the VOR [40,97] were found in vermis, a cerebellar structure external to the circuitry mediating the vestibulo-ocular reflex [120,122].
Understandings the mechanisms underlying the nE2 action on cerebellum in rodent models could have an impact on human health, offering potential insights into the possibility of developing innovative intervention strategies in diseases-involving cerebellar abnormalities [130,131,132,133,134] and supporting neuroprotection [135,136]. The similarity of the estradiol pathway in the cerebellum and brainstem between rodents and humans [66,79,80], indeed, suggests similar action mechanisms of nE2 upon cerebellar networks in these species. Furthermore, it is reasonable to assume that a role of nE2 in controlling neuronal functioning via modulation of synaptic transmission, similar to that detected in the cerebellum, could also be extended to other brain regions where so far, due to low aromatase expression, estradiol synthesis has been not identified or considered not functionally relevant.

4. The Impact of nE2 upon Cerebellar Functioning in Adult Rodents: Multiple Lines of Evidence

The discovery that cerebellar PCs are one of the main sites of neurosteroidogenesis in the brain of neonate mammals has provided the opportunity to investigate the impact of nE2 in regulating brain neural networks develop during neonatal life [85,137]. One of the best-known effects of nE2 in the developing cerebellum includes the modulation of dendritic growth, spinogenesis and, synaptogenesis via activation of estrogen cognate nuclear receptor (mostly ERβ) in PCs. Such nE2 effect is a “classical” delayed genomic action mediated by the neurotrophic factor BDNF and involving Purkinje and external granular cells, both expressing a high level of aromatase [84,85,86,114,137].
Notably, in the cerebellum of adult rodents, the expression and distribution of the enzymatic machinery to produce estradiol, and the distribution of estrogen receptors drastically change compared to a neonatal animal (Table 1). As discussed above, aromatase expression becomes very low, with the enzyme substantially restricted to PCs [66,83,84]. ERαs are expressed at a negligible level and primarily localized within the PCs nucleus and cytoplasm [88,89,100]. The expression of ERβs is kept high with the receptors present in the soma (cytoplasm and nucleus) and dendrites of Purkinje cells and mostly in the cytoplasm of granular, Golgi, and basket type neurons [88,89,90,97,101,105]. GPER-1 is also present in the cerebellum with high, moderate, and low expression in PCs, granular cell layer, and molecular layer, respectively. However, its potential role in cerebellar functioning is almost unexplored [29]. To our knowledge, no information is available on the presence in the adults’ rodent’s cerebellum of estradiol membrane and the G-protein-coupled receptors ER-X and Gαq-mER, respectively. However, may the reorganization of the estradiol pathway into the cerebellar cortex during adulthood still indicate a functional relevance for nE2 in controlling neuronal functions? Very likely, the answer to this question is yes. This type of nE2 pathway reorganization in adults suggests that estradiol may be locally synthesized and potentially exert influence, at least, at the level of synaptic transmission as it could be the case for the synapses converging onto PCs. Furthermore, based on ERs distribution, this influence should possibly be mediated via PCs, granular, and Golgi cell ERβs or PCs-granular cells GPER-1 activation. Although the studies of the impact of nE2 in regulating cerebellar neuronal activity are only at the beginning, some general rules describing the nE2 action on the cerebellum are emerging.
The works of Smith and colleagues first demonstrated the involvement of estrogen in influencing the glutamatergic neurotransmission between parallel fibers and PCs by applying estradiol directly to the cerebellum or administering it via subcutaneous injection in the adult ovariectomized rat [138]. Further, these studies showed that locomotor-induced Purkinje cell activity could be affected by either administration of estradiol or high level of circulating estrogen during proestrus [126,139]. Afterward, Andreescu et al., showed that endogenously induced high level of plasma estradiol in ovariectomized mice might affect long-term potentiation at the parallel fiber to Purkinje cell synapse without impacting the long-tern depression and basal neurotransmission. Furthermore, they found that the high level of plasma estradiol increased the size of the postsynaptic complex at the parallel fiber to PC synapse and regulated VOR adaptation through activation of ERβs signaling in PCs [97]. Interestingly, the authors did not observe the modulatory effects of estradiol on the synapses between climbing fiber and PCs. Besides, they showed differences in VOR adaptation during the natural estrous cycle. Thus, the authors pointed to the importance of the gonadal-dependent plasma estradiol fluctuation in modulating cerebellar functioning [97].
However, the initial concept that the cerebellum might be a mere target of circulating hormones has been recently further developed to extend the action of estrogen into the adult cerebellum to the de novo synthesized estradiol . Hedges et al. demonstrated that nE2, as well as gonad-derived estradiol (at high plasma level), might regulate parallel fiber-Purkinje cell synapses transmission by enhancing the glutamatergic signaling in the adult of both male and female mouse [39,87]. Using in vivo imaging and in vitro electrophysiological techniques, they found that the effect of estradiol on glutamatergic neurotransmission was mainly mediated by an estrogen-dependent activation of metabotropic glutamate receptors type 1a (mGluR1a) signaling into the cerebellar cortex [39], a result supported by the already proposed role of metabotropic glutamate receptors (mGluR1a, 2, and 3) in mediating the estradiol effect on functional change (Figure 2) [140,141]. Compelling, mGluR1s is a leading candidate to play a cardinal role in the cerebellar mechanism involved in the regulation of parallel fiber to PCs synapses and in the adaptation of the VOR [26,112]. Besides, it has been shown that the estradiol-mediated modulation of glutamatergic neurotransmission took part at a postsynaptic level and that the GPER signaling was not involved. Further, male mice underwent chronic and systemic administration of the aromatase inhibitor fadrozole showed an impairment of locomotor performance that indirectly indicates a possible role played by nE2 in controlling this cerebellar-influenced motor behavior [39].
Dieni et al., have found that de novo synthesized estradiol acutely regulates cerebellar synaptic plasticity affecting long term potentiation at the parallel fiber-Purkinje cell synapse (PF-LTP) and modulates the encoding of VOR adaptation in adult male rats [40,74]. In these in vitro studies on cerebellar slices, using the aromatase blocker letrozole (LTZ), the authors demonstrated that nE2 specifically affects the PF-LTP acting at postsynaptic level but influences neither the long term depression at the parallel fiber-Purkinje cell synapse (PF-LTD) nor the intrinsic and basal synaptic proprieties of PCs. Further, nE2 seems not to affect the synaptic transmission between climbing fibers and PCs (data not published), a result supported by the previous finding that estradiol does not alter the structure of climbing fibers to PCs synapses [97].
The authors hypothesized that the nE2 effect on cerebellar LTP could be due to either a transient or an acute increase of estradiol synthesis during PF-LTP induction (Figure 4). Indeed, the acute application of the aromatase inhibitor LTZ in cerebellar slices did not influence basal neuronal activity [40] but, the perfusion of estradiol changes the efficacy of glutamatergic transmission at the parallel fiber to PC synapse [39,87].
Furthermore, these effects were probably mainly mediated by the activation of ERβs (due to the much higher ERβs expression than ERαs), as suggested by the fact that estradiol-dependent effects vanished when electrophysiological recordings were performed in the presence of the estrogen receptors antagonist ICI 182,780 (data not published).
Overall, nE2 does not seems essential for basal neurotransmission. Nevertheless, it is reasonable to assume that in the cerebellum estradiol availability may locally and dynamically fluctuate on demand in response to specific signals and that this dynamic nE2 fluctuation may influence in turn basal neurotransmission at parallel fiber-PC synapse if required and affects synaptic plasticity at the same synaptic site [39,40,87]. The fact that in adult rodents was detected not differences in estradiol concentration between plasma and cerebellum tissues in some studies (concentration on average referred in the order of a hundred picomolar) and considerable diversity in estradiol level in cerebellar explants among others may support the assumption of an on-demand fluctuation of nE2 concentration [39,62,84,100,142]. On the other hand, it is well known that phosphorylation processes and glutamatergic, dopaminergic, and calcium signaling can rapidly modify aromatase activity, thereby causing changes in the capability to synthesize nE2 (Figure 3) [3,21,38,143,144,145,146]. Furthermore, activation of motor pathways can locally elevate aromatase activity, and the nE2 concentration may rapidly fluctuate within sensorimotor circuits [21,53,147]. It is worth mentioning that neurotransmitter signaling modulating aromatase activity, and intracellular calcium transient takes part in cerebellar PCs during motor learning phenomena [26,113,115,122]. Thus, potentially, a variety of signals could rapidly and dynamically modulate nE2 synthesis to regulate the estradiol action at the parallel fiber-PC synapse. This scenario implies that classical neurotransmitters and nE2 signaling may influence each other with the consequence that the balance of this cross-influence could be a requisite for the homeostatic control of the synaptic function (Figure 2 and Figure 3).
However, the essential point is that in the in vitro models, the nE2-dependent modulation of parallel fiber-PCs synapses activity is characterized by a low aromatase expression and a localized estradiol synthesis over PCs which modulation is cardinal for the cerebellar dependent learning [39,40,66,84]. Of course, in intact animals, the circulating estradiol, especially when at high plasma level (e.g., during female proestrus), could also contribute to the nE2 effect or interact with it to determine more complex multiple modes and time-course changes in cerebellar network functioning [40,97]. Besides, to date, it has not been possible to rule out the possibility that in physiological conditions, other nearby structures presenting high nE2 synthesis levels, such as inferior olive and hypothalamus, might supply an amount of nE2 to the cerebellum via penetration-diffusion mechanisms [44,83]. The recent discovery of the pineal gland as an active site of the de novo synthesized neurosteroids in birds also adds this endocrine organ closely located to the cerebellum as a potential structure that could contribute to providing estradiol to the cerebellum [148,149].
Consistent with the impact of nE2 on the plasticity of the cerebellar circuitry, it has also been shown that the inhibition of estradiol synthesis by acute and systemic administration of LTZ affects gain increase and decrease in VOR adaptation without influencing basal oculomotor reflexes of adult male rats [40,74] (Figure 4). This finding provides one of the most persuasive evidence indicating a role played by nE2 in regulating adaptive behavior depending on cerebellar plasticity, and diverse observations support it. First, differently than other motor behavior, the VOR depends on a specific cerebellar circuitry and plasticity (Figure 4). Second, cerebellar and vestibular plasticity mediating the VOR adaptation is selectively regulated by nE2 (see below) [40,41,42,92]. Then, in adult male rodents, the blood level of estradiol (often referred to less than a hundred picomolar) is too low to influence, per se, neuronal plasticity and excluding, therefore, a useful contribution of the circulating hormone in regulating cerebellar function for this experimental model [4,44,62,142,150,151].
Moreover, systemically blocking the estradiol synthesis by LTZ produces a rapid impairment in the adaptation of the VOR, which is not consistent with and has a different timescale of the LTZ effect on the depletion of peripherally derived estrogen [74,152]. Further, LTZ, per se, influence neither basal ocular reflexes nor the basal activity of VOR circuitry [40,42,74]. Finally, cerebellar tissue from castrated rodents contains a similar amount of estrogen than intact animals (intact: ~6 pg/mg; castrated: ~9 pg/mg) [39], suggesting an active estradiol synthesis by the cerebellum and possibly nearby regions.
A critical lack for a better understanding of the estradiol action on cerebellar function is the absence of information about a possible effect of nE2 on the glutamatergic neurotransmission within inferior olive, a crucial structure in influencing the activity of cerebellum (Figure 4) and presenting high nE2 synthesis and estradiol receptors (GPR-1 and ERβs) [29,70,89]. The inferior olive computes both GABAergic and glutamatergic signals, and the balance between these two neurotransmitters signalings is an essential factor for the proposed role of the olivocerebellar system in controlling motor coordination [153,154,155,156]. So far, it has been found that endogenous estradiol, as well as an endogenously induced high level of plasma estradiol, exerts a neuroprotective role against a form of cerebellar ataxia experimentally induced through the degeneration of the inferior olivary nucleus in rats [70]. However, to the best of our knowledge, no studies investigated the nE2 influence on synaptic transmission in the inferior olivary nucleus, although exogenous estradiol could affect the pattern discharge in neurons of dorsal accessory olive [157]. Thus, behind the neuroprotective function, the extent to which the estradiol affects the neuronal activity in inferior olive is an unanswered question. Potentially, nE2 might influence cerebellar-dependent motor behaviors also via modulation of neurotransmission within the inferior olivary nucleus.
Estrogens could also contribute to regulating the cerebellar cortex circuitry via modulation of inhibitory GABAergic signaling [158], similar to what happens in other brain regions [53,159]. Due to structural and functional relationships with the cerebellum and the very similar nE2 pathway, vestibular nuclei may be a good model for inferring a potential action of estrogen on GABAergic/glutamatergic neurotransmission balance in the cerebellum. In medial vestibular nuclei slices, estradiol may potentiate the glutamatergic neurotransmission via ERs (ERα and ERβ) activation and induction of intracellular pre and postsynaptic NMDARs-dependent signaling. While a high concentration of exogenous-derived estradiol affects basal neurotransmission, the locally synthesized nE2 plays an essential role only in the LTP induction [42,92]. However, nE2 may regulate the LTP in vestibular nuclei by diminishing the GABAergic neurotransmission as well as enhancing glutamatergic neurotransmission [41]. An action of nE2 on both cerebellar GABAergic and glutamatergic neurotransmission similar to that found in vestibular nuclei is possible and consistent with the nE2-mediated effects discussed in the next section.

4.1. Implications for the Impact of nE2 on the Cerebellar Function

The general implication of the findings discussed here is that the cerebellum in adult rodents uses de novo estradiol syntheses to potentiate glutamatergic signaling and regulate cortical synaptic plasticity and related learning phenomena.
In terms of impact on synaptic plasticity, nE2 seems to selectively affect the glutamatergic synapse at parallel fiber-PCs acting on PF-LTP without influencing the PF-LTD. Similarly, nE2 plays a central role in the modulation of the LTP but not the LTD in vestibular nuclei, hippocampus, and striatum [41,54,55,56,92]. The selective effect of nE2 on PF-LTP in the cerebellum is consistent with the unique mechanisms leading to long-term synaptic plasticity at this synaptic site. The free intracellular calcium concentration, along with phosphorylation/dephosphorylation processes within PCs, is the critical factor in determining the polarity of synaptic plasticity [26]. nE2 signaling could facilitate the rapid and moderate increase of intracellular calcium concentration associated with the LTP induction in PCs and consequent to parallel fiber stimulation [22,115,160,161]. Conversely, PF-LTD needs more massive and prolonged cytoplasmatic calcium concentration fluctuation than those requested for the LTP, and it can be produced, per se, through the recruitment of climbing fibers used paired with parallel fiber stimulation for the PF-LTD induction [115,162,163,164]. On the other hand, the calcium-dependent phosphatase/kinase switch determining the polarity of synaptic plasticity may activate the aromatase during PF-LTP by dephosphorylation processes and inactivate it in the case of PF-LTD via phosphorylation (Figure 3) [26,38,76,112,165,166].
Thus, it has been suggested that nE2 fluctuation mediates by aromatase activation/deactivation not only facilitates the LTP but also could prevent that inappropriate calcium increments would switch the LTP in LTD [40].
In the schema proposed here, the brain may use localized nE2 synthesis, independently by the peripheral production of estradiol, to rapidly and selectively regulate cerebellar synaptic plasticity and influence related learning processes that are essential to adapt behavior to changing contexts.
For the behavior mainly considered in this discussion, the VOR (Figure 4), it is mostly accepted that the encoding of gain decrease and increase in the adaptation of this reflex depends on PF-LTP and PF-LTD, respectively [110,112,117,121,122].
However, the observation that nE2 influences both the gain increase and decrease in VOR adaptation but at a cellular level only impacts the PF-LTP suggests that extra-cerebellar, possibly in the vestibular nuclei of the brainstem, as well as cerebellar synaptic plastic sites sensitive to nE2 (Figure 4), should be required for this reflex to adapt [40,41,113,116,117].
Consistent with the “transferred hypothesis” [118,167,168,169], motor memory may be first encoded into the cerebellar cortex and then immediately transferred to vestibular nuclei in a process detectable within minutes after the beginning of the learning process [170]. In this scenario, cerebellar LTP and vestibular LTD would mediate the expression of the VOR adaptive increase, while cerebellar LTD and vestibular LTP might regulate the VOR increase (Figure 4) [40,113,117]. Thus, nE2 may take part in the VOR gain decrease, mainly influencing cerebellar PF-LTP, whereas the gain increase of the reflex would be modulated via nE2-dependent regulation of vestibular LTP (Figure 4) [40,74]. Consistently, nE2 is necessary for LTP induction not only in the cerebellum but even in the vestibular nuclei via induction of potentiation in glutamatergic and inhibition in GABAergic signaling [41,42,91]. Immunohistochemical analysis has already demonstrated the presence of aromatase in the vestibular nuclei of rodents during developing and post-developing periods [91,93]. Furthermore, in adult rats, quantitative real-time PCR measures found similar low aromatase mRNA levels in the brainstem and cerebellum [66]. Similarly, only a few aromatase-immunoreactive neuronal structures were observed in the vestibular system of adult male quail [171]. In the cerebellum and vestibular nuclei, therefore, the maintenance of low aromatase expression during adulthood could have the physiological significance to control, via localized nE2 synthesis, the neurotransmission at a specific synaptic site with particular relevance as those controlling adaptive processes.
Since the aromatase in the cerebellum and brainstem in rodents has very similar expression to that detected in humans, one can also hypothesize similar vestibulo-cerebellar mechanisms controlled by nE2 and leading to motor learning in these two species [66,79,80,83]. A cardinal consequence that extends this concept is that nE2 could influence human perceptive and motor phenomena that require adaptive process and plasticity in the vestibulo-cerebellar networks in physiological as well as pathological conditions [172,173,174,175,176]. Thus, nE2 may exert its influence not only on VOR adaptation but even on the rebalancing after labyrinthine lesion of vestibular reflexes and perception [177,178,179] and on other motor behaviors like balance and locomotion requiring the computational power of cerebellar networks to adapt responses to the context [120,128,180,181]. For instance, the degradation of postural stability observed with age could involve the impairment of the nE2 pathway in the vestibulo-cerebellar circuitry or its interaction with circulating estrogen, as suggested by the improvement of balance in post-menopausal women treated with estrogen [182]. Furthermore, the decrease of locomotory performance in mice after the administration of fadrozole could be a consequence, at least in part, of its effects in blocking the aromatase activity in PCs [39]. Notably, the treatment of breast cancer with aromatase inhibitors has beneficial for the course of the disease; however, side effects might be ataxia, a form of motor coordination impairment typically named dysmetria, and dizziness [183,184]. Nevertheless, it is not possible to rule out that after the administration of aromatase inhibitors, the inhibition of nE2 synthesis in extra-cerebellar structures could contribute to impair the motor function and perception.
Unanswered questions over some aspects related to the cerebellar mechanisms underlying changes in motor performance and behavior could be clarified using the manipulation of estradiol synthesis as a sort of investigation tool. For example, although the synaptic plasticity is known as the primary mechanism on cerebellar dependent motor learning, recent studies on VOR adaptation suggest that also intrinsic neural plasticity may play an essential role in it [117]. However, the relationship between synaptic and intrinsic plasticity in the cerebellum has not been established yet. A possible approach to clarify the latter point could be through the use of estrogens or blockers of their synthesis as it has been shown that estradiol may impact the cerebellar synaptic transmission but not the intrinsic proprieties of neurons in the cerebellum [40,97].
Besides, in the “transferred hypothesis” discussed above, the temporal order of cerebellar and vestibular nuclei synaptic plasticity in the acquisition and maintenance of motor memory is still questioned. Again, to answer this question, one could separately manipulate the estradiol synthesis in the cerebellum and vestibular nuclei in pre, during, and post adaptation. This kind of manipulation, indeed, should specifically affect the LTP at either cerebellar or vestibular nuclei synaptic sites depending on the case [40,41]. If, on the one hand, results from such investigations could help to clarify the temporal relationship between cerebellar and extra cerebellar plastic sites in motor memory formation, on the other hand, they may also be relevant for a better understanding of general mechanisms underlying the cerebellar dependent learning.

4.2. The Possible Relationship between Cerebellar nE2 Pathway and Other Brain Structures: Implication for Health

Generally speaking, the control of the strength of synapse transmission at parallel fibers to PCs, is essential to regulate the cerebellar cortex output to deep nuclei and extra-cerebellar areas in the brain [120,185,186]. Besides the brainstem, cerebellar communication via closed loops has been found with prefrontal, parietal, temporal, motor, and premotor cortex areas of the cerebral cortex [180,187,188,189]. The cerebellum, therefore, interacts with cortical networks that play a central role in many aspects of cognition, including attention, language, working memory, executive control, temporal representation, emotion, and addiction, as well as the movement [127,128,190]. Moreover, even for the vestibular system has been demonstrated interaction with thalamocortical pathways involved in the cognitive process [191,192,193]. Thus estradiol, including nE2, could widely impact motor and non-motor functions influenced by the cerebellum [120,194] (Figure 5).
There is general agreement that cerebellar output provides to the brain a forward function critical for effective control of movement and cognition [195,196]. The impairment of the forward control should lead to motor dysmetria, a primary cerebellar motor deficit characterized by the loss of the capability to perform rapid and coordinate movements, and for an analogy with the motor domain to cognitive dysmetria [197,198]. The latter indicates an alteration in the synchronous coordination of cognitive capacities leading to inability in controlling thought processes. Not surprisingly, cerebellar dysfunctions, with particular regards to vermis, are linked with schizophrenia and autism spectrum disorder [199,200,201]. Furthermore, schizophrenia and autism spectrum disorder both present dysregulated estrogen pathways during the neurodevelopment of the illness, and adjunctive estradiol therapy has provided a promising outlook in the treatment of schizophrenia [202,203]. Besides, mounting evidence indicates that in Alzheimer’s disease, the cerebellum accumulates Amyloid-β deposits with the consequence of losing PCs and altering cortical synaptic information processing that could contribute to the cognitive and motor deficits that characterize Alzheimer’s disease [132,204]. Concerning the latter issue, it is worthy of mentioning that estradiol may exert neuroprotection against toxic damage induced by Amyloid-β in the cerebellum [205]. Perhaps, despite the intricate estradiol’s interaction with the brain, also the nE2 action in modulating the computational power of the cerebellar cortex could play a significant role in the regulation of the cognitive as well as motor domain. (Figure 5). Further investigations are required for a better understanding of the interaction mechanism between nE2 and the cerebellar function and the possible significance for the cerebellum involvement in sensorimotor and mental processes. Furthermore, additional studies addressing whether sex differences in developing cognitive decline [194,202,203] are influenced by possible diverse interactions between the estrogen hormone and the cerebellar nE2 in males and females are recommended.

5. Conclusions

In summary, in contrast with the idea that in the nervous system, the estradiol synthesizing enzyme aromatase requires a high expression to produce useful amounts of nE2, recent investigations in the cerebellum indicate that even a low aromatase expression can be related to the fast nE2 modulation of neuronal physiology. It is now evident that nE2 and not just the gonadal-derived hormone may affect the cerebellar functioning not only during the neonatal life but also in adulthood. The evidence is that in the cerebellum of adult rodents, a brain structure typically characterized by low aromatase expression, nE2 rapidly impacts the glutamatergic neurotransmission at specific synaptic sites as those between the parallel fibers and PCs dendritic spines probably via activation of “non-classical” estradiol signaling in PCs. Furthermore, emerging data indicate that nE2 plays an essential role in the cerebellar-dependent behaviors and motor memory formation, affecting multiple synaptic plastic sites at the cerebellar and possibly extra-cerebellar networks. As an extension of the latter concept, the suggestion is that the nE2 synthesized in the cerebellum (and perhaps in nearby areas) could influence a variety of sensorimotor phenomena depending on cerebellar synaptic plasticity.
Moreover, the fact that the nE2 synthesis influences the computational processes of the cerebellar cortex leads to the consequence that it may even play a role in cognitive as well as motor functions involving the cerebellum. Since altered cortical cerebellar synaptic processing may be involved in different forms of cognitive decline, we believe that the findings reported here may offer absorbing insight into the possibility of developing novel adjunctive estrogen therapy strategies that can be useful in supporting neuroprotection.

Author Contributions

Conceptualization, R.P.; Formal analysis, R.P., and C.V.D.; Funding acquisition, C.V.D., and R.P.; Investigation, C.V.D., and R.P.; Software, C.V.D., A.B., S.C., and R.P.; Writing—original draft, R.P.; Writing—review and editing, C.V.D., R.P., S.C., and A.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are supported by University of Alabama (UAB) 2018–2019 Faculty Development Grant Program (FDGP) and National Science Foundation (NSF) #1539034 RII Track-2 Focused EPSCoR Collaborations (FEC): Bridging Cognitive Science and Neuroscience Using Innovative Imaging Technologies (CVD); Fondazione Cassa di Risparmio di Perugia #2015.0328.021 (R.P.).

Conflicts of Interest

The authors declare that they have no conflict of interest.

Abbreviations

nE2estradiol synthesized in the nervous system
ERsestrogen receptors
ERαestrogen receptors isoform α
ERβestrogen receptors isoform β
GPCRsG-protein-coupled receptors
GPER-1estrogen-G-protein-coupled receptors-1
Gαq-mERGαq-membrane estrogen receptor
ER-Xestrogen receptor-X
GABAGamma aminobutyric acid
VORvestibulo-ocular reflex
PCsPurkinje cells
BDNFbrain-derived neurotrophic factor; mGluR1a: metabotropic glutamate receptors type 1a
PF-LTPlong term potentiation at the parallel fiber-Purkinje cell synapse
PF-LTDlong term depression at the parallel fiber-Purkinje cell synapse
LTZletrozole
ICI 182,780estrogen receptor antagonist
NMDARN-methyl-D-aspartate receptor
LTPlong term potentiation
LTDlong term depression

References

  1. Cornil, C.A.; Ball, G.F.; Balthazart, J. Rapid control of male typical behaviors by brain-derived estrogens. Front. Neuroendocr. 2012, 33, 425–446. [Google Scholar] [CrossRef] [Green Version]
  2. Balthazart, J.; Baillien, M.; Cornil, C.A.; Ball, G.F. Preoptic aromatase modulates male sexual behavior: Slow and fast mechanisms of action. Physiol. Behav. 2004, 83, 247–270. [Google Scholar] [CrossRef]
  3. Hojo, Y.; Murakami, G.; Mukai, H.; Higo, S.; Hatanaka, Y.; Ogiue-Ikeda, M.; Ishii, H.; Kimoto, T.; Kawato, S. Estrogen synthesis in the brain—Role in synaptic plasticity and memory. Mol. Cell. Endocrinol. 2008, 290, 31–43. [Google Scholar] [CrossRef]
  4. Woolley, C.S. Acute effects of estrogen on neuronal physiology. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 657–680. [Google Scholar] [CrossRef]
  5. Ervin, K.S.; Phan, A.; Gabor, C.S.; Choleris, E. Rapid oestrogenic regulation of social and nonsocial learning. J. Neuroendocr. 2013, 25, 1116–1132. [Google Scholar] [CrossRef] [Green Version]
  6. Cooke, B.M.; Woolley, C.S. Gonadal hormone modulation of dendrites in the mammalian CNS. J. Neurobiol. 2005, 64, 34–46. [Google Scholar] [CrossRef]
  7. McEwen, B.S.; Alves, S.E. Estrogen actions in the central nervous system. Endocr. Rev. 1999, 20, 279–307. [Google Scholar] [CrossRef]
  8. Konkle, A.T.M.; McCarthy, M.M. Developmental time course of estradiol, testosterone, and dihydrotestosterone levels in discrete regions of male and female rat brain. Endocrinology 2011, 152, 223–235. [Google Scholar] [CrossRef] [Green Version]
  9. Hojo, Y.; Hattori, T.A.; Enami, T.; Furukawa, A.; Suzuki, K.; Ishii, H.T.; Mukai, H.; Morrison, J.H.; Janssen, W.G.; Kominami, S.; et al. Adult male rat hippocampus synthesizes estradiol from pregnenolone by cytochromes P45017alpha and P450 aromatase localized in neurons. Proc. Natl. Acad. Sci. USA 2004, 101, 865–870. [Google Scholar] [CrossRef] [Green Version]
  10. Baulieu, E.E. Neurosteroids: Of the nervous system, by the nervous system, for the nervous system. Recent Prog. Horm. Res. 1997, 52, 1–32. [Google Scholar]
  11. Compagnone, N.A.; Mellon, S.H. Neurosteroids: Biosynthesis and function of these novel neuromodulators. Front. Neuroendocr. 2000, 21, 1–56. [Google Scholar] [CrossRef]
  12. Kimoto, T.; Tsurugizawa, T.; Ohta, Y.; Makino, J.; Tamura, H.; Hojo, Y.; Takata, N.; Kawato, S. Neurosteroid synthesis by cytochrome p450-containing systems localized in the rat brain hippocampal neurons: N-methyl-D-aspartate and calcium-dependent synthesis. Endocrinology 2001, 142, 3578–3589. [Google Scholar] [CrossRef]
  13. Cornil, C.A.; Ball, G.F.; Balthazart, J. The dual action of estrogen hypothesis. Trends Neurosci. 2015, 38, 408–416. [Google Scholar] [CrossRef] [Green Version]
  14. Rosenfeld, C.S.; Shay, D.A.; Vieira-Potter, V.J. Cognitive effects of aromatase and possible role in memory disorders. Front. Endocrinol. (Lausanne) 2018, 9, 610. [Google Scholar] [CrossRef]
  15. Morris, J.A.; Jordan, C.L.; Breedlove, S.M. Sexual differentiation of the vertebrate nervous system. Nat. Neurosci. 2004, 7, 1034–1039. [Google Scholar] [CrossRef]
  16. Duarte-Guterman, P.; Yagi, S.; Chow, C.; Galea, L.A. Hippocampal learning, memory, and neurogenesis: Effects of sex and estrogens across the lifespan in adults. Horm. Behav. 2015, 74, 37–52. [Google Scholar] [CrossRef] [Green Version]
  17. Clayton, D.F. The genomic action potential. Neurobiol. Learn. Mem. 2000, 74, 185–216. [Google Scholar] [CrossRef] [Green Version]
  18. Jacob, J.; Sebastian, K.S.; Devassy, S.; Priyadarsini, L.; Farook, M.F.; Shameem, A.; Mathew, D.; Sreeja, S.; Thampan, R.V. Membrane estrogen receptors: Genomic actions and post transcriptional regulation. Mol. Cell. Endocrinol. 2006, 246, 34–41. [Google Scholar] [CrossRef]
  19. Vasudevan, N.; Pfaff, D.W. Non-genomic actions of estrogens and their interaction with genomic actions in the brain. Front. Neuroendocr. 2008, 29, 238–257. [Google Scholar] [CrossRef]
  20. Balthazart, J.; Choleris, E.; Remage-Healey, L. Steroids and the brain: 50 years of research, conceptual shifts and the ascent of non-classical and membrane-initiated actions. Horm. Behav. 2018, 99, 1–8. [Google Scholar] [CrossRef]
  21. Rudolph, L.M.; Cornil, C.A.; Mittelman-Smith, M.A.; Rainville, J.R.; Remage-Healey, L.; Sinchak, K.; Micevych, P.E. Actions of steroids: New neurotransmitters. J. Neurosci. 2016, 36, 11449–11458. [Google Scholar] [CrossRef] [Green Version]
  22. Ronnekleiv, O.K.; Kelly, M.J. Membrane-initiated effects of estradiol in the central nervous system. In Hormones, Brain and Behavior; Pfaff, D.W., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–22. [Google Scholar]
  23. Oberlander, J.G.; Woolley, C.S. 17beta-estradiol acutely potentiates glutamatergic synaptic transmission in the hippocampus through distinct mechanisms in males and females. J. Neurosci. 2016, 36, 2677–2690. [Google Scholar] [CrossRef]
  24. Kelly, M.J.; Moss, R.L.; Dudley, C.A. Differential sensitivity of preoptic-septal neurons to microelectrophoresed estrogen during the estrous cycle. Brain Res. 1976, 114, 152–157. [Google Scholar] [CrossRef]
  25. Mittelman-Smith, M.A.; Rudolph, L.M.; Mohr, M.A.; Micevych, P.E. Rodent models of non-classical progesterone action regulating ovulation. Front. Endocrinol. 2017, 8. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, D.J.; Su, L.D.; Wang, Y.N.; Yang, D.; Sun, C.L.; Zhou, L.; Wang, X.X.; Shen, Y. Long-term potentiation at cerebellar parallel fiber-Purkinje cell synapses requires presynaptic and postsynaptic signaling cascades. J. Neurosci. 2014, 34, 2355–2364. [Google Scholar] [CrossRef] [Green Version]
  27. Morissette, M.; Le Saux, M.; D’Astous, M.; Jourdain, S.; Al Sweidi, S.; Morin, N.; Estrada-Camarena, E.; Mendez, P.; Garcia-Segura, L.M.; Di Paolo, T.; et al. Contribution of estrogen receptors alpha and beta to the effects of estradiol in the brain. J. Steroid Biochem. Mol. Biol. 2008, 108, 327–338. [Google Scholar] [CrossRef] [Green Version]
  28. Kuo, J.; Hamid, N.; Bondar, G.; Prossnitz, E.R.; Micevych, P. Membrane estrogen receptors stimulate intracellular calcium release and progesterone synthesis in hypothalamic astrocytes. J. Neurosci. 2010, 30, 12950–12957. [Google Scholar] [CrossRef]
  29. Hazell, G.G.; Yao, S.T.; Roper, J.A.; Prossnitz, E.R.; O’Carroll, A.M.; Lolait, S.J. Localisation of GPR30, a novel G protein-coupled oestrogen receptor, suggests multiple functions in rodent brain and peripheral tissues. J. Endocrinol. 2009, 202, 223–236. [Google Scholar] [CrossRef]
  30. Levin, E.R. Plasma membrane estrogen receptors. Trends Endocrinol. Metab. 2009, 20, 477–482. [Google Scholar] [CrossRef] [Green Version]
  31. Toran-Allerand, C.D.; Guan, X.; MacLusky, N.J.; Horvath, T.L.; Diano, S.; Singh, M.; Connolly, E.S., Jr.; Nethrapalli, I.S.; Tinnikov, A.A. ER-X: A novel, plasma membrane-associated, putative estrogen receptor that is regulated during development and after ischemic brain injury. J. Neurosci. 2002, 22, 8391–8401. [Google Scholar] [CrossRef]
  32. Toran-Allerand, C.D. Estrogen and the brain—Beyond ER-alpha, ER-beta, and 17 beta-estradiol. Ann. N. Y. Acad. Sci. 2005, 1052, 136–144. [Google Scholar] [CrossRef]
  33. Qiu, J.; Ronnekleiv, O.K.; Kelly, M.J. Modulation of hypothalamic neuronal activity through a novel G-protein-coupled estrogen membrane receptor. Steroids 2008, 73, 985–991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Razandi, M.; Pedram, A.; Greene, G.L.; Levin, E.R. Cell membrane and nuclear estrogen receptors (ERs) originate from a single transcript: Studies of ER alpha and ER beta expressed in Chinese hamster ovary cells. Mol. Endocrinol. 1999, 13, 307–319. [Google Scholar] [CrossRef]
  35. Soltysik, K.; Czekaj, P. Membrane estrogen receptors—Is it an alternative way of estrogen action? J. Physiol. Pharmacol. 2013, 64, 129–142. [Google Scholar]
  36. Levin, E.R.; Hammes, S.R. Nuclear receptors outside the nucleus: Extranuclear signalling by steroid receptors. Nat. Rev. Mol. Cell Biol. 2016, 17, 783–797. [Google Scholar] [CrossRef] [Green Version]
  37. Melcangi, R.C.; Panzica, G.; Garcia-Segura, L.M. Neuroactive steroids: Focus on human brain. Neuroscience 2011, 191, 1–5. [Google Scholar] [CrossRef] [Green Version]
  38. Charlier, T.D.; Cornil, C.A.; Patte-Mensah, C.; Meyer, L.; Mensah-Nyagan, A.G.; Balthazart, J. Local modulation of steroid action: Rapid control of enzymatic activity. Front. Neurosci. 2015, 9, 83. [Google Scholar] [CrossRef] [Green Version]
  39. Hedges, V.L.; Chen, G.; Yu, L.; Krentzel, A.A.; Starrett, J.R.; Zhu, J.N.; Suntharalingam, P.; Remage-Healey, L.; Wang, J.J.; Ebner, T.J.; et al. Local estrogen synthesis regulates parallel fiber-purkinje cell neurotransmission within the cerebellar cortex. Endocrinology 2018, 159, 1328–1338. [Google Scholar] [CrossRef] [Green Version]
  40. Dieni, C.V.; Ferraresi, A.; Sullivan, J.A.; Grassi, S.; Pettorossi, V.E.; Panichi, R. Acute inhibition of estradiol synthesis impacts vestibulo-ocular reflex adaptation and cerebellar long-term potentiation in male rats. Brain Struct. Funct. 2018, 223, 837–850. [Google Scholar] [CrossRef] [Green Version]
  41. Grassi, S.; Frondaroli, A.; Dieni, C.; Scarduzio, M.; Pettorossi, V.E. Long-term potentiation in the rat medial vestibular nuclei depends on locally synthesized 17beta-estradiol. J. Neurosci. 2009, 29, 10779–10783. [Google Scholar] [CrossRef] [Green Version]
  42. Scarduzio, M.; Panichi, R.; Pettorossi, V.E.; Grassi, S. Synaptic long-term potentiation and depression in the rat medial vestibular nuclei depend on neural activation of estrogenic and androgenic signals. PLoS ONE 2013, 8, e80792. [Google Scholar] [CrossRef]
  43. Chen, B.S.; Roche, K.W. Regulation of NMDA receptors by phosphorylation. Neuropharmacology 2007, 53, 362–368. [Google Scholar] [CrossRef] [Green Version]
  44. Hojo, Y.; Higo, S.; Ishii, H.; Ooishi, Y.; Mukai, H.; Murakami, G.; Kominami, T.; Kimoto, T.; Honma, S.; Poirier, D.; et al. Comparison between hippocampus-synthesized and circulation-derived sex steroids in the hippocampus. Endocrinology 2009, 150, 5106–5112. [Google Scholar] [CrossRef]
  45. Mukai, H.; Kimoto, T.; Hojo, Y.; Kawato, S.; Murakami, G.; Higo, S.; Hatanaka, Y.; Ogiue-Ikeda, M. Modulation of synaptic plasticity by brain estrogen in the hippocampus. Biochim. Biophys. Acta 2010, 1800, 1030–1044. [Google Scholar] [CrossRef]
  46. Cornil, C.A.; Ball, G.F.; Balthazart, J. Functional significance of the rapid regulation of brain estrogen action: Where do the estrogens come from? Brain Res. 2006, 1126, 2–26. [Google Scholar] [CrossRef] [Green Version]
  47. Storman, E.M.; Liu, N.J.; Wessendorf, M.W.; Gintzler, A.R. Physical linkage of estrogen receptor alpha and aromatase in rat: Oligocrine and endocrine actions of CNS-produced estrogens. Endocrinology 2018, 159, 2683–2697. [Google Scholar] [CrossRef]
  48. Kretz, O.; Fester, L.; Wehrenberg, U.; Zhou, L.; Brauckmann, S.; Zhao, S.; Prange-Kiel, J.; Naumann, T.; Jarry, H.; Frotscher, M.; et al. Hippocampal synapses depend on hippocampal estrogen synthesis. J. Neurosci. 2004, 24, 5913–5921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Bailey, D.J.; Makeyeva, Y.V.; Paitel, E.R.; Pedersen, A.L.; Hon, A.T.; Gunderson, J.A.; Saldanha, C.J. Hippocampal aromatization modulates spatial memory and characteristics of the synaptic membrane in the male zebra finch. Endocrinology 2017, 158, 852–859. [Google Scholar] [CrossRef] [Green Version]
  50. Azcoitia, I.; Arevalo, M.A.; Garcia-Segura, L.M. Neural-derived estradiol regulates brain plasticity. J. Chem. Neuroanat. 2018, 89, 53–59. [Google Scholar] [CrossRef]
  51. Sellers, K.J.; Erli, F.; Raval, P.; Watson, I.A.; Chen, D.; Srivastava, D.P. Rapid modulation of synaptogenesis and spinogenesis by 17beta-estradiol in primary cortical neurons. Front. Cell. Neurosci. 2015, 9, 137. [Google Scholar] [CrossRef] [Green Version]
  52. Brocca, M.E.; Garcia-Segura, L.M. Non-reproductive functions of aromatase in the central nervous system under physiological and pathological conditions. Cell. Mol. Neurobiol. 2019, 39, 473–481. [Google Scholar] [CrossRef]
  53. Tremere, L.A.; Jeong, J.K.; Pinaud, R. Estradiol shapes auditory processing in the adult brain by regulating inhibitory transmission and plasticity-associated gene expression. J. Neurosci. 2009, 29, 5949–5963. [Google Scholar] [CrossRef]
  54. Pettorossi, V.E.; Di Mauro, M.; Scarduzio, M.; Panichi, R.; Tozzi, A.; Calabresi, P.; Grassi, S. Modulatory role of androgenic and estrogenic neurosteroids in determining the direction of synaptic plasticity in the CA1 hippocampal region of male rats. Physiol. Rep. 2013, 1, e00185. [Google Scholar] [CrossRef]
  55. Tozzi, A.; de Iure, A.; Tantucci, M.; Durante, V.; Quiroga-Varela, A.; Giampa, C.; Di Mauro, M.; Mazzocchetti, P.; Costa, C.; Di Filippo, M.; et al. Endogenous 17beta-estradiol is required for activity-dependent long-term potentiation in the striatum: Interaction with the dopaminergic system. Front. Cell. Neurosci. 2015, 9, 192. [Google Scholar] [CrossRef] [Green Version]
  56. Di Mauro, M.; Tozzi, A.; Calabresi, P.; Pettorossi, V.E.; Grassi, S. Neo-synthesis of estrogenic or androgenic neurosteroids determine whether long-term potentiation or depression is induced in hippocampus of male rat. Front. Cell. Neurosci. 2015, 9, 376. [Google Scholar] [CrossRef] [Green Version]
  57. Mermelstein, P.G.; Becker, J.B.; Surmeier, D.J. Estradiol reduces calcium currents in rat neostriatal neurons via a membrane receptor. J. Neurosci. 1996, 16, 595–604. [Google Scholar] [CrossRef]
  58. Li, J.; Siegel, M.; Yuan, M.K.; Zeng, Z.Y.; Finnucan, L.; Persky, R.; Hurn, P.D.; McCullough, L.D. Estrogen enhances neurogenesis and behavioral recovery after stroke. J. Cereb. Blood Flow Metab. 2011, 31, 413–425. [Google Scholar] [CrossRef] [Green Version]
  59. Zhou, L.; Fester, L.; von Blittersdorff, B.; Hassu, B.; Nogens, H.; Prange-Kiel, J.; Jarry, H.; Wegscheider, K.; Rune, G.M. Aromatase inhibitors induce spine synapse loss in the hippocampus of ovariectomized mice. Endocrinology 2010, 151, 1153–1160. [Google Scholar] [CrossRef] [Green Version]
  60. Roselli, C.E.; Horton, L.E.; Resko, J.A. Distribution and regulation of aromatase activity in the rat hypothalamus and limbic system. Endocrinology 1985, 117, 2471–2477. [Google Scholar] [CrossRef]
  61. Lymer, J.M.; Sheppard, P.A.S.; Kuun, T.; Blackman, A.; Jani, N.; Mahbub, S.; Choleris, E. Estrogens and their receptors in the medial amygdala rapidly facilitate social recognition in female mice. Psychoneuroendocrinology 2018, 89, 30–38. [Google Scholar] [CrossRef]
  62. Li, J.; Gibbs, R.B. Detection of estradiol in rat brain tissues: Contribution of local versus systemic production. Psychoneuroendocrinology 2019, 102, 84–94. [Google Scholar] [CrossRef] [PubMed]
  63. Wirth, S.; Yanike, M.; Frank, L.M.; Smith, A.C.; Brown, E.N.; Suzuki, W.A. Single neurons in the monkey hippocampus and learning of new associations. Science 2003, 300, 1578–1581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Liu, T.; Xu, D.; Ashe, J.; Bushara, K. Specificity of inferior olive response to stimulus timing. J. Neurophysiol. 2008, 100, 1557–1561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Reeber, S.L.; White, J.J.; George-Jones, N.A.; Sillitoe, R.V. Architecture and development of olivocerebellar circuit topography. Front. Neural Circuits 2012, 6, 115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Tabatadze, N.; Sato, S.M.; Woolley, C.S. Quantitative analysis of long-form aromatase mRNA in the male and female rat brain. PLoS ONE 2014, 9, e100628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Seredynski, A.L.; Balthazart, J.; Christophe, V.J.; Ball, G.F.; Cornil, C.A. Neuroestrogens rapidly regulate sexual motivation but not performance. J. Neurosci. 2013, 33, 164–174. [Google Scholar] [CrossRef] [Green Version]
  68. De Groof, G.; Balthazart, J.; Cornil, C.A.; Van der Linden, A. Topography and lateralized effect of acute aromatase inhibition on auditory processing in a seasonal songbird. J. Neurosci. 2017, 37, 4243–4254. [Google Scholar] [CrossRef]
  69. Cornil, C.A.; Taziaux, M.; Baillien, M.; Ball, G.F.; Balthazart, J. Rapid effects of aromatase inhibition on male reproductive behaviors in Japanese quail. Horm. Behav. 2006, 49, 45–67. [Google Scholar] [CrossRef] [Green Version]
  70. Sierra, A.; Azcoitia, I.; Garcia-Segura, L. Endogenous estrogen formation is neuroprotective in model of cerebellar ataxia. Endocrine 2003, 21, 43–51. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, Q.G.; Wang, R.M.; Tang, H.; Dong, Y.; Chan, A.; Sareddy, G.R.; Vadlamudi, R.K.; Brann, D.W. Brain-derived estrogen exerts anti-inflammatory and neuroprotective actions in the rat hippocampus. Mol. Cell. Endocrinol. 2014, 389, 84–91. [Google Scholar] [CrossRef] [Green Version]
  72. Luine, V. Estradiol: Mediator of memories, spine density and cognitive resilience to stress in female rodents. J. Steroid Biochem. Mol. Biol. 2016, 160, 189–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ratner, M.H.; Kumaresan, V.; Farb, D.H. Neurosteroid actions in memory and neurologic/neuropsychiatric disorders. Front. Endocrinol. (Lausanne) 2019, 10, 169. [Google Scholar] [CrossRef] [PubMed]
  74. Dieni, C.V.; Sullivan, J.A.; Faralli, M.; Contemori, S.; Biscarini, A.; Pettorossi, V.E.; Panichi, R. 17 beta-estradiol synthesis modulates cerebellar dependent motor memory formation in adult male rats. Neurobiol. Learn. Mem. 2018, 155, 276–286. [Google Scholar] [CrossRef] [PubMed]
  75. Graham, B.M.; Milad, M.R. Inhibition of estradiol synthesis impairs fear extinction in male rats. Learn. Mem. 2014, 21, 347–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Remage-Healey, L.; Dong, S.; Maidment, N.T.; Schlinger, B.A. Presynaptic control of rapid estrogen fluctuations in the songbird auditory forebrain. J. Neurosci. 2011, 31, 10034–10038. [Google Scholar] [CrossRef] [Green Version]
  77. Engler-Chiurazzi, E.B.; Brown, C.M.; Povroznik, J.M.; Simpkins, J.W. Estrogens as neuroprotectants: Estrogenic actions in the context of cognitive aging and brain injury. Prog. Neurobiol. 2017, 157, 188–211. [Google Scholar] [CrossRef] [Green Version]
  78. Hara, Y.; Waters, E.M.; McEwen, B.S.; Morrison, J.H. Estrogen effects on cognitive and synaptic health over the lifecourse. Physiol. Rev. 2015, 95, 785–807. [Google Scholar] [CrossRef] [Green Version]
  79. Biegon, A.; Kim, S.W.; Alexoff, D.L.; Jayne, M.; Carter, P.; Hubbard, B.; King, P.; Logan, J.; Muench, L.; Pareto, D.; et al. Unique distribution of aromatase in the human brain: In vivo studies with PET and [N-methyl-11C]vorozole. Synapse 2010, 64, 801–807. [Google Scholar] [CrossRef] [Green Version]
  80. Azcoitia, I.; Yague, J.G.; Garcia-Segura, L.M. Estradiol synthesis within the human brain. Neuroscience 2011, 191, 139–147. [Google Scholar] [CrossRef]
  81. Lavaque, E.; Mayen, A.; Azcoitia, I.; Tena-Sempere, M.; Garcia-Segura, L.M. Sex differences, developmental changes, response to injury and cAMP regulation of the mRNA levels of steroidogenic acute regulatory protein, cytochrome p450scc, and aromatase in the olivocerebellar system. J. Neurobiol. 2006, 66, 308–318. [Google Scholar] [CrossRef]
  82. Roselli, C.E.; Ellinwood, W.E.; Resko, J.A. Regulation of brain aromatase activity in rats. Endocrinology 1984, 114, 192–200. [Google Scholar] [CrossRef] [PubMed]
  83. Munetomo, A.; Hojo, Y.; Higo, S.; Kato, A.; Yoshida, K.; Shirasawa, T.; Shimizu, T.; Barron, A.; Kimoto, T.; Kawato, S.; et al. Aging-induced changes in sex-steroidogenic enzymes and sex-steroid receptors in the cortex, hypothalamus and cerebellum. J. Physiol. Sci. 2015, 65, 253–263. [Google Scholar] [CrossRef] [PubMed]
  84. Sakamoto, H.; Mezaki, Y.; Shikimi, H.; Ukena, K.; Tsutsui, K. Dendritic growth and spine formation in response to estrogen in the developing Purkinje cell. Endocrinology 2003, 144, 4466–4477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Tsutsui, K. Neurosteroids in the Purkinje cell: Biosynthesis, mode of action and functional significance. Mol. Neurobiol. 2008, 37, 116–125. [Google Scholar] [CrossRef] [PubMed]
  86. Tsutsui, K. Neurosteroid biosynthesis and action during cerebellar development. Cerebellum 2012, 11, 414–415. [Google Scholar] [CrossRef]
  87. Hedges, V.L.; Ebner, T.J.; Meisel, R.L.; Mermelstein, P.G. The cerebellum as a target for estrogen action. Front. Neuroendocr. 2012, 33, 403–411. [Google Scholar] [CrossRef] [Green Version]
  88. Ikeda, Y.; Nagai, A. Differential expression of the estrogen receptors alpha and beta during postnatal development of the rat cerebellum. Brain Res. 2006, 1083, 39–49. [Google Scholar] [CrossRef]
  89. Mitra, S.W. Immunolocalization of estrogen receptor beta in the mouse brain: Comparison with estrogen receptor alpha. Endocrinology 2003, 144, 2844. [Google Scholar] [CrossRef]
  90. Price, R.H.; Handa, R.J. Expression of estrogen receptor-beta protein and mRNA in the cerebellum of the rat. Neurosci. Lett. 2000, 288, 115–118. [Google Scholar] [CrossRef]
  91. Grassi, S.; Scarduzio, M.; Panichi, R.; Dall’Aglio, C.; Boiti, C.; Pettorossi, V.E. Opposite long-term synaptic effects of 17beta-estradiol and 5alpha-dihydrotestosterone and localization of their receptors in the medial vestibular nucleus of rats. Brain Res. Bull. 2013, 97, 1–7. [Google Scholar] [CrossRef]
  92. Grassi, S.; Frondaroli, A.; Scarduzio, M.; Dutia, M.B.; Dieni, C.; Pettorossi, V.E. Effects of 17beta-estradiol on glutamate synaptic transmission and neuronal excitability in the rat medial vestibular nuclei. Neuroscience 2010, 165, 1100–1114. [Google Scholar] [CrossRef] [PubMed]
  93. Horvath, T.L.; Wikler, K.C. Aromatase in developing sensory systems of the rat brain. J. Neuroendocr. 1999, 11, 77–84. [Google Scholar] [CrossRef] [PubMed]
  94. Dean, S.L.; Knutson, J.F.; Krebs-Kraft, D.L.; McCarthy, M.M. Prostaglandin E2 is an endogenous modulator of cerebellar development and complex behavior during a sensitive postnatal period. Eur. J. Neurosci. 2012, 35, 1218–1229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Abel, J.M.; Witt, D.M.; Rissman, E.F. Sex differences in the cerebellum and frontal cortex: Roles of estrogen receptor alpha and sex chromosome genes. Neuroendocrinology 2011, 93, 230–240. [Google Scholar] [CrossRef] [Green Version]
  96. Biamonte, F.; Assenza, G.; Marino, R.; D’Amelio, M.; Panteri, R.; Caruso, D.; Scurati, S.; Yague, J.G.; Garcia-Segura, L.M.; Cesa, R.; et al. Interactions between neuroactive steroids and reelin haploinsufficiency in Purkinje cell survival. Neurobiol. Dis. 2009, 36, 103–115. [Google Scholar] [CrossRef]
  97. Andreescu, C.E.; Milojkovic, B.A.; Haasdijk, E.D.; Kramer, P.; De Jong, F.H.; Krust, A.; De Zeeuw, C.I.; De Jeu, M.T. Estradiol improves cerebellar memory formation by activating estrogen receptor beta. J. Neurosci. 2007, 27, 10832–10839. [Google Scholar] [CrossRef]
  98. Yamada, N.M.; Hirata, S.; Kato, J. Distribution and postnatal changes of aromatase mRNA in the female rat brain. J. Steroid Biochem. Mol. Biol. 1994, 48, 529–533. [Google Scholar] [CrossRef]
  99. Jakab, R.L.; Wong, J.K.; Belcher, S.M. Estrogen receptor beta immunoreactivity in differentiating cells of the developing rat cerebellum. J. Comp. Neurol. 2001, 430, 396–409. [Google Scholar] [CrossRef]
  100. Mohamed, M.K.; Abdel-Rahman, A.A. Effect of long-term ovariectomy and estrogen replacement on the expression of estrogen receptor gene in female rats. Eur. J. Endocrinol. 2000, 142, 307–314. [Google Scholar] [CrossRef] [Green Version]
  101. Shughrue, P.J.; Lane, M.V.; Merchenthaler, I. Comparative distribution of estrogen receptor-alpha and -beta mRNA in the rat central nervous system. J. Comp. Neurol. 1997, 388, 507–525. [Google Scholar] [CrossRef]
  102. Zhang, J.Q.; Cai, W.Q.; Zhou, D.S.; Su, B.Y. Distribution and differences of estrogen receptor beta immunoreactivity in the brain of adult male and female rats. Brain Res. 2002, 935, 73–80. [Google Scholar] [CrossRef]
  103. Cardona-Gomez, G.P.; DonCarlos, L.; Garcia-Segura, L.M. Insulin-like growth factor I receptors and estrogen receptors colocalize in female rat brain. Neuroscience 2000, 99, 751–760. [Google Scholar] [CrossRef] [Green Version]
  104. Perez, S.E.; Chen, E.Y.; Mufson, E.J. Distribution of estrogen receptor alpha and beta immunoreactive profiles in the postnatal rat brain. Brain Res. Dev. Brain Res. 2003, 145, 117–139. [Google Scholar] [CrossRef]
  105. Shughrue, P.J.; Merchenthaler, I. Distribution of estrogen receptor beta immunoreactivity in the rat central nervous system. J. Comp. Neurol. 2001, 436, 64–81. [Google Scholar] [CrossRef]
  106. Boyden, E.S.; Raymond, J.L. Active reversal of motor memories reveals rules governing memory encoding. Neuron 2003, 39, 1031–1042. [Google Scholar] [CrossRef] [Green Version]
  107. Lisberger, S.G.; Miles, F.A. Role of primate medial vestibular nucleus in long-term adaptive plasticity of vestibuloocular reflex. J. Neurophysiol. 1980, 43, 1725–1745. [Google Scholar] [CrossRef]
  108. Robinson, D.A. Adaptive gain control of vestibuloocular reflex by the cerebellum. J. Neurophysiol. 1976, 39, 954–969. [Google Scholar] [CrossRef]
  109. Ito, M. Cerebellar learning in the vestibulo-ocular reflex. Trends Cogn. Sci. 1998, 2, 313–321. [Google Scholar] [CrossRef]
  110. Kimpo, R.R.; Boyden, E.S.; Katoh, A.; Ke, M.C.; Raymond, J.L. Distinct patterns of stimulus generalization of increases and decreases in VOR gain. J. Neurophysiol. 2005, 94, 3092–3100. [Google Scholar] [CrossRef] [Green Version]
  111. Titley, H.K.; Heskin-Sweezie, R.; Chung, J.Y.; Kassardjian, C.D.; Razik, F.; Broussard, D.M. Rapid consolidation of motor memory in the vestibuloocular reflex. J. Neurophysiol. 2007, 98, 3809–3812. [Google Scholar] [CrossRef] [Green Version]
  112. Titley, H.K.; Heskin-Sweezie, R.; Broussard, D.M. The bidirectionality of motor learning in the vestibulo-ocular reflex is a function of cerebellar mGluR1 receptors. J. Neurophysiol. 2010, 104, 3657–3666. [Google Scholar] [CrossRef] [Green Version]
  113. Boyden, E.S.; Katoh, A.; Pyle, J.L.; Chatila, T.A.; Tsien, R.W.; Raymond, J.L. Selective engagement of plasticity mechanisms for motor memory storage. Neuron 2006, 51, 823–834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Sasahara, K.; Shikimi, H.; Haraguchi, S.; Sakamoto, H.; Honda, S.; Harada, N.; Tsutsui, K. Mode of action and functional significance of estrogen-inducing dendritic growth, spinogenesis, and synaptogenesis in the developing Purkinje cell. J. Neurosci. 2007, 27, 7408–7417. [Google Scholar] [CrossRef] [PubMed]
  115. Coesmans, M.; Weber, J.T.; De Zeeuw, C.I.; Hansel, C. Bidirectional parallel fiber plasticity in the cerebellum under climbing fiber control. Neuron 2004, 44, 691–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. De Zeeuw, C.I.; Yeo, C.H. Time and tide in cerebellar memory formation. Curr. Opin. Neurobiol. 2005, 15, 667–674. [Google Scholar] [CrossRef] [PubMed]
  117. Jang, D.C.; Kim, S.J. Plasticity leading to cerebellum-dependent learning: Two different regions, two different types. Pflugers Arch. 2019, 471, 927–934. [Google Scholar] [CrossRef] [PubMed]
  118. Kassardjian, C.D.; Tan, Y.F.; Chung, J.Y.; Heskin, R.; Peterson, M.J.; Broussard, D.M. The site of a motor memory shifts with consolidation. J. Neurosci. 2005, 25, 7979–7985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Lisberger, S.G. Neural basis for motor learning in the vestibuloocular reflex of primates. III. Computational and behavioral analysis of the sites of learning. J. Neurophysiol. 1994, 72, 974–998. [Google Scholar] [CrossRef]
  120. Ito, M. Cerebellar circuitry as a neuronal machine. Prog. Neurobiol. 2006, 78, 272–303. [Google Scholar] [CrossRef]
  121. Boyden, E.S.; Katoh, A.; Raymond, J.L. Cerebellum-dependent learning: The role of multiple plasticity mechanisms. Annu. Rev. Neurosci. 2004, 27, 581–609. [Google Scholar] [CrossRef] [Green Version]
  122. Broussard, D.M.; Titley, H.K.; Antflick, J.; Hampson, D.R. Motor learning in the VOR: The cerebellar component. Exp. Brain Res. 2011, 210, 451–463. [Google Scholar] [CrossRef] [PubMed]
  123. Blazquez, P.M.; Hirata, Y.; Highstein, S.M. The vestibulo-ocular reflex as a model system for motor learning: What is the role of the cerebellum? Cerebellum 2004, 3, 188–192. [Google Scholar] [CrossRef] [PubMed]
  124. Gittis, A.H.; du Lac, S. Intrinsic and synaptic plasticity in the vestibular system. Curr. Opin. Neurobiol. 2006, 16, 385–390. [Google Scholar] [CrossRef] [PubMed]
  125. Broussard, D.M.; Kassardjian, C.D. Learning in a simple motor system. Learn. Mem. 2004, 11, 127–136. [Google Scholar] [CrossRef] [Green Version]
  126. Smith, S.S. Sensorimotor-correlated discharge recorded from ensembles of cerebellar Purkinje cells varies across the estrous cycle of the rat. J. Neurophysiol. 1995, 74, 1095–1108. [Google Scholar] [CrossRef]
  127. Strick, P.L.; Dum, R.P.; Fiez, J.A. Cerebellum and nonmotor function. Annu. Rev. Neurosci. 2009, 32, 413–434. [Google Scholar] [CrossRef] [Green Version]
  128. Manto, M.; Oulad Ben Taib, N. The contributions of the cerebellum in sensorimotor control: What are the prevailing opinions which will guide forthcoming studies? Cerebellum 2013, 12, 313–315. [Google Scholar] [CrossRef]
  129. Doya, K. Complementary roles of basal ganglia and cerebellum in learning and motor control. Curr. Opin. Neurobiol. 2000, 10, 732–739. [Google Scholar] [CrossRef]
  130. Ma, Q.; Zeng, L.L.; Shen, H.; Liu, L.; Hu, D. Altered cerebellar-cerebral resting-state functional connectivity reliably identifies major depressive disorder. Brain Res. 2013, 1495, 86–94. [Google Scholar] [CrossRef]
  131. Stoodley, C.J.; D’Mello, A.M.; Ellegood, J.; Jakkamsetti, V.; Liu, P.; Nebel, M.B.; Gibson, J.M.; Kelly, E.; Meng, F.; Cano, C.A.; et al. Altered cerebellar connectivity in autism and cerebellar-mediated rescue of autism-related behaviors in mice. Nat. Neurosci. 2017, 20, 1744–1751. [Google Scholar] [CrossRef]
  132. Jacobs, H.I.L.; Hopkins, D.A.; Mayrhofer, H.C.; Bruner, E.; van Leeuwen, F.W.; Raaijmakers, W.; Schmahmann, J.D. The cerebellum in Alzheimer’s disease: Evaluating its role in cognitive decline. Brain 2018, 141, 37–47. [Google Scholar] [CrossRef] [PubMed]
  133. Wu, T.; Hallett, M. The cerebellum in Parkinson’s disease. Brain 2013, 136, 696–709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Walter, J.T.; Alvina, K.; Womack, M.D.; Chevez, C.; Khodakhah, K. Decreases in the precision of Purkinje cell pacemaking cause cerebellar dysfunction and ataxia. Nat. Neurosci. 2006, 9, 389–397. [Google Scholar] [CrossRef] [PubMed]
  135. Firozan, B.; Goudarzi, I.; Elahdadi Salmani, M.; Lashkarbolouki, T.; Rezaei, A.; Abrari, K. Estradiol increases expression of the brain-derived neurotrophic factor after acute administration of ethanol in the neonatal rat cerebellum. Eur. J. Pharmacol. 2014, 732, 1–11. [Google Scholar] [CrossRef]
  136. Mirzatoni, A.; Spence, R.D.; Naranjo, K.C.; Saldanha, C.J.; Schlinger, B.A. Injury-induced regulation of steroidogenic gene expression in the cerebellum. J. Neurotrauma 2010, 27, 1875–1882. [Google Scholar] [CrossRef] [Green Version]
  137. Tsutsui, K.; Ukena, K.; Sakamoto, H.; Okuyama, S.; Haraguchi, S. Biosynthesis, mode of action, and functional significance of neurosteroids in the purkinje cell. Front. Endocrinol. (Lausanne) 2011, 2, 61. [Google Scholar] [CrossRef] [Green Version]
  138. Smith, S.S.; Waterhouse, B.D.; Woodward, D.J. Locally applied estrogens potentiate glutamate-evoked excitation of cerebellar Purkinje cells. Brain Res. 1988, 475, 272–282. [Google Scholar] [CrossRef]
  139. Smith, S.S.; Woodward, D.J.; Chapin, J.K. Sex steroids modulate motor-correlated increases in cerebellar discharge. Brain Res. 1989, 476, 307–316. [Google Scholar] [CrossRef]
  140. Dewing, P.; Boulware, M.I.; Sinchak, K.; Christensen, A.; Mermelstein, P.G.; Micevych, P. Membrane estrogen receptor-alpha interactions with metabotropic glutamate receptor 1a modulate female sexual receptivity in rats. J. Neurosci. 2007, 27, 9294–9300. [Google Scholar] [CrossRef] [Green Version]
  141. Boulware, M.I.; Weick, J.P.; Becklund, B.R.; Kuo, S.P.; Groth, R.D.; Mermelstein, P.G. Estradiol activates group I and II metabotropic glutamate receptor signaling, leading to opposing influences on cAMP response element-binding protein. J. Neurosci. 2005, 25, 5066–5078. [Google Scholar] [CrossRef]
  142. Luine, V.N.; Beck, K.D.; Bowman, R.E.; Frankfurt, M.; Maclusky, N.J. Chronic stress and neural function: Accounting for sex and age. J. Neuroendocr. 2007, 19, 743–751. [Google Scholar] [CrossRef] [PubMed]
  143. Remage-Healey, L. Frank beach award winner: Steroids as neuromodulators of brain circuits and behavior. Horm. Behav. 2014, 66, 552–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Balthazart, J.; Baillien, M.; Ball, G.F. Rapid control of brain aromatase activity by glutamatergic inputs. Endocrinology 2006, 147, 359–366. [Google Scholar] [CrossRef] [PubMed]
  145. Do Rego, J.L.; Seong, J.Y.; Burel, D.; Leprince, J.; Luu-The, V.; Tsutsui, K.; Tonon, M.C.; Pelletier, G.; Vaudry, H. Neurosteroid biosynthesis: Enzymatic pathways and neuroendocrine regulation by neurotransmitters and neuropeptides. Front. Neuroendocr. 2009, 30, 259–301. [Google Scholar] [CrossRef]
  146. Charlier, T.D.; Harada, N.; Balthazart, J.; Cornil, C.A. Human and quail aromatase activity is rapidly and reversibly inhibited by phosphorylating conditions. Endocrinology 2011, 152, 4199–4210. [Google Scholar] [CrossRef] [Green Version]
  147. Remage-Healey, L.; Oyama, R.K.; Schlinger, B.A. Elevated aromatase activity in forebrain synaptic terminals during song. J. Neuroendocr. 2009, 21, 191–199. [Google Scholar] [CrossRef] [Green Version]
  148. Haraguchi, S.; Hara, S.; Ubuka, T.; Mita, M.; Tsutsui, K. Possible role of pineal allopregnanolone in Purkinje cell survival. Proc. Natl. Acad. Sci. USA 2012, 109, 21110–21115. [Google Scholar] [CrossRef] [Green Version]
  149. Tsutsui, K. Kobayashi award: Discovery of cerebellar and pineal neurosteroids and their biological actions on the growth and survival of Purkinje cells during development (review). Gen. Comp. Endocrinol. 2019, 284, 113051. [Google Scholar] [CrossRef]
  150. Gu, Q.; Moss, R.L. 17 beta-Estradiol potentiates kainate-induced currents via activation of the cAMP cascade. J. Neurosci. 1996, 16, 3620–3629. [Google Scholar] [CrossRef]
  151. Oliveira, C.A.; Mahecha, G.A.; Carnes, K.; Prins, G.S.; Saunders, P.T.; Franca, L.R.; Hess, R.A. Differential hormonal regulation of estrogen receptors ERalpha and ERbeta and androgen receptor expression in rat efferent ductules. Reproduction 2004, 128, 73–86. [Google Scholar] [CrossRef] [Green Version]
  152. Wempe, M.F.; Buchanan, C.M.; Buchanan, N.L.; Edgar, K.J.; Hanley, G.A.; Ramsey, M.G.; Skotty, J.S.; Rice, P.J. Pharmacokinetics of letrozole in male and female rats: Influence of complexation with hydroxybutenyl-beta cyclodextrin. J. Pharm. Pharmacol. 2007, 59, 795–802. [Google Scholar] [CrossRef] [PubMed]
  153. Lang, E.J. Organization of olivocerebellar activity in the absence of excitatory glutamatergic input. J. Neurosci. 2001, 21, 1663–1675. [Google Scholar] [CrossRef] [PubMed]
  154. Mathy, A.; Clark, B.A.; Hausser, M. Synaptically induced long-term modulation of electrical coupling in the inferior olive. Neuron 2014, 81, 1290–1296. [Google Scholar] [CrossRef] [Green Version]
  155. Hoge, G.J.; Davidson, K.G.; Yasumura, T.; Castillo, P.E.; Rash, J.E.; Pereda, A.E. The extent and strength of electrical coupling between inferior olivary neurons is heterogeneous. J. Neurophysiol. 2011, 105, 1089–1101. [Google Scholar] [CrossRef] [Green Version]
  156. Llinas, R.R. The Noncontinuous nature of movement execution. Mot. Control Concepts Issues 1991, 50, 223–242. [Google Scholar]
  157. Smith, S.S. Oestrogen effects in olivo-cerebellar and hippocampal circuits. In Novartis Foundation Symposium; Chadwick, D.J., Ed.; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2000. [Google Scholar]
  158. Cavallotti, C.; Iacopino, L.; Amenta, F. Stimulatory effect of beta-estradiol treatment on GABA-degradative enzymes within rat cerebellar cortex. Neurosci. Lett. 1983, 39, 205–209. [Google Scholar] [CrossRef]
  159. Rudick, C.N.; Woolley, C.S. Estrogen regulates functional inhibition of hippocampal CA1 pyramidal cells in the adult female rat. J. Neurosci. 2001, 21, 6532–6543. [Google Scholar] [CrossRef] [Green Version]
  160. Jorntell, H.; Hansel, C. Synaptic memories upside down: Bidirectional plasticity at cerebellar parallel fiber-Purkinje cell synapses. Neuron 2006, 52, 227–238. [Google Scholar] [CrossRef] [Green Version]
  161. Nilsen, J.; Chen, S.; Brinton, R.D. Dual action of estrogen on glutamate-induced calcium signaling: Mechanisms requiring interaction between estrogen receptors and src/mitogen activated protein kinase pathway. Brain Res. 2002, 930, 216–234. [Google Scholar] [CrossRef]
  162. Wang, Y.T.; Linden, D.J. Expression of cerebellar long-term depression requires postsynaptic clathrin-mediated endocytosis. Neuron 2000, 25, 635–647. [Google Scholar] [CrossRef] [Green Version]
  163. Tempia, F.; Alojado, M.E.; Strata, P.; Knopfel, T. Characterization of the mGluR(1)-mediated electrical and calcium signaling in Purkinje cells of mouse cerebellar slices. J. Neurophysiol. 2001, 86, 1389–1397. [Google Scholar] [CrossRef] [PubMed]
  164. Vogt, K.E.; Canepari, M. On the induction of postsynaptic granule cell-Purkinje neuron LTP and LTD. Cerebellum 2010, 9, 284–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Hansel, C.; de Jeu, M.; Belmeguenai, A.; Houtman, S.H.; Buitendijk, G.H.; Andreev, D.; De Zeeuw, C.I.; Elgersma, Y. alphaCaMKII Is essential for cerebellar LTD and motor learning. Neuron 2006, 51, 835–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Schonewille, M.; Belmeguenai, A.; Koekkoek, S.K.; Houtman, S.H.; Boele, H.J.; van Beugen, B.J.; Gao, Z.; Badura, A.; Ohtsuki, G.; Amerika, W.E.; et al. Purkinje cell-specific knockout of the protein phosphatase PP2B impairs potentiation and cerebellar motor learning. Neuron 2010, 67, 618–628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Galiana, H.L. A new approach to understanding adaptive visual-vestibular interactions in the central nervous system. J. Neurophysiol. 1986, 55, 349–374. [Google Scholar] [CrossRef] [PubMed]
  168. Raymond, J.L.; Lisberger, S.G.; Mauk, M.D. The cerebellum: A neuronal learning machine? Science 1996, 272, 1126–1131. [Google Scholar] [CrossRef]
  169. Anzai, M.; Kitazawa, H.; Nagao, S. Effects of reversible pharmacological shutdown of cerebellar flocculus on the memory of long-term horizontal vestibulo-ocular reflex adaptation in monkeys. Neurosci. Res. 2010, 68, 191–198. [Google Scholar] [CrossRef]
  170. Menzies, J.R.; Porrill, J.; Dutia, M.; Dean, P. Synaptic plasticity in medial vestibular nucleus neurons: Comparison with computational requirements of VOR adaptation. PLoS ONE 2010, 5, e13182. [Google Scholar] [CrossRef] [Green Version]
  171. Evrard, H.C.; Harada, N.; Balthazart, J. Immunocytochemical localization of aromatase in sensory and integrating nuclei of the hindbrain in Japanese quail (Coturnix japonica). J. Comp. Neurol. 2004, 473, 194–212. [Google Scholar] [CrossRef]
  172. Smith, P.F.; Agrawal, Y.; Darlington, C.L. Sexual dimorphism in vestibular function and dysfunction. J. Neurophysiol. 2019, 121, 2379–2391. [Google Scholar] [CrossRef]
  173. Panichi, R.; Botti, F.M.; Ferraresi, A.; Faralli, M.; Kyriakareli, A.; Schieppati, M.; Pettorossi, V.E. Self-motion perception and vestibulo-ocular reflex during whole body yaw rotation in standing subjects: The role of head position and neck proprioception. Hum. Mov. Sci. 2011, 30, 314–332. [Google Scholar] [CrossRef] [PubMed]
  174. Pettorossi, V.E.; Panichi, R.; Botti, F.M.; Kyriakareli, A.; Ferraresi, A.; Faralli, M.; Schieppati, M.; Bronstein, A.M. Prolonged asymmetric vestibular stimulation induces opposite, long-term effects on self-motion perception and ocular responses. J. Physiol. 2013, 591, 1907–1920. [Google Scholar] [CrossRef] [PubMed]
  175. Panichi, R.; Occhigrossi, C.; Ferraresi, A.; Faralli, M.; Lucertini, M.; Pettorossi, V.E. Adaptive changes in the perception of fast and slow movement at different head positions. Aerosp. Med. Hum. Perform. 2017, 88, 463–468. [Google Scholar] [CrossRef] [PubMed]
  176. Pettorossi, V.E.; Panichi, R.; Botti, F.M.; Biscarini, A.; Filippi, G.M.; Schieppati, M. Long-lasting effects of neck muscle vibration and contraction on self-motion perception of vestibular origin. Clin. Neurophysiol. 2015, 126, 1886–1900. [Google Scholar] [CrossRef] [PubMed]
  177. Scarduzio, M.; Panichi, R.; Pettorossi, V.E.; Grassi, S. The repetition timing of high frequency afferent stimulation drives the bidirectional plasticity at central synapses in the rat medial vestibular nuclei. Neuroscience 2012, 223, 1–11. [Google Scholar] [CrossRef] [PubMed]
  178. Panichi, R.; Faralli, M.; Bruni, R.; Kiriakarely, A.; Occhigrossi, C.; Ferraresi, A.; Bronstein, A.M.; Pettorossi, V.E. Asymmetric vestibular stimulation reveals persistent disruption of motion perception in unilateral vestibular lesions. J. Neurophysiol. 2017, 118, 2819–2832. [Google Scholar] [CrossRef]
  179. Gunther, L.; Beck, R.; Xiong, G.; Potschka, H.; Jahn, K.; Bartenstein, P.; Brandt, T.; Dutia, M.; Dieterich, M.; Strupp, M.; et al. N-acetyl-L-leucine accelerates vestibular compensation after unilateral labyrinthectomy by action in the cerebellum and thalamus. PLoS ONE 2015, 10, e0120891. [Google Scholar] [CrossRef]
  180. Sokolov, A.A.; Miall, R.C.; Ivry, R.B. The cerebellum: Adaptive prediction for movement and cognition. Trends Cogn. Sci. 2017, 21, 313–332. [Google Scholar] [CrossRef] [Green Version]
  181. Morton, S.M.; Bastian, A.J. Cerebellar control of balance and locomotion. Neuroscientist 2004, 10, 247–259. [Google Scholar] [CrossRef]
  182. Cangussu, L.M.; Nahas-Neto, J.; Petri Nahas, E.A.; Rodrigues Barral, A.B.; Buttros Dde, A.; Uemura, G. Evaluation of postural balance in postmenopausal women and its relationship with bone mineral density—A cross sectional study. BMC Musculoskelet. Disord. 2012, 13, 2. [Google Scholar] [CrossRef] [Green Version]
  183. Mayer, I.A.; Abramson, V.G.; Isakoff, S.J.; Forero, A.; Balko, J.M.; Kuba, M.G.; Sanders, M.E.; Yap, J.T.; Van den Abbeele, A.D.; Li, Y.; et al. Stand up to cancer phase Ib study of pan-phosphoinositide-3-kinase inhibitor buparlisib with letrozole in estrogen receptor-positive/human epidermal growth factor receptor 2-negative metastatic breast cancer. J. Clin. Oncol. 2014, 32, 1202–1209. [Google Scholar] [CrossRef] [PubMed]
  184. Santen, R.J.; Manni, A.; Harvey, H.; Redmond, C. Endocrine treatment of breast cancer in women. Endocr. Rev. 1990, 11, 221–265. [Google Scholar] [CrossRef]
  185. Kelly, R.M.; Strick, P.L. Cerebellar loops with motor cortex and prefrontal cortex of a nonhuman primate. J. Neurosci. 2003, 23, 8432–8444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Salmi, J.; Pallesen, K.J.; Neuvonen, T.; Brattico, E.; Korvenoja, A.; Salonen, O.; Carlson, S. Cognitive and motor loops of the human cerebro-cerebellar system. J. Cogn. Neurosci. 2010, 22, 2663–2676. [Google Scholar] [CrossRef] [PubMed]
  187. O’Reilly, J.X.; Beckmann, C.F.; Tomassini, V.; Ramnani, N.; Johansen-Berg, H. Distinct and overlapping functional zones in the cerebellum defined by resting state functional connectivity. Cereb. Cortex 2010, 20, 953–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Dum, R.P.; Strick, P.L. An unfolded map of the cerebellar dentate nucleus and its projections to the cerebral cortex. J. Neurophysiol. 2003, 89, 634–639. [Google Scholar] [CrossRef] [PubMed]
  189. Sokolov, A.A.; Erb, M.; Grodd, W.; Pavlova, M.A. Structural loop between the cerebellum and the superior temporal sulcus: Evidence from diffusion tensor imaging. Cereb. Cortex 2014, 24, 626–632. [Google Scholar] [CrossRef] [Green Version]
  190. Schmahmann, J.D. The cerebrocerebellar system: Anatomic substrates of the cerebellar contribution to cognition and emotion. Int. Rev. Psychiatr. 2001, 13, 247–260. [Google Scholar] [CrossRef]
  191. Besnard, S.; Lopez, C.; Brandt, T.; Denise, P.; Smith, P.F. Editorial: The vestibular system in cognitive and memory processes in mammalians. Front. Integr. Neurosci. 2015, 9, 55. [Google Scholar] [CrossRef] [Green Version]
  192. Mast, F.W.; Preuss, N.; Hartmann, M.; Grabherr, L. Spatial cognition, body representation and affective processes: The role of vestibular information beyond ocular reflexes and control of posture. Front. Integr. Neurosci. 2014, 8, 44. [Google Scholar] [CrossRef] [Green Version]
  193. Smith, P.F.; Darlington, C.L. Personality changes in patients with vestibular dysfunction. Front. Hum. Neurosci. 2013, 7, 678. [Google Scholar] [CrossRef] [Green Version]
  194. Dean, S.L.; McCarthy, M.M. Steroids, sex and the cerebellar cortex: Implications for human disease. Cerebellum 2008, 7, 38–47. [Google Scholar] [CrossRef]
  195. Ishikawa, T.; Tomatsu, S.; Izawa, J.; Kakei, S. The cerebro-cerebellum: Could it be loci of forward models? Neurosci. Res. 2016, 104, 72–79. [Google Scholar] [CrossRef] [Green Version]
  196. Ito, M. Control of mental activities by internal models in the cerebellum. Nat. Rev. Neurosci. 2008, 9, 304–313. [Google Scholar] [CrossRef]
  197. Schmahmann, J.D. Disorders of the cerebellum: Ataxia, dysmetria of thought, and the cerebellar cognitive affective syndrome. J. Neuropsychiatry Clin. Neurosci. 2004, 16, 367–378. [Google Scholar] [CrossRef] [PubMed]
  198. Hore, J.; Wild, B.; Diener, H.C. Cerebellar dysmetria at the elbow, wrist, and fingers. J. Neurophysiol. 1991, 65, 563–571. [Google Scholar] [CrossRef] [PubMed]
  199. Hampson, D.R.; Blatt, G.J. Autism spectrum disorders and neuropathology of the cerebellum. Front. Neurosci. 2015, 9, 420. [Google Scholar] [CrossRef] [Green Version]
  200. Parker, K.L.; Narayanan, N.S.; Andreasen, N.C. The therapeutic potential of the cerebellum in schizophrenia. Front. Syst. Neurosci. 2014, 8, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Andreasen, N.C.; Paradiso, S.; O’Leary, D.S. “Cognitive dysmetria” as an integrative theory of schizophrenia: A dysfunction in cortical-subcortical-cerebellar circuitry? Schizophr. Bull. 1998, 24, 203–218. [Google Scholar] [CrossRef] [PubMed]
  202. Crider, A.; Pillai, A. Estrogen signaling as a therapeutic target in neurodevelopmental disorders. J. Pharmacol. Exp. Ther. 2017, 360, 48–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Gogos, A.; Sbisa, A.M.; Sun, J.; Gibbons, A.; Udawela, M.; Dean, B. A role for estrogen in schizophrenia: Clinical and preclinical findings. Int. J. Endocrinol. 2015, 2015, 615356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Hoxha, E.; Lippiello, P.; Zurlo, F.; Balbo, I.; Santamaria, R.; Tempia, F.; Miniaci, M.C. The emerging role of altered cerebellar synaptic processing in Alzheimer’s disease. Front. Aging Neurosci. 2018, 10, 396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Napolitano, M.; Costa, L.; Piacentini, R.; Grassi, C.; Lanzone, A.; Gulino, A. 17beta-estradiol protects cerebellar granule cells against beta-amyloid-induced toxicity via the apoptotic mitochondrial pathway. Neurosci. Lett. 2014, 561, 134–139. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biosynthetic pathway for neurosteroids in the brain. The arrows indicate biosynthetic pathways of neurosteroids identified in the brain. P450scc, Cytochrome P450 cholesterol side-chain cleavage enzyme; p450c17, cytochrome P450 17a-hydroxylase/C17; DHEA, dehydroepiandrosterone; 17β-HSD, 17beta-hydroxysteroid dehydrogenase; 3β-HSD 3beta-hydroxysteroid dehydrogenase D5–D4 isomerase; 5α-R, 5alfa-reductase; p450ARO, cytochrome P450 aromatase; 3α-HSD 3alfa-hydroxysteroid dehydrogenase D5–D4 isomerase.
Figure 1. Biosynthetic pathway for neurosteroids in the brain. The arrows indicate biosynthetic pathways of neurosteroids identified in the brain. P450scc, Cytochrome P450 cholesterol side-chain cleavage enzyme; p450c17, cytochrome P450 17a-hydroxylase/C17; DHEA, dehydroepiandrosterone; 17β-HSD, 17beta-hydroxysteroid dehydrogenase; 3β-HSD 3beta-hydroxysteroid dehydrogenase D5–D4 isomerase; 5α-R, 5alfa-reductase; p450ARO, cytochrome P450 aromatase; 3α-HSD 3alfa-hydroxysteroid dehydrogenase D5–D4 isomerase.
Ijms 21 03316 g001
Figure 2. Schematic representation of the classic and non-classic mode of action of the estrogen estradiol. The schema represents possible estradiol effects on neural targets. Both classical and non-classical effects require estrogen receptors activation by estradiol (nE2). In the classic genomic mode of action, nE2 binds to cytoplasmatic estrogen receptors beta or alpha (ERs: ERα, ERβ), receptors dimerize (not shown), and translocate to the nucleus. Once bound specific estrogen response element on the DNA, the dimer possibly recruits transcriptional coregulator to modulates the gene transcription. The results may be a slow and long-lasting effect on cells and, ultimately, on the entire organism. In the non-classical mode of action, nE2 binds cytoplasmatic or membrane-associated estrogen receptors (ERs, GPER-1, Gαq). The binding triggers intracellular signaling cascades involving several kinases (e.g., PCA, Protein kinase A; PKC, Protein kinases C; MAPK, mitogen-activated protein kinase; PI3K, phosphoinositide 3-kinase), adenylyl cyclase (AC), phospholipase C (PLC), and fluctuation in intracellular calcium concentration. Such intracellular signaling may result in the rapid modulation of synaptic function (the schema also shows possible neurotrophic effects) and, ultimately, in behavioral modifications. Finally, in the estrogen-mediated indirect genetic effect, nE2 binding to estrogen receptors induces signaling cascades that might activate transcription factors (e.g., CREB, cAMP response element-binding protein) to regulate the gene transcription. AKT, Protein kinase B; AMPA, ionotropic glutamate receptor; ER, classical estrogen receptor; ERK, extracellular signal-regulated kinase; GPER-1, G protein-coupled transmembrane estrogen receptor-1; Gαq-mER, Gαq-coupled membrane-associated estrogen receptor; mER, membrane-associated classical estrogen receptor; mGluR, metabotropic glutamate receptor; NMDA, ionotropic glutamate receptor.
Figure 2. Schematic representation of the classic and non-classic mode of action of the estrogen estradiol. The schema represents possible estradiol effects on neural targets. Both classical and non-classical effects require estrogen receptors activation by estradiol (nE2). In the classic genomic mode of action, nE2 binds to cytoplasmatic estrogen receptors beta or alpha (ERs: ERα, ERβ), receptors dimerize (not shown), and translocate to the nucleus. Once bound specific estrogen response element on the DNA, the dimer possibly recruits transcriptional coregulator to modulates the gene transcription. The results may be a slow and long-lasting effect on cells and, ultimately, on the entire organism. In the non-classical mode of action, nE2 binds cytoplasmatic or membrane-associated estrogen receptors (ERs, GPER-1, Gαq). The binding triggers intracellular signaling cascades involving several kinases (e.g., PCA, Protein kinase A; PKC, Protein kinases C; MAPK, mitogen-activated protein kinase; PI3K, phosphoinositide 3-kinase), adenylyl cyclase (AC), phospholipase C (PLC), and fluctuation in intracellular calcium concentration. Such intracellular signaling may result in the rapid modulation of synaptic function (the schema also shows possible neurotrophic effects) and, ultimately, in behavioral modifications. Finally, in the estrogen-mediated indirect genetic effect, nE2 binding to estrogen receptors induces signaling cascades that might activate transcription factors (e.g., CREB, cAMP response element-binding protein) to regulate the gene transcription. AKT, Protein kinase B; AMPA, ionotropic glutamate receptor; ER, classical estrogen receptor; ERK, extracellular signal-regulated kinase; GPER-1, G protein-coupled transmembrane estrogen receptor-1; Gαq-mER, Gαq-coupled membrane-associated estrogen receptor; mER, membrane-associated classical estrogen receptor; mGluR, metabotropic glutamate receptor; NMDA, ionotropic glutamate receptor.
Ijms 21 03316 g002
Figure 3. Schematic cartoon representing the post-translational mechanisms modulating aromatase activity. Fluctuations in intracellular calcium (Ca2+) concentration, along with glutamatergic and dopaminergic signals, might rapidly modulate intracellular pathways to regulate aromatase activity via phosphorylation/dephosphorylation processes. As a result, the transformation rate of testosterone (T) into 17 beta-estradiol (nE2) may be rapidly regulated.
Figure 3. Schematic cartoon representing the post-translational mechanisms modulating aromatase activity. Fluctuations in intracellular calcium (Ca2+) concentration, along with glutamatergic and dopaminergic signals, might rapidly modulate intracellular pathways to regulate aromatase activity via phosphorylation/dephosphorylation processes. As a result, the transformation rate of testosterone (T) into 17 beta-estradiol (nE2) may be rapidly regulated.
Ijms 21 03316 g003
Figure 4. Impact of estradiol on plasticity sites in the vestibulo-cerebellar network. (A) Schematic representation of the neural connections mediating the VOR and possible sites of influence of circulating estradiol and nE2 (dashed lines); (B) Effect of nE2 on the plasticity mechanisms that could be responsible for the adaptive modification of the VOR. The red and green pathways indicate the modifications in the VOR circuitry underlying the gain down and gain up adaptation, respectively. In the bottom are reported representative VOR position traces recorded by an infrared technique in adult rats during horizontal head rotation (Hr, black dashed traces) in the dark in the control condition (black) and after adaptation paradigms (red and green). AOS, accessory optic system; Br, brainstem; Cc, cerebellar cortex; Cf, climbing fiber; Em, extra-ocular muscles; Es, error signal; Gc, granule cell; Io, inferior olive; Mc, Motor command; Mf, mossy fibers; nE2, neurosteroid estradiol; Om, ocular motoneurons in the midbrain; Pc Purkinje cell; Pf, parallel fiber; PF-LTP, LTP at the parallel fiber-Purkinje cell synapse; PF-LTD, LTD at the parallel fiber-Purkinje cell synapse; Qp, quick eye movement phases; Si, sensory inflow; Sp, slow eye movement phases; Vn, vestibular nuclei; VN-LTP, LTP in the vestibular nuclei; VN-LTD, LTD in the vestibular nuclei; VOR, vestibulo-ocular reflex.
Figure 4. Impact of estradiol on plasticity sites in the vestibulo-cerebellar network. (A) Schematic representation of the neural connections mediating the VOR and possible sites of influence of circulating estradiol and nE2 (dashed lines); (B) Effect of nE2 on the plasticity mechanisms that could be responsible for the adaptive modification of the VOR. The red and green pathways indicate the modifications in the VOR circuitry underlying the gain down and gain up adaptation, respectively. In the bottom are reported representative VOR position traces recorded by an infrared technique in adult rats during horizontal head rotation (Hr, black dashed traces) in the dark in the control condition (black) and after adaptation paradigms (red and green). AOS, accessory optic system; Br, brainstem; Cc, cerebellar cortex; Cf, climbing fiber; Em, extra-ocular muscles; Es, error signal; Gc, granule cell; Io, inferior olive; Mc, Motor command; Mf, mossy fibers; nE2, neurosteroid estradiol; Om, ocular motoneurons in the midbrain; Pc Purkinje cell; Pf, parallel fiber; PF-LTP, LTP at the parallel fiber-Purkinje cell synapse; PF-LTD, LTD at the parallel fiber-Purkinje cell synapse; Qp, quick eye movement phases; Si, sensory inflow; Sp, slow eye movement phases; Vn, vestibular nuclei; VN-LTP, LTP in the vestibular nuclei; VN-LTD, LTD in the vestibular nuclei; VOR, vestibulo-ocular reflex.
Ijms 21 03316 g004
Figure 5. Depiction representing cerebellar nE2 pathways that could influence motor behavior and cognitive functions. Cec, cerebral cortex; Dn, cerebellar deep nuclei; nE2, neurosteroid estradiol; Pcs, Purkinje cells; Pf-LTP, LTP at the parallel fiber-Purkinje cell synapse; Vn, vestibular nuclei; Vn-LTP, LTP in the vestibular nuclei.
Figure 5. Depiction representing cerebellar nE2 pathways that could influence motor behavior and cognitive functions. Cec, cerebral cortex; Dn, cerebellar deep nuclei; nE2, neurosteroid estradiol; Pcs, Purkinje cells; Pf-LTP, LTP at the parallel fiber-Purkinje cell synapse; Vn, vestibular nuclei; Vn-LTP, LTP in the vestibular nuclei.
Ijms 21 03316 g005
Table 1. Summary of findings indicating aromatase and estrogen receptors expression and distribution within the cerebellar cortex in rodents.
Table 1. Summary of findings indicating aromatase and estrogen receptors expression and distribution within the cerebellar cortex in rodents.
Cerebellar Cortex LayerAromataseERαERβGPER-1
ChildhoodAdultChildhoodAdultChildhoodAdultChildhoodAdult
Gc1, 71 11, 12, 208, 11, 14, 15, 19, 21 18
PC1, 7111, 20117, 11, 1211, 8, 14, 15–17, 19, 21 18
Ml 11, 1211, 19 18
NI4, 5, 102, 3, 4, 9, 3, 6, 13 3, 6
Data are mainly based on immunohistochemical and mRNA detection techniques. Childhood indicates the period from neonatal life to prepuberty. Adult indicates the period of life starting from peri puberty. ERα, estrogen receptors alpha; ERβ, estrogen receptors beta; Gc, granular cell; GPER-1, G protein estrogen receptors (G protein-coupled transmembrane receptors); Ml, molecular layer; NI, localization not identified (studies only based on mRNA detection techniques); PC, Purkinje cell. 1, Sakamoto et al., 2003 [84]; 2 Tabatadze et al., 2014 [66]; 3 Munetomo et al., 2015 [83]; 4 Lavaque et al., 2005 [81]; 5 Dean et al., 2012 [94]; 6 Abel et al., 2011 [95]; 7 Biamonte et al., 2009 [96]; 8 Andreescu et al., 2008 [97]; 9 Roselli et al. 1984 [82]; 10 Yamada et al. 1994 [98]; 11 Ikeda and Nakai 2006 [88]; 12 Jakab et al. 2001 [99]; 13 Mohamed et al., 2000 [100]; 14 Mitra et al., 2003 [89]; 15 Price and Handa et al. 2000 [90]; 16 Shughrue et al. 1997 [101]; 17 Zhang et al. 2002 [102]; 18 Hazell et al., 2009 [29]; 19 Cardona-Gomez et al. 2000 [103]; 20 Perez et al. 2003 [104]; 21 Shughrue and Merchenthaler 2001 [105]. See the text for detail.

Share and Cite

MDPI and ACS Style

Dieni, C.V.; Contemori, S.; Biscarini, A.; Panichi, R. De Novo Synthesized Estradiol: A Role in Modulating the Cerebellar Function. Int. J. Mol. Sci. 2020, 21, 3316. https://doi.org/10.3390/ijms21093316

AMA Style

Dieni CV, Contemori S, Biscarini A, Panichi R. De Novo Synthesized Estradiol: A Role in Modulating the Cerebellar Function. International Journal of Molecular Sciences. 2020; 21(9):3316. https://doi.org/10.3390/ijms21093316

Chicago/Turabian Style

Dieni, Cristina V., Samuele Contemori, Andrea Biscarini, and Roberto Panichi. 2020. "De Novo Synthesized Estradiol: A Role in Modulating the Cerebellar Function" International Journal of Molecular Sciences 21, no. 9: 3316. https://doi.org/10.3390/ijms21093316

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

Article Metrics

Back to TopTop