Skip to main content

The peri-menopause in a woman’s life: a systemic inflammatory phase that enables later neurodegenerative disease

Abstract

The peri-menopause or menopausal transition—the time period that surrounds the final years of a woman’s reproductive life—is associated with profound reproductive and hormonal changes in a woman’s body and exponentially increases a woman’s risk of cerebral ischemia and Alzheimer’s disease. Although our understanding of the exact timeline or definition of peri-menopause is limited, it is clear that there are two stages to the peri-menopause. These are the early menopausal transition, where menstrual cycles are mostly regular, with relatively few interruptions, and the late transition, where amenorrhea becomes more prolonged and lasts for at least 60 days, up to the final menstrual period. Emerging evidence is showing that peri-menopause is pro-inflammatory and disrupts estrogen-regulated neurological systems. Estrogen is a master regulator that functions through a network of estrogen receptors subtypes alpha (ER-α) and beta (ER-β). Estrogen receptor-beta has been shown to regulate a key component of the innate immune response known as the inflammasome, and it also is involved in regulation of neuronal mitochondrial function. This review will present an overview of the menopausal transition as an inflammatory event, with associated systemic and central nervous system inflammation, plus regulation of the innate immune response by ER-β-mediated mechanisms.

Introduction

Aging is a complex, predetermined natural process. This natural process of aging is associated with reproductive senescence in most vertebrates, including mammals of both sexes. For most of the vertebrates, including lower mammals, life ends around the attainment of reproductive senescence. Among lower mammals, in only five species of whale (including killer whales) females and women experience true menopause and survive years after menopausal transition occurs. Menopause is defined as the cessation of the menstrual cycle due to anovulation. It is verified retrospectively after one year of amenorrhea [1]. The phenomenon of menopause brings multiple physiological changes in the body, and the age at which menopause occurs is increasingly recognized as an indicator for health outcomes in later life. An average age of menopause is between 45 and 51 years in the USA. The transition to menopause usually lasts about 7 years but can last as long as 14 years. A 2012 study of stroke risk in women found that if natural menopause occurs before 42 years of age, then the risk of stroke doubles [2]. In general, women’s risk of stroke and cardiac arrest increases exponentially after the onset of menopause. Both stroke and cardiac arrest cause focal and global cerebral ischemia (CI), respectively, with a major complication of cognitive decline [3,4,5,6,7]. Along with loss of ovarian functions, the endocrine transition from the estrogen cycling of the reproductive phase to the estrogen decline of the reproductively quiescent, post-menopausal phase is associated with mild cognitive dysfunction, which has been proposed to be a prodromal phase of Alzheimer’s disease (AD) [8]. The endocrine transition is also associated with a rise in chronic low-grade inflammation [9]. The persistent low-grade inflammation in turn accelerates ovarian failure [10]. A recent study suggested that the menopausal transition prompts an innate immune inflammatory response in the female reproductive organs that propagates to the brain, making the brain more susceptible to ischemic damage [11]. Therefore, an understanding of the biological mechanisms of menopause transition can better equip us to lower the risk of CI and AD in women and develop strategies to protect them from menopause-associated health complications.

Understanding the biological mechanisms of menopause transition or peri-menopause is complicated [12]. Systemic evaluation is difficult in humans owing to multiple life factors such as age, parity, diet, environmental factors (e.g., toxin exposures, drug abuse), genetic background, and overlapping medical comorbidities [13]. This points to the need to conduct studies using animal models of menopause; however, as routinely used laboratory rodents do not undergo menopause, we have included some discussion on how to mimic the condition of menopause in these animals in this review. Although our understanding of menopausal transition induced physiological changes is limited, it is clear that a foundational indicator of human menopause is complete ovarian failure. Menopausal women have very low circulating levels of estrogens (estrone (E1), 17β-estradiol (E2), estriol (E3)) and progesterone but significantly elevated follicle stimulating hormone (FSH) and luteinizing hormone (LH) levels [14, 15]. It is also well known that during the premenopausal phase of a woman’s life, estrogen confers natural protection against cerebrovascular diseases. Estrogens exert beneficial effects on a myriad of body systems, including cardiovascular, bone, and brain. It has also been shown that decline in circulating E2 after menopause is associated with an increased risk for cardiovascular disease, osteoporosis, cancer, diabetes, stroke, sleep disturbances, AD, and cognitive decline [8, 16]. A study investigating a cohort of healthy women transitioning into menopause showed increase in abdominal obesity, triglycerides, total cholesterol and LDL cholesterol, fasting glucose, insulin resistance, and body mass index (BMI), and increased blood pressure [17].

Numerous pharmacological strategies of the past several decades that target menopause and associated cerebro- and cardiovascular or metabolic disorders have focused on substituting or replacing lost estrogen/+progesterone. The trials focused on primary prevention—the Women’s Health Initiative (WHI)—and secondary prevention—the Heart and Estrogen/progesterone Replacement Study (HERS) and the Women’s Estrogen for Stroke Trial (WEST) [18]. These trials indicate that postmenopausal hormone therapy is not effective for reducing the risk of a recurrent stroke or death among women with established vascular disease or for prevention of a first stroke. Similar results exist for cardiovascular disease, and even recent trials of postmenopausal hormone treatment to improve cognitive outcomes have been inconsistent [19, 20]. Later trial outcomes indicate that postmenopausal hormone use may not benefit verbal cognitive function, although current and past hormone use is associated with differences in neural pathways used while assessing verbal semantic distinctions. Overall, outcomes of these trials suggested that the hormone therapy should not be initiated to prevent vascular disease among postmenopausal women. Even trials of postmenopausal hormone treatment to improve cognitive outcomes have been inconsistent. Furthermore, therapies based on estrogen substitution have been challenged by several risks associated with treatment, including heart disease, stroke, blood clots, and breast cancer.

Although estrogen disappointed in the clinic, multiple basic science studies in the field of stroke showed beneficial effects of estrogen therapy [21,22,23,24] and therefore provided better understanding of how estrogen(s) exert beneficial effects on cerebro-, cardio-, and vascular systems. It is now known that one of the key functions of estrogen is to work as a potent anti-inflammatory factor [25,26,27], and therefore, disturbances in the cyclic pattern of circulating estrogens at the menopausal transition activate systemic innate and adaptive immune responses [28]. The inflammasome is a key element of the innate immune response [29, 30]. The inflammasome is a multiprotein complex responsible for the activation of caspase-1 and processing of pro-inflammatory cytokines such as IL-1β and IL-18 [31, 32]. As an element of the innate immune response, the inflammasome complex is a sensor of damage associated molecular patterns (DAMPs) [33]. The onset of the innate immune responses leads to activation of the adaptive immune response, which response results in infiltration of peripheral immune cells, particularly T cell invasion of the brain [34]. Ultimately repeated or sustained activation of innate and adaptive immune responses can create the chronic low-grade inflammation typical of aging. The presence of the inflammasome complex in the cerebrospinal fluid of post-menopausal women suggests that the decline in estrogens induces a pro-inflammatory state [35]. Inflammasomes could be an important indicator of the effect of menopause on the immune system. This review assesses (1) our understanding of the menopausal transition and associated estrogen decline, (2) how these impact systemic and central nervous system (CNS) inflammatory responses, and (3) possible mechanisms by which estrogen receptor(s) regulate inflammasome activation in the brain to provide protection against ischemic damage.

The menopausal transition process

According to the classification of the American Society for Reproductive Medicine’s Stages of Reproductive Aging Workshop (STRAW), a woman’s life is delineated into seven stages ranging from the onset of menstrual cycles at menarche and the reproductive age to the peri-menopausal and postmenopausal phases [1]. In the USA, approximately 1.3 million women become menopausal each year. The overall process of menopause transition lasts about 14 years. As mentioned in introduction, the average age of menopause is between 45 and 51 years, and the mean life span of women continues to increase beyond 80 years, which is about 5 years longer than that of men [36, 37]. Therefore, women are likely to spend at least one third of their life in the post-menopause stage, a stage that is vulnerable to the morbidities caused by immune and metabolic dysfunction and neurodegenerative disease.

An additional 1% of women experience premature menopause, before the age of 40 [38, 39]. Heredity appears to be the most important determinant of age at menopause [40]. Premature menopause due to permanent ovarian failure may be associated with sex chromosome abnormalities [39]. However, changes from the body’s natural fluctuating levels of estrogens through surgical removal of the ovaries or through natural menopause, have been independently linked to an altered immune profile, bone and blood vessel health, and changes to cognitive processes [41]. Apart from surgical removal of ovaries, cigarette smoking/tobacco use, exposure to environmental toxins, and malnutrition have been associated with premature menopause [39]. Premature menopause is defined by presence of amenorrhea, increased gonadotrophin levels, and estrogen deficiency earlier in life before actual age of menopause [42, 43]. Women with premature menopause are at higher risk of premature death, neurological diseases, psychosexual dysfunction, mood disorders, osteoporosis, ischemic heart disease, and infertility [39]. Additionally, environmental xenoestrogens, by-products of industrialization—Bisphenol A (BPA), bis(2-ethylhexyl)phthalate (DEHP), and di(n-butyl)phthalate (DBP)—modulate systemic estrogens and induce systemic and CNS inflammation [44]. Xenoestrogens (XEs) mimic or block the synthesis, metabolism, and transport of normal endogenous hormones, disturbing normal endocrine function [45]. Xenoestrogens exert their effects through estrogen receptor signaling, resulting in epigenetic changes [46]. Multiple reviews on the environmental chemical(s) have suggested that the exposure to these chemicals depletes the ovarian reserve, leading to impaired functioning of the ovary and a shortening of the reproductive lifespan [47, 48] and early menopause [49].

Menopausal transition and medical comorbidities

Increasingly, multiple lines of experimental and public health evidence suggest that the chronic inflammation associated with estrogen decline can potentiate immune and metabolic dysfunction and neurodegenerative disease, confounding peri-menopause and posing major health challenges for twenty-first century women. Women’s risk of cardiac arrest and stroke increases exponentially after the onset of menopause. There are three main types of stroke: transient ischemic attack, ischemic stroke, and hemorrhagic stroke. It is estimated that 87 percent of strokes are ischemic. Although men have an increased stroke risk, more women than men will experience a stroke during their lifetime because of their increased life span [50, 51]. Women account for 60% of all stroke events [2, 52]. Studies consistently show that women are more functionally impaired after stroke and are less likely to receive thrombolytic therapy with tissue plasminogen activator compared with men [52]. Given the increased stroke burden and barriers to acute stroke therapy in women, it is critical to understand risk factors unique to women so that new strategies for stroke prevention can be considered. Beyond age at natural menopause, duration of ovarian activity may be a marker of stroke risk. A recent case-control study found that a longer lifetime estrogen exposure, defined as the difference between age at menopause and age at menarche, was associated with decreased stroke risk [53]. Although the underlying biological mechanisms driving increased stroke risk in women often remain unclear, they may be dependent on decline in estrogens levels around peri-menopause.

One in five women develop AD in the seventh decade of life. Late-onset AD is the most common form of dementia, and two thirds of late-onset AD patients are women. The higher longevity is one of the explanations for the late-onset AD in women; however, increasing evidence suggests that longevity alone is not the only explanation, and there may be other underlying mechanisms. Recent multi-modality brain imaging studies have compared cognitively normal 40- to 60-year-old peri-menopausal and post-menopausal women versus age- and education-matched men. The studies indicate that as women go through menopause, multiple imaging findings indicative of AD endophenotype emerge, including reduced brain glucose metabolism in frontal cortex, increased amyloid-β (Aβ) accumulation, and gray matter and white matter loss. The patterns of brain hypometabolism correlated with measured reduction in platelet mitochondrial cytochrome oxidase (COX) activity, which suggests the emergence of AD-like bioenergetic deficits in peri- and post-menopausal women [54,55,56]. Further studies indicate that systemic inflammation and estrogen decline associated with peri-menopause can contribute to accumulation of Aβ. Direct effects of E2 on neuronal Aβ have been demonstrated, showing that E2 decreased the generation and secretion of Aβ in primary neuronal culture and that administration of estrogen in estrogen-deprived mice reversed the elevated levels of brain Aβ [57].

As mentioned above, comorbidities such as tobacco smoking cause premature menopause and aggravated systemic inflammation. Nicotine is a potent addictive agent which inhibits aromatase enzyme activity. The aromatase catalyzes the conversion of androgens into estrogens [58]. Therefore, chronic nicotine exposure reduces circulating estrogen levels and triggers premature menopause in women [59,60,61,62,63,64,65,66,67]. This epidemiological finding has been modeled in female rats, in which chronic nicotine exposure reduced endogenous E2 levels [68]. Since E2 mediates its neuroprotective effects via ligand-activated estrogen receptors (ERs) subtype alpha (ER-α) and beta (ER-β), inhibition/knockdown of either of these ERs in the brain abolishes E2-induced ischemic protection, suggesting a key role of ER-α and/or ER-β-activation [69,70,71]. Estrogen receptor-β was first reported to localize to the mitochondria in 2004 [72]. Since then, it has been shown that long-term nicotine exposure selectively decreased membrane-bound and mitochondrial ER-β but not the nuclear ER-β. In a separate study, we observed that ER-β modulates inflammasome activation in the brain. However, in this study it remained to be identified which subcellular location is responsible for ER-β’s inhibitory effect on inflammasomes [68, 73,74,75]. Since nicotine reduces membrane-bound and mitochondrial ER-β availability and increases inflammasome activation and exacerbates post-ischemic damage in the brain of female rats, ER-β located at these two subcellular sites may be playing a role in regulation of inflammasome activation [35]. It is also likely that ER-β is translocated to cytoplasm following nicotine treatment, thus reducing the presence of ER-β at these subcellular locations and resulting in increased inflammasome activation. It has been demonstrated that ERs require palmitoylation for their transport to various subcellular sites including the plasma membrane [76,77,78]. Palmitoylation is a post-translational modification that regulates membrane-protein interactions and is a reversible process [79,80,81,82]. In our unpublished study in progress, nicotine reduced ER-β palmitoylation in the hippocampus. Therefore, analyzing ER-β palmitoylation levels in the subcellular fractions and investigating the role of membrane-bound ER-β with non-permeable estrogen-conjugate may provide insight into the role of membrane-bound ER-β in inflammasome activation. In an in vitro study using organotypic slice cultures, inhibition of inflammasome activation using Isoliquiritigenin (ILG) attenuated nicotine-induced ischemic cell death after oxygen-glucose deprivation [35]. Isoliquiritigenin inhibits inflammasome activation by retarding oligomerization of Apoptosis-associated speck-like protein containing a caspase recruitment domain (ASC) and NOD-like receptor-3 (NLRP3) activation. The neuroprotective effect of ILG reflects the suppression of nicotine-induced inflammasome assembly in the brain [83, 84]. Since mitochondria play pivotal roles in initiation and regulation of the NLRP3, and NLRP3 activators induce mitochondrial destabilization, NLRP3 deubiquitination, and linear ubiquitination of ASC [85], it is likely that the presence of mitochondrial ER-β may be hindering NLRP3-mitochondrial interactions. Nicotine-induced loss of mitochondrial ER-β, therefore, provides mitochondrial access to NLRP3, leading to inflammasome oligomerization and activation; however, these hypotheses remain to be tested. These findings suggest that comorbidities like environmental exposures to toxic compounds and drug addiction also need to be taken into consideration when investigating menopause-associated inflammatory responses.

Menopause and immune responses

Estrogens are a key influence on immune and inflammatory processes, summarized graphically in Fig. 1. The role of estrogens is shown by increased inflammatory responses to infection and a higher rate of autoimmune diseases in post-menopausal women when compared to men, as well as by the variation of chronic inflammatory disease activity with the menstrual cycle, pregnancy, and menopause [86, 87]. It is very well known by now that the paucity of ovarian steroidal hormones enhances the inflammatory process predisposing menopausal women to immune disorders such as rheumatoid arthritis [88], that the pathology of multiple sclerosis worsens after menopause [89], and that post-menopausal women are more prone to stronger immune responses [90]. Additionally, studies from various laboratories have demonstrated at least a trend for increases in circulating pro-inflammatory cytokines IL-6 and TNF-alpha after natural or surgical menopause [91,92,93,94]. Deficiency of ovarian steroidal hormones potentiates the pro-inflammatory state, predisposing menopausal women to immune disorders [90, 95, 96]. The endocrine transition of the peri-menopause to the post-menopause stage, and the associated rise in chronic low-grade inflammation, is suggested to accelerate ovarian failure [10]. Declining ovarian steroidal hormones at menopausal transition coincide with higher levels of circulating interleukins IL-6, sIL-6, IL-4, IL-2, and tumor necrosis Factor (TNF) in postmenopausal women, and the latter two are also shown to reverse by hormone therapy [9, 94, 97]. Some cytokines, viz. IL-4 and IL-2 levels, are shown to increase with menopause, but the increases can be reversed by hormone therapy and are clearly suggestive of hormone dependence [9]. Synchronous to the systemic inflammatory responses, the peri-menopausal transition exhibits a decline in brain glucose metabolism and mitochondrial respiration [54, 98,99,100], myelin catabolism [101], and reduction of brain white matter volume [56]. These changes in systemic inflammation are also associated with beta-amyloid deposition in brain [56] and changes in neurological function [102]. Inflammatory signaling also changes T cell response, causing a reduction in CD4-positive T cell numbers in menopausal women and eventually inversion in the CD4/CD8 T cell ratio, which is indicative of aging and can be correlated with increased oxidative stress [103,104,105]. Overall these changes cause deterioration of the adaptive component of immune system and lead to decreased numbers of circulating B cells with menopause, especially during late menopause in comparison to peri-menopause [106]. Hormone replacement therapy retards the progress of immunosenescence by increasing the production of the CD5-B-2 subset of B cells, once more supporting the role for estrogen in the immune response [106].

Fig. 1
figure 1

At the reproductive senescence/menopausal transition the ovarian failure is associated with release of extracellular vesicles containing inflammasomes, which may be responsible for low-grade systemic inflammation. This low-grade inflammation may compromise the blood-brain barrier (BBB), making the brain more susceptible to inflammation and neurodegenerative diseases

Recapitulating human female reproductive phases in laboratory animals

Estrogen’s actions are complex, and it is often challenging to systematically evaluate the biological underpinnings associated with estrogen’s actions in menopausal women. Therefore, multiple laboratories around the globe use rodent models that are invaluable tools for studying the impact of estrogen fluctuations on a variety of body systems including brain. Animal models provide researchers with opportunities to gain a fundamental understanding of the key elements underlying reproduction and aging processes, paving the way to explore novel pathways for intervention associated with known health risks. It is essential to keep in mind that some of the mechanisms associated with aging and the transition into a reproductively senescent state can differ when translating from one species to another.

The key difference between human and rodent reproductive senescence is that rodents have an estrous cycle rather than a menstrual cycle and that rodents’ uterine lining is reabsorbed rather than shed via menstruation. The estrous cycle of rat is usually 4–5 days as against the 28-day menstrual cycle in the human female. The estrous cycle is comprised of 4 main stages, viz. proestrus, estrus, metestrus, and diestrus. The transition from the diestrus to proestrus stage represents the follicular phase of the menstrual cycle, and during this transition the estrogen levels increase. The next transition from the proestrus to estrus stage is identified by a preovulatory surge of luteinizing hormone. The gradually increasing influence of progesterone corresponds to the ovulatory phase in humans. Lastly, the transition from estrus to metestrus represents the luteal phase, and this transition is associated with higher titers of circulating progesterone. The circulating ovarian hormone concentrations are lowest at the diestrus stage, and this stage correlates to the late luteal phase and menstruation in the human female [107,108,109]. Between the age of 9 and 12 months of age, rodents typically experience irregular estrous cycles. This phase is also known as estropause, when a persistent and prolonged estrus phase may be associated with anovulatory cycles. Eventually, animals transition into an anestrous state, where ovulatory cycles cease and low levels of gonadal steroids are evident [13, 110,111,112]. The fact that moderate and persistent elevations in estrogens can occur in the rat is one difference between the rodent from human peri-menopause that can complicate the interpretation of rodent models of human reproductive phases.

A review by Galea et al. reported studies showing that the naturally occurring fluctuations in ovarian hormones across the rodent estrous cycle influence hippocampal neurogenesis in adult virgin females [113]. Adult female rats have 50% more newly proliferating cells and fewer pyknotic cells in the dentate gyrus (DG) during proestrus (the high estrogen stage) compared to male rats or adult female rats in either the estrous or diestrus stage when estradiol levels are much lower [113]. These findings from ovaries-intact rodents demonstrate mitogenic and survival effects of estrogens in the hippocampal DG [113]. Studies also showed that estrous cycle and gonadal hormonal fluctuation affect the densities of dendritic spines on rodent cortical and hippocampal pyramidal neurons [114, 115]. A study also observed differential expression of genes (DEGs) in the hippocampus of rodent depending on the stage of the estrous cycle. A transcriptome analysis of the hippocampus over the course of the four consecutive stages of the estrous cycle demonstrated that sixty-seven unique genes are differentially expressed. The majority of the differentially expressed genes occur over a single stage transition: thirty-one genes were found to change from diestrus to proestrus, five genes from proestrus to estrus, and seven genes from metestrus to diestrus. An exception to this is that there are twelve differentially expressed genes with decreased expression during the proestrus to estrus transition that are also differentially expressed during the diestrus to proestrus transition [116]. Naturally, such differentially expressed genes may have correlation with the function of the hippocampus, and the hippocampus is widely believed to be essential for learning about the context in which conditioning occurs [117]. A study assessed fear conditioning procedures on naturally cycling female rats, where cued and contextual-based fear learning were tested at the same stage of the estrous cycle, during either estrus or proestrus. Female proestrus rats showed less spatial-contextual conditioning than did male or estrous female rats. These results suggest that the changes found during the proestrus part of the cycle are related to hippocampal information processing and not to general changes in learning ability, to shock sensitivity, or to state-dependent learning [118]. Another study of similar approach demonstrated that naturally cycling ovarian hormones influence fear extinction in that elevated ovarian hormone levels during the proestrus phase appear to facilitate extinction recall [119]. This study also reported sex differences in extinction recall when accounting for cycle phase in females and suggested that the elevated fear observed in female relative to male rats during extinction recall may parallel the higher prevalence of anxiety disorders in women [119]. It is now known that exposure to stressors such as foot shock during fear conditioning paradigm leads to increased expression of multiple inflammatory factors, including the pro-inflammatory cytokine interleukin-1 (IL-1) in the brain [120, 121]. IL-1 expression dependent on endogenous estrogens may be responsible for different responses to fear conditioning.

It has been shown that these hormonal fluctuations during different stages of the normal estrous cycle influence different pathological outcomes of the focal and global CI. In stroke-prone spontaneously hypertensive female rats (SHRSPs), middle cerebral artery occlusion (MCAO) during proestrus induced infarcts that were 20% smaller compared with SHRSPs in metestrus [122]. Using a similar approach in young virgin female rats, a subsequent study from our laboratory demonstrated that the increasing milieu of circulating estrogens during the transition from diestrus to the proestrus of the estrous cycle induced ischemic protection to the hippocampal CA1 neurons [123]. In the same study it was shown that increased levels of circulating estrogens during the transition from diestrus to proestrus protects the neurons from ischemic damage, by activating the cyclic AMP response element binding (CREB) protein-mediated signaling [123]. Furthermore, the study design experimentally mimicked the endogenous variations of E2 in ovariectomized (OVX) rats by replacing E2 either at 48, 72, or 96 h intervals prior to induction of an ischemic episode. In ovariectomized rats, a single bolus of exogenous E2 replacement 48 h prior to an ischemic episode provided maximum neuroprotection via phosphorylation of CREB protein, which requires activation of ER-β [124]. Even the long-term periodic E2-treatment (every 48 h) prior to global ischemia improved cognition and reduced hippocampal neuronal loss by means of ER-β activation [124]. Direct activation of ER-β by its agonist treatments every 48 h for a month improved spatial learning, memory, and ischemic neuronal survival in OVX rats, confirming role of ER-β signaling [124]. On the contrary, silencing of hippocampal ER-β using antisense approach attenuated E2-mediated ischemic protection, suggesting that ER-β plays a key role in mediating the beneficial effects of periodic E2 treatments [125]. In support of our study, a study by Vegeto et al. demonstrated that physiological concentrations of E2 pretreatment prevent brain inflammatory responses to lipopolysaccharide, a powerful inflammatory agent [126]. E2 treatment inhibits microglia activation, as demonstrated by the lack of acquisition of the typical reactive morphology of these cells, by the impaired expression of proteins associated with phagocytosis and cell migration, and by the reduced infiltration of leukocytes through the brain parenchyma [126]. These studies motivate investigation of prophylactic ER-β replacement regimens to reduce menopause or post-menopause-associated cognitive dysfunction, ranging from subjective symptoms (“brain fog”) to measurable post-ischemic cognitive decline.

In addition to aforementioned pretreatment strategies, studies from various laboratories using a model of global or focal ischemia in OVX rats show post-ischemic E2 treatment is neuroprotective [21, 23, 127]. The activation of either/both ER-α and ER-β are proven neuroprotective against ischemic insult. However, the mechanisms of neuroprotection conferred by these receptors seem to be different. Estrogen receptor-dependent mechanisms of neuroprotection could vary depending on the experimental injury model used, sex of the animal, age of the animal, pre- or post-treatment, time of E2 administration, type of estrogens administered, the level of estrogen administrated, and the mode of administration of the steroid (see review [128]). One of the caveats of most of the aforementioned studies, including ours, is that they were performed in young OVX female rats. Ovariectomy in the young female rat mimics the condition of surgical menopause and may lack typical physiological changes in systems and CNS that are due to natural slow decline in ovarian functions. Therefore, an approach to use naturally reproductive senescence rats could be the more appropriate model to investigate effects of menopause.

At approximately a year old, the female rat transitions into an anestrous state, where ovulatory cycles halt and low levels of gonadal steroids are present [110,111,112]. This situation mimics the human menopause. Utilizing the ovary-intact rat model facilitates the evaluation of the natural age-related cellular and molecular changes in brain regions involved in normal reproductive functioning and feedback. A study from Sohrabji’s laboratory using an animal model of menopause (reproductive senescence) shows that MCAO causes a larger cortical-striatal infarct in the older, ovary-intact acyclic group compared with younger females [129]. Studies from that laboratory also showed that E2 treatment is neuroprotective in younger females, while E2 paradoxically increases infarct volume in middle-aged acyclic female rats [129]. Their study also suggested that there is an age-associated loss of Insulin-Like Growth Factor-1, a neuroprotectant that decreases with advancing age and is downregulated by E2 treatment [129]. Consistent with the findings in menopausal women, studies have demonstrated increased pro-inflammatory cytokine levels in middle-aged female rats [130]. Elevation in inflammasome proteins has been previously reported in the hippocampus of aged rats [131]. Utilizing the ovaries-intact rat model of reproductive senescence, our recent study demonstrated the ovarian release of inflammasome-containing extracellular vesicles, which reside via blood and CSF in the CNS. This study also confirmed increased expression of inflammasome complex proteins in the CSF of peri-menopausal women, providing evidence that the observed increase in inflammation occurs in both rodents and human females [11]. Therefore, employing the ovaries-intact rat model of reproductive senescence could help understand inflammatory changes in the menopausal brain. In a recently published study, we demonstrated that adoptive transfer of serum-derived extracellular vesicles (EVs) of peri-menopausal women to young female rats triggered inflammasome activation in the rat brain, suggesting that these EVs carry the factors that can prompt inflammatory responses in the body, including CNS as they can cross the BBB [11]. It has been shown that EVs carry inflammasome proteins that play a role in the pathology of brain and spinal cord injury [132] as well as stroke [133]. It has been demonstrated that high levels of pro-inflammatory interleukin 1β exist in serum and brain of reproductively senescent female rats as compared to younger females and age matched males [11]. In an unpublished study, we observed a sex difference in systemic pro-inflammatory cytokines IL-1 and 10 profiles before and at 24 h following induction of transient MCAO; these were much higher in middle-aged female compared to male rats. These systemic inflammatory mediators released after an ischemic episode could compromise the BBB and govern the overall ischemic outcome, as middle-aged females have been shown to have severe ischemic damage as compared to young female or male counter parts.

Estrogen receptor-beta, the inflammasome, and mitochondria

Numerous research studies, including ours, have established that E2 mediates ischemic neuroprotection through activation of estrogen receptor subtypes alpha (ER-α), beta (ER-β), and G Protein-Coupled Estrogen Receptor 1 (GPER-1; also known as GPR30) [124, 134,135,136]. GPER is a newly identified member of the estrogen receptor family and is shown to localize in the cerebral cortex and hippocampus, basal forebrain, thalamus, and dorsal striatum [137, 138]. Initial studies demonstrated that GPER binds with E2, leading to rapid activation of extracellular regulated kinases (ERKs) and cAMP generation [139,140,141]. Subsequently studies reported that GPER activation via G1 administration could rapidly activate PI3K-Akt and MEK-ERK, which are rapid kinase signaling pathways in the hippocampus, and exert strong neuroprotection against global cerebral ischemia (GCI) [71, 137, 142]. A recently published study shows that GPER activation upregulates interleukin-1 receptor antagonism in the hippocampus after GCI and thus limits ischemic cell death [143]. Importantly, this study suggests that GPER preserves cognitive function following GCI via enhancing the anti-inflammatory defense mechanism of neurons by upregulating interleukin-1β receptor antagonist (IL1RA) [143].

In the brain, ERα regulates reproductive neuroendocrine functions; however, ERβ plays a definitive role in a variety of neurobiological functions [144]. Utilizing subtype-selective estrogen receptor agonists has helped determine the roles for these receptors in ischemic neuroprotection. Although ER-α and ER-β share similar ligand binding domains, ER-β possesses a relative binding affinity for several steroid hormones that differs from that of ER-α [144, 145]. Propylpyrazole triol (PPT) is selective for ER-α, with a 400-fold relative binding affinity for ER-α over ER-β [146]. Diarylpropionitrile (DPN) is a subtype-selective agonist with a 70-fold greater relative binding affinity and 170-fold greater relative potency in transcription assays for ER-β than for ER-α [147, 148]. Studies have shown that both ER-α or ER-β concentration in the brain varies during aging [149, 150]. Both ER-α and ER-β decrease in the synapses of the rat hippocampal CA1 region with age but, in contrast to ER-α, the expression of ER-β is increased in response to E2 in older animals [150]. Because ER-β remains responsive to E2, understanding the mechanisms by which ER-β protects the brain from ischemic damage in reproductively senescent females could help develop future therapeutic targets. It has now been shown that periodic ER-β activation using DPN protects hippocampal neurons from ischemic cell death in reproductively senescent female rats [125]. The observed ischemic protection conferred by periodic ER-β agonist exposure reduced the inflammasome activation and decreased IL-1β proteins in the hippocampus [125]. Silencing of hippocampal ER-β using intracerebroventricular (ICV) antisense injections, increased inflammasome activation, supporting the role of ER-β in inflammasome regulation [125]. Therefore, mechanisms by which ER-β reduces post-ischemic inflammasome activation need to be further investigated.

It has been demonstrated that ER-β is localized and involved in regulation of mitochondrial function in neurons [73], summarized graphically in Fig. 2. Mitochondrial estrogen receptors play a direct role in estrogen-mediated preservation and regulation of mitochondrial structure and function [151,152,153,154,155,156,157]. The mitochondrial oxidative phosphorylation (OXPHOS) system is located in the inner mitochondrial membrane and is composed of 5 multi-subunit complexes (complexes I-V or CI-CV). Biogenesis of the mitochondrial oxidative phosphorylation system (OXPHOS) depends on both mitochondrial and nuclear genomes (see reviews [158, 159]). Estrogen receptors bind to the estrogen responsive element (EREs) located in D-loop mitochondrial DNA (mtDNA), suggesting that estrogen receptors are involved in modulation of mitochondrial gene expression [160]. In this context, a study using a human breast epithelia cell line showed that E2-stimulated increase in mRNA levels of the mtDNA-encoded genes cytochrome c oxidase subunits I and II was inhibited by ICI 182,780 (an estrogen antagonist; also known as Fulvestrant), indicating estrogen receptor dependence [155]. Another study demonstrated that cytochrome c oxidase (complex IV; CIV) subunit III mRNA levels significantly increased in the hippocampus within 3 h of E2 treatment of OVX female rats [156]. The presence of estrogen receptors in both the nuclear and mitochondrial compartments of the cell suggests regulation of mitochondrial biogenesis and function through nuclear-mitochondrial cross-talk [161, 162]. The concept of nuclear-mitochondrial cross-talk is also supported by the fact that the ER-β is provided with a mitochondrial targeting protein sequence (mTPS; aa 220-270) while ER-α lacks an mTPS [155]. In further support of the cross-talk concept, an in vitro study demonstrated shuttling of ER-β between the mitochondria and nucleus [163]. In contrast to direct regulation of mtDNA, other studies showed that mitochondrial ER-β mediates its effect through CREB phosphorylation, and, in turn, pCREB can bind directly to the D-loop of mtDNA (the control region of mtDNA) and regulates gene expression of OXPHOS subunits [164,165,166]. Our study showed that silencing of ER-β reduced nuclear and mitochondrial pCREB following E2 treatment, suggesting that the ER-β is essential for CREB phosphorylation (pCREB) at both subcellular locations [73]. Furthermore, silencing of ER-β lowered protein levels of mitochondria-encoded complex IV (CIV) subunits 1, 2, and 3 (Cox 1, 2, and 3), indicating the role of ER-β in pCREB-mediated mitochondrial OXPHOS protein expression [73].

Fig. 2
figure 2

Putative mechanism of inflammasome activation in the neuron during pre- and peri-menopause. During pre-menopause, cyclic estradiol-17β(E2) maintains expression of nuclear, membrane, and mitochondrial estrogen receptor-beta (ER-β) expression, which in turn inhibits inflammasome activation by regulating mitochondrial functions, regulating biogenesis through cyclic AMP response element binding (CREB), and by reducing mitochondrial reactive oxygen species (ROS) formation. ER-β also increases expression of anti-inflammatory protein expression and reduces pro-inflammatory proteins. Decline in circulating estradiol-17β decreases estrogen receptor-beta (ER-β), causing activation of the inflammasome by reactive oxygen species (ROS). The inflammasome activates pro-caspase-1 into caspase-1, resulting in the processing of pro-IL-1β into IL-1β. Once active, IL-1β is secreted, resulting in a spread of the inflammatory response into neighboring cells. Similarly, extracellular vesicles containing inflammasome proteins get secreted, thus also contributing to the spread of the inflammatory response. ASC, Apoptosis-Associated Speck-Like Protein Containing CARD; ER-β, estrogen receptor subtype beta; ILR, interleukin receptors; IL-1β, interleukin 1β; NLR, nod-like receptor; NF-κB, nuclear factor κB; ROS, reactive oxygen species; TLR, toll-like receptors; TNFα, tumor necrosis factor alpha

The mammalian CIV is highly complex and ER-β regulates this complex in multiple possible ways that could affect oxidative phosphorylation. The CIV is a large integral membrane protein composed of several metal prosthetic sites and 13-14 protein subunits. First, subunits of this complex are partially encoded by both mitochondrial DNA and the nuclear genome [158, 159], and the assembly process of the CIV complex is very complicated and highly regulated because of dual origin. Since the silencing of ER-β lowered protein levels of mitochondria-encoded Cox 1-3, ER-β may also regulate nuclear DNA dependent subunit expression [158]. Mitochondrial ER-β mediates its effect through CREB phosphorylation, and phosphorylated CREB can bind directly to the control region of mitochondrial DNA and regulate gene expression of mitochondrial respiratory chain protein subunits [164,165,166]. Knockdown of ER-β reduced nuclear and mitochondrial pCREB following E2 treatment in rat hippocampus, which suggests ER-β is essential for CREB phosphorylation at both subcellular locations, and the mechanism by which it must be regulating CIV subunit expression [73]. Secondly, function of CIV also depends on phosphorylation of its subunits, and the CIV subunits Cox 1 and 4 undergo phosphorylation [167,168,169]. Thirdly, CIV subunits need correct assembly, and any defects in CIV assembly and/or stability of the enzyme result in mitochondrial dysfunction [167, 169,170,171]. Although it is clear that ER-β is located on mitochondria, our understanding of its presence on outer or inner mitochondrial membrane remains limited and needs investigation. Using isolated mitochondria from female brain, our study showed that CIV activity is directly regulated by ER-β and involved in regulation of glucose metabolism in the brain [73, 172].

The mitochondria are not only the cell’s powerhouses; they integrate a large number of signal transduction pathways for a wide variety of biologically active molecules. In this context, mitochondria could be considered a cellular arsenal since they (1) enclose a potent cocktail of pro-apoptotic proteins, (2) are a major site for production of reactive oxygen species, and (3) maintain calcium homeostasis. Disturbance(s) in fine-tuning of mitochondrial functions could release these “loaded weapons” thus activating cell death pathway(s). Mitochondrial dysfunction owing to ischemia triggers the generation of mitochondrial reactive oxygen species (mitoROS). Therefore, maintaining normal mitochondrial function is crucial for cell survival. Despite numerous studies conducted to understand mechanisms of mitochondrial function, there are multiple gaps, and new areas are constantly emerging which require further investigation.

Emerging studies are showing a major role of mitochondria in regulation of innate inflammation induced by inflammasome protein NLRP3, which is a sensor for disrupted homoeostasis, including perturbed mitochondrial function [173, 174]. One of the mechanisms of NLRP3 activation supported by the most studies includes the generation of mitoROS and translocation of NLRP3 to the mitochondria, leading to the release of mitochondrial DNA (mtDNA) [31, 175,176,177]. NLRP3 activators induce mitochondrial destabilization, NLRP3 deubiquitination, linear ubiquitination of inflammasome protein ASC, and externalization or release of mitochondria-derived molecules such as mitochondrial DNA. These molecules bind to NLRP3 that is translocated to the mitochondria and activate the NLRP3 inflammasome [85]. Mitochondria are proposed to harbor NLRP3 and be able to regulate the activity of the inflammasome complex, and mitochondrial ROS can exacerbate inflammasome immunogenic signals. In contrast, activation of mitophagy reduces inflammation by clearing mitochondrial bound NLRP3 complexes [178,179,180]. It is apparent that NLRP3 plays a role in ischemic pathology, as NLRP3 knockout animals have significantly reduced infarct size and neurovascular damage after focal cerebral ischemia [181]. With respect to E2 regulation of NLRP3 inflammasome activation, E2 has been reported in one study to suppress NLRP3 inflammasome gene expression in the cerebral cortex after focal cerebral ischemia [182]. Estrogen modulation of inflammation in hippocampus and of depression- and anxiety-like behavior is ER-β dependent [183]. However, the role of ER-β in activation of the NLRP3 inflammasome in the CNS remains unknown. The fact that ER-β activation confers ischemic protection, stimulates mitochondrial functions, and inhibits inflammasome activation is suggestive of its central role in cross-talk between inflammasome and mitochondria. A better understanding of the underlying mechanisms can ultimately lead to therapeutic strategies.

Conclusion

There is increasing and compelling evidence showing that estrogen decline during the menopausal transition drives a systemic inflammatory state. This state is characterized by systemic pro-inflammatory cytokines derived from reproductive tissues, alteration in the cellular immune profile, increased availability of inflammasome proteins in the CNS, and a pro-inflammatory microenvironment which makes the brain more susceptible to ischemic and other stressors. These pro-inflammatory processes appear to compromise ER-β’s role in protecting the brain from ischemic damage and to compromise mitochondrial functions that modulate inflammasome activation. This state sets the stage for late life neurodegenerative/neurovascular disease with co-morbid cognitive dysfunction or decline. The use of ER-β-selective agonists may constitute a safer and more effective target for future therapeutic research than an ER-α agonist or E2. ER-β activation in the brain confers ischemic protection, stimulates mitochondrial functions, and inhibits inflammasome activation. ER-β agonists may be safer in that ER-β lacks the ability to stimulate the proliferation of breast or endometrial tissue. The ER-β agonist may be able to act both on the cerebro- and cardiovascular system to reduce the ischemic burden. Thus, ER-β signaling is a guide for future translational research to reduce cognitive decline and cerebral ischemia incidents and impact in post-menopausal women, while avoiding the side effects produced by chronic E2 treatment. Therefore, the model of reproductive senescence as a systemic inflammatory phase of life is crucial to understanding neurological changes that can occur in menopausal women, and to the development of novel therapeutic targets to mitigate morbidities associated with age and reproductive senescence.

Availability of data and materials

Not applicable

Abbreviations

AD:

Alzheimer’s disease

ASC:

Apoptosis-Associated Speck-Like Protein Containing CARD

BBB:

Blood brain barrier

BMI:

Body mass index

CNS:

Central nervous system

CI:

Cerebral ischemia

CIV:

mitochondrial OXPHOS system complexes I-IV

DAMPS:

Damage-associated molecular patterns

DEG:

Differentially expressed genes

DG:

Dentate gyrus

ER-α:

Estrogen receptor subtype alpha

ER-β:

Estrogen receptor subtype beta

E1:

Estrone

E2:

17β-estradiol

E3:

Estriol

ERES:

Estrogen responsive element

EV:

Extracellular vesicles

FSH:

Follicle stimulating hormone

GCI:

Global cerebral ischemia

GPER:

G protein-coupled estrogen receptor

ICV:

Intracerebroventricular

ILG:

Isoliquiritigenin

ILR:

Interleukin receptors

IL-1β:

Interleukin 1β

LH:

Luteinizing hormone

MCAO:

Middle cerebral artery occlusion

mtDNA:

Mitochondrial DNA

mTPS:

mitochondrial targeting protein sequences

NLR:

Nod-like receptor

NF-κB:

Nuclear factor κB

OXPHOS:

Oxidative phosphorylation

ROS:

Reactive oxygen species

TLR:

Toll-like receptors

TNFα:

Tumor necrosis factor alpha

XE:

Xenoestrogens

References

  1. Davis SR, et al. Menopause. Nat Rev Dis Primers. 2015;1:15004.

    Article  PubMed  Google Scholar 

  2. Lisabeth L, Bushnell C. Stroke risk in women: the role of menopause and hormone therapy. Lancet Neurol. 2012;11(1):82–91.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Stephens S, et al. Neuropsychological characteristics of mild vascular cognitive impairment and dementia after stroke. Int J Geriatr Psychiatry. 2004;19(11):1053–7.

    Article  CAS  PubMed  Google Scholar 

  4. Go AS, et al. Heart disease and stroke statistics--2014 update: a report from the American Heart Association. Circulation. 2014;129(3):e28–e292.

    PubMed  Google Scholar 

  5. Levine DA, et al. Trajectory of cognitive decline after incident stroke. JAMA. 2015;314(1):41–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Barba R, et al. Poststroke dementia : clinical features and risk factors. Stroke. 2000;31(7):1494–501.

    Article  CAS  PubMed  Google Scholar 

  7. Tatemichi TK, et al. Clinical determinants of dementia related to stroke. Ann Neurol. 1993;33(6):568–75.

    Article  CAS  PubMed  Google Scholar 

  8. Caldwell CC, Yao J, Brinton RD. Targeting the prodromal stage of Alzheimer’s disease: bioenergetic and mitochondrial opportunities. Neurotherapeutics. 2015;12(1):66–80.

    Article  CAS  PubMed  Google Scholar 

  9. Yasui T, et al. Changes in serum cytokine concentrations during the menopausal transition. Maturitas. 2007;56(4):396–403.

    Article  CAS  PubMed  Google Scholar 

  10. Agacayak E, et al. Role of inflammation and oxidative stress in the etiology of primary ovarian insufficiency. Turk J Obstet Gynecol. 2016;13(3):109–15.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Raval AP, et al. Sexual dimorphism in inflammasome-containing extracellular vesicles and the regulation of innate immunity in the brain of reproductive senescent females. Neurochem Int. 2019;127:29–37.

    Article  CAS  PubMed  Google Scholar 

  12. Santoro N. Perimenopause: from research to practice. J Women's Health (Larchmt). 2016;25(4):332–9.

    Article  Google Scholar 

  13. Koebele SV, Bimonte-Nelson HA. Modeling menopause: the utility of rodents in translational behavioral endocrinology research. Maturitas. 2016;87:5–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Rannevik G, et al. A longitudinal study of the perimenopausal transition: altered profiles of steroid and pituitary hormones, SHBG and bone mineral density. Maturitas. 1995;21(2):103–13.

    Article  CAS  PubMed  Google Scholar 

  15. Burger HG, et al. Cycle and hormone changes during perimenopause: the key role of ovarian function. Menopause. 2008;15(4 Pt 1):603–12.

    Article  PubMed  Google Scholar 

  16. Dalal PK, Agarwal M. Postmenopausal syndrome. Indian J Psychiatry. 2015;57(Suppl 2):S222–32.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Matthews KA, et al. Changes in cardiovascular risk factors during the perimenopause and postmenopause and carotid artery atherosclerosis in healthy women. Stroke. 2001;32(5):1104–11.

    Article  CAS  PubMed  Google Scholar 

  18. Brass LM. Hormone replacement therapy and stroke: clinical trials review. Stroke. 2004;35(11 Suppl 1):2644–7.

    Article  CAS  PubMed  Google Scholar 

  19. Berent-Spillson A, et al. Postmenopausal hormone treatment alters neural pathways but does not improve verbal cognitive function. Menopause. 2018;25(12):1424–31.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Pinkerton JV. Hormone therapy for postmenopausal women. N Engl J Med. 2020;382(5):446–55.

    Article  PubMed  Google Scholar 

  21. Rusa R, et al. 17beta-estradiol reduces stroke injury in estrogen-deficient female animals. Stroke. 1999;30(8):1665–70.

    Article  CAS  PubMed  Google Scholar 

  22. Wang Q, et al. Estrogen provides neuroprotection in transient forebrain ischemia through perfusion-independent mechanisms in rats. Stroke. 1999;30(3):630–7.

    Article  CAS  PubMed  Google Scholar 

  23. Dubal DB, et al. Estradiol protects against ischemic injury. J Cereb Blood Flow Metab. 1998;18(11):1253–8.

    Article  CAS  PubMed  Google Scholar 

  24. Yang SH, et al. Estradiol exerts neuroprotective effects when administered after ischemic insult. Stroke. 2000;31(3):745–9 discussion 749-50.

    Article  CAS  PubMed  Google Scholar 

  25. Cushman M, et al. Effect of postmenopausal hormones on inflammation-sensitive proteins: the Postmenopausal Estrogen/Progestin Interventions (PEPI) Study. Circulation. 1999;100(7):717–22.

    Article  CAS  PubMed  Google Scholar 

  26. Edwards BJ, Li J. Endocrinology of menopause. Periodontol. 2013;61(1):177–94.

    Article  Google Scholar 

  27. Vegeto E, Benedusi V, Maggi A. Estrogen anti-inflammatory activity in brain: a therapeutic opportunity for menopause and neurodegenerative diseases. Front Neuroendocrinol. 2008;29(4):507–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Giannoni E, et al. Estradiol and progesterone strongly inhibit the innate immune response of mononuclear cells in newborns. Infect Immun. 2011;79(7):2690–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Tomura S, et al. Effects of therapeutic hypothermia on inflammasome signaling after traumatic brain injury. J Cereb Blood Flow Metab. 2012;32(10):1939–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. de Rivero Vaccari JP, Dietrich WD, Keane RW. Activation and regulation of cellular inflammasomes: gaps in our knowledge for central nervous system injury. J Cereb Blood Flow Metab. 2014;34(3):369–75.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Lamkanfi M, Dixit VM. Mechanisms and functions of inflammasomes. Cell. 2014;157(5):1013–22.

    Article  CAS  PubMed  Google Scholar 

  32. Broz P, Dixit VM. Inflammasomes: mechanism of assembly, regulation and signalling. Nat Rev Immunol. 2016;16(7):407–20.

    Article  CAS  PubMed  Google Scholar 

  33. Land WG. The role of damage-associated molecular patterns (DAMPs) in human diseases: Part II: DAMPs as diagnostics, prognostics and therapeutics in clinical medicine. Sultan Qaboos Univ Med J. 2015;15(2):e157–70.

    PubMed  PubMed Central  Google Scholar 

  34. Unger MS, et al. Microglia prevent peripheral immune cell invasion and promote an anti-inflammatory environment in the brain of APP-PS1 transgenic mice. J Neuroinflammation. 2018;15(1):274.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. d'Adesky ND, et al. Nicotine alters estrogen receptor-beta-regulated inflammasome activity and exacerbates ischemic brain damage in female rats. Int J Mol Sci. 2018;19(5):1330.

    Article  PubMed Central  CAS  Google Scholar 

  36. Harlow SD, et al. Executive summary of the stages of reproductive aging workshop + 10: addressing the unfinished agenda of staging reproductive aging. Menopause. 2012;19(4):387–95.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Kung HC, et al. Deaths: final data for 2005. Natl Vital Stat Rep. 2008;56(10):1–120.

    PubMed  Google Scholar 

  38. Burkard T, et al. Utilization pattern of hormone therapy in UK general practice between 1996 and 2015: a descriptive study. Menopause. 2019;26(7):741–9.

    Article  PubMed  Google Scholar 

  39. Okeke T, Anyaehie U, Ezenyeaku C. Premature menopause. Ann Med Health Sci Res. 2013;3(1):90–5.

    Article  PubMed  PubMed Central  Google Scholar 

  40. de Bruin JP, et al. The role of genetic factors in age at natural menopause. Hum Reprod. 2001;16(9):2014–8.

    Article  PubMed  Google Scholar 

  41. Henderson VW. Cognitive changes after menopause: influence of estrogen. Clin Obstet Gynecol. 2008;51(3):618–26.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Jewelewicz R, Schwartz M. Premature ovarian failure. Bull N Y Acad Med. 1986;62(3):219–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Coulam CB, Adamson SC, Annegers JF. Incidence of premature ovarian failure. Obstet Gynecol. 1986;67(4):604–6.

    CAS  PubMed  Google Scholar 

  44. Park C, et al. The effects of bisphenol A, benzyl butyl phthalate, and di(2-ethylhexyl) phthalate on estrogen receptor alpha in estrogen receptor-positive cells under hypoxia. Environ Pollut. 2019;248:774–81.

    Article  CAS  PubMed  Google Scholar 

  45. Xu Z, et al. Nonmonotonic responses to low doses of xenoestrogens: a review. Environ Res. 2017;155:199–207.

    Article  CAS  PubMed  Google Scholar 

  46. Rosenfeld CS, Cooke PS. Endocrine disruption through membrane estrogen receptors and novel pathways leading to rapid toxicological and epigenetic effects. J Steroid Biochem Mol Biol. 2019;187:106–17.

    Article  CAS  PubMed  Google Scholar 

  47. Ge W, et al. Establishment and depletion of the ovarian reserve: physiology and impact of environmental chemicals. Cell Mol Life Sci. 2019;76(9):1729–46.

    Article  CAS  PubMed  Google Scholar 

  48. Gore AC. Endocrine-disrupting chemicals. JAMA Intern Med. 2016;176(11):1705–6.

    Article  PubMed  Google Scholar 

  49. Brown S. Endocrine disrupting chemicals associated with earlier menopause. Post Reprod Health. 2015;21(1):5–6.

    Article  PubMed  Google Scholar 

  50. Carandang R, et al. Trends in incidence, lifetime risk, severity, and 30-day mortality of stroke over the past 50 years. JAMA. 2006;296(24):2939–46.

    Article  CAS  PubMed  Google Scholar 

  51. Petrea RE, et al. Gender differences in stroke incidence and poststroke disability in the Framingham heart study. Stroke. 2009;40(4):1032–7.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Reeves MJ, et al. Sex differences in stroke: epidemiology, clinical presentation, medical care, and outcomes. Lancet Neurol. 2008;7(10):915–26.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Alonso de Lecinana M, et al. Risk of ischemic stroke and lifetime estrogen exposure. Neurology. 2007;68(1):33–8.

    Article  CAS  PubMed  Google Scholar 

  54. Mosconi L, et al. Perimenopause and emergence of an Alzheimer's bioenergetic phenotype in brain and periphery. PLoS One. 2017;12(10):e0185926.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Mosconi L, et al. Correction: Perimenopause and emergence of an Alzheimer's bioenergetic phenotype in brain and periphery. PLoS One. 2018;13(2):e0193314.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Mosconi L, et al. Sex differences in Alzheimer risk: brain imaging of endocrine vs chronologic aging. Neurology. 2017;89(13):1382–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Xu H, et al. Estrogen, beta-amyloid metabolism/trafficking, and Alzheimer’s disease. Ann N Y Acad Sci. 2006;1089:324–42.

    Article  CAS  PubMed  Google Scholar 

  58. Rune GM, Frotscher M. Neurosteroid synthesis in the hippocampus: role in synaptic plasticity. Neuroscience. 2005;136(3):833–42.

    Article  CAS  PubMed  Google Scholar 

  59. Barbieri RL, Gochberg J, Ryan KJ. Nicotine, cotinine, and anabasine inhibit aromatase in human trophoblast in vitro. J Clin Invest. 1986;77(6):1727–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Cassidenti DL, et al. Short-term effects of smoking on the pharmacokinetic profiles of micronized estradiol in postmenopausal women. Am J Obstet Gynecol. 1990;163(6 Pt 1):1953–60.

    Article  CAS  PubMed  Google Scholar 

  61. Cramer DW, et al. Cross-sectional and case-controlled analyses of the association between smoking and early menopause. Maturitas. 1995;22(2):79–87.

    Article  CAS  PubMed  Google Scholar 

  62. Grainge MJ, et al. Cigarette smoking, alcohol and caffeine consumption, and bone mineral density in postmenopausal women. The Nottingham EPIC Study Group. Osteoporos Int. 1998;8(4):355–63.

    Article  CAS  PubMed  Google Scholar 

  63. Greenberg G, Thompson SG, Meade TW. Relation between cigarette smoking and use of hormonal replacement therapy for menopausal symptoms. J Epidemiol Community Health. 1987;41(1):26–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Jensen J, Christiansen C, Rodbro P. Cigarette smoking, serum estrogens, and bone loss during hormone-replacement therapy early after menopause. N Engl J Med. 1985;313(16):973–5.

    Article  CAS  PubMed  Google Scholar 

  65. Michnovicz JJ, et al. Increased urinary catechol estrogen excretion in female smokers. Steroids. 1988;52(1-2):69–83.

    Article  CAS  PubMed  Google Scholar 

  66. Mueck AO, Seeger H. Smoking, estradiol metabolism and hormone replacement therapy. Curr Med Chem Cardiovasc Hematol Agents. 2005;3(1):45–54. https://doi.org/10.2174/1568016052773270.

    Article  CAS  PubMed  Google Scholar 

  67. Windham GC, et al. Cigarette smoking and effects on menstrual function. Obstet Gynecol. 1999;93(1):59–65.

    CAS  PubMed  Google Scholar 

  68. Raval AP, et al. Nicotine and estrogen synergistically exacerbate cerebral ischemic injury. Neuroscience. 2011;181:216–25.

    Article  CAS  PubMed  Google Scholar 

  69. Zhang QG, et al. Estrogen attenuates ischemic oxidative damage via an estrogen receptor alpha-mediated inhibition of NADPH oxidase activation. J Neurosci. 2009;29(44):13823–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Dubal DB, et al. Differential modulation of estrogen receptors (ERs) in ischemic brain injury: a role for ERalpha in estradiol-mediated protection against delayed cell death. Endocrinology. 2006;147(6):3076–84.

    Article  CAS  PubMed  Google Scholar 

  71. Lebesgue D, et al. Estradiol rescues neurons from global ischemia-induced cell death: multiple cellular pathways of neuroprotection. Steroids. 2009;74(7):555–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Yang SH, et al. Mitochondrial localization of estrogen receptor beta. Proc Natl Acad Sci U S A. 2004;101(12):4130–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Raval AP, et al. Synergistic inhibitory effect of nicotine plus oral contraceptive on mitochondrial complex-IV is mediated by estrogen receptor-beta in female rats. J Neurochem. 2012;121(1):157–67.

    Article  CAS  PubMed  Google Scholar 

  74. Shi FD, et al. Nicotinic attenuation of central nervous system inflammation and autoimmunity. J Immunol. 2009;182(3):1730–9.

    Article  CAS  PubMed  Google Scholar 

  75. Piao WH, et al. Nicotine and inflammatory neurological disorders. Acta Pharmacol Sin. 2009;30(6):715–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Marino M, Ascenzi P. Steroid hormone rapid signaling: the pivotal role of S-palmitoylation. IUBMB Life. 2006;58(12):716–9.

    Article  CAS  PubMed  Google Scholar 

  77. Acconcia F, et al. Palmitoylation-dependent estrogen receptor alpha membrane localization: regulation by 17beta-estradiol. Mol Biol Cell. 2005;16(1):231–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Acconcia F, et al. S-palmitoylation modulates human estrogen receptor-alpha functions. Biochem Biophys Res Commun. 2004;316(3):878–83.

    Article  CAS  PubMed  Google Scholar 

  79. Draper JM, Smith CD. Palmitoyl acyltransferase assays and inhibitors (Review). Mol Membr Biol. 2009;26(1):5–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Greaves J, Chamberlain LH. Palmitoylation-dependent protein sorting. J Cell Biol. 2007;176(3):249–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Iwanaga T, et al. Dynamic protein palmitoylation in cellular signaling. Prog Lipid Res. 2009;48(3-4):117–27.

    Article  CAS  PubMed  Google Scholar 

  82. Linder ME, Deschenes RJ. Palmitoylation: policing protein stability and traffic. Nat Rev Mol Cell Biol. 2007;8(1):74–84.

    Article  CAS  PubMed  Google Scholar 

  83. Zeng J, et al. Isoliquiritigenin alleviates early brain injury after experimental intracerebral hemorrhage via suppressing ROS- and/or NF-kappaB-mediated NLRP3 inflammasome activation by promoting Nrf2 antioxidant pathway. J Neuroinflammation. 2017;14(1):119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  84. Gomes PX, et al. Differences in vulnerability to nicotine-induced kindling between female and male periadolescent rats. Psychopharmacology. 2013;225(1):115–26.

    Article  CAS  PubMed  Google Scholar 

  85. Liu Q, et al. The role of mitochondria in NLRP3 inflammasome activation. Mol Immunol. 2018;103:115–24.

    Article  CAS  PubMed  Google Scholar 

  86. Straub RH. The complex role of estrogens in inflammation. Endocr Rev. 2007;28(5):521–74.

    Article  CAS  PubMed  Google Scholar 

  87. Sternberg EM. Neuroendocrine regulation of autoimmune/inflammatory disease. J Endocrinol. 2001;169(3):429–35.

    Article  CAS  PubMed  Google Scholar 

  88. Doran MF, et al. Trends in incidence and mortality in rheumatoid arthritis in Rochester, Minnesota, over a forty-year period. Arthritis Rheum. 2002;46(3):625–31.

    Article  PubMed  Google Scholar 

  89. Tutuncu M, et al. Onset of progressive phase is an age-dependent clinical milestone in multiple sclerosis. Mult Scler. 2013;19(2):188–98.

    Article  PubMed  Google Scholar 

  90. Sharma N, et al. Association of periodontal inflammation, systemic inflammation, and duration of menopausal years in postmenopausal women. Quintessence Int. 2018;49(2):123–31.

    PubMed  Google Scholar 

  91. Girasole G, et al. Oestrogens prevent the increase of human serum soluble interleukin-6 receptor induced by ovariectomy in vivo and decrease its release in human osteoblastic cells in vitro. Clin Endocrinol. 1999;51(6):801–7.

    Article  CAS  Google Scholar 

  92. Woodward M, et al. Associations of blood rheology and interleukin-6 with cardiovascular risk factors and prevalent cardiovascular disease. Br J Haematol. 1999;104(2):246–57.

    Article  CAS  PubMed  Google Scholar 

  93. Deswal A, et al. Cytokines and cytokine receptors in advanced heart failure: an analysis of the cytokine database from the Vesnarinone trial (VEST). Circulation. 2001;103(16):2055–9.

    Article  CAS  PubMed  Google Scholar 

  94. Pfeilschifter J, et al. Changes in proinflammatory cytokine activity after menopause. Endocr Rev. 2002;23(1):90–119.

    Article  CAS  PubMed  Google Scholar 

  95. Benedusi V, et al. A lack of ovarian function increases neuroinflammation in aged mice. Endocrinology. 2012;153(6):2777–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Kireev RA, et al. Melatonin and oestrogen treatments were able to improve neuroinflammation and apoptotic processes in dentate gyrus of old ovariectomized female rats. Age (Dordr). 2014;36(5):9707.

    Article  CAS  Google Scholar 

  97. Giuliani N, et al. Serum interleukin-6, soluble interleukin-6 receptor and soluble gp130 exhibit different patterns of age- and menopause-related changes. Exp Gerontol. 2001;36(3):547–57.

    Article  CAS  PubMed  Google Scholar 

  98. Yao J, et al. Decline in mitochondrial bioenergetics and shift to ketogenic profile in brain during reproductive senescence. Biochim Biophys Acta. 2010;1800(10):1121–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Ding F, et al. Early decline in glucose transport and metabolism precedes shift to ketogenic system in female aging and Alzheimer’s mouse brain: implication for bioenergetic intervention. PLoS One. 2013;8(11):e79977.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Mishra A, Brinton RD. Inflammation: bridging age, menopause and APOEepsilon4 genotype to Alzheimer’s disease. Front Aging Neurosci. 2018;10:312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Klosinski LP, et al. White matter lipids as a ketogenic fuel supply in aging female brain: implications for Alzheimer’s disease. EBioMedicine. 2015;2(12):1888–904.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Brinton RD, et al. Perimenopause as a neurological transition state. Nat Rev Endocrinol. 2015;11(7):393–405.

    Article  CAS  PubMed  Google Scholar 

  103. Larbi A, et al. Aging of the immune system as a prognostic factor for human longevity. Physiology. 2008;23:64–74.

    Article  CAS  PubMed  Google Scholar 

  104. Gameiro CM, Romao F, Castelo-Branco C. Menopause and aging: changes in the immune system--a review. Maturitas. 2010;67(4):316–20.

    Article  CAS  PubMed  Google Scholar 

  105. Muller GC, et al. The inverted CD4:CD8 ratio is associated with gender-related changes in oxidative stress during aging. Cell Immunol. 2015;296(2):149–54.

    Article  CAS  PubMed  Google Scholar 

  106. Kamada M, et al. B cell subsets in postmenopausal women and the effect of hormone replacement therapy. Maturitas. 2001;37(3):173–9.

    Article  CAS  PubMed  Google Scholar 

  107. Hotchkiss J, Knobil E. The physiology of reproduction, vol. 2. 2nd ed. New York: Ravan Press; 1994. p. 711–49.

    Google Scholar 

  108. Knobil E. The neuroendocrine control of the menstrual cycle. Recent Prog Horm Res. 1980;36:53–88.

    CAS  PubMed  Google Scholar 

  109. Knobil E, Neill JD. The physiology of reproduction, vol. 2. 2nd ed. New York: Ravaen Press; 1994.

    Google Scholar 

  110. Lu KH, et al. Chronological changes in sex steroid, gonadotropin and prolactin secretions in aging female rats displaying different reproductive states. Biol Reprod. 1979;21(1):193–203.

    Article  CAS  PubMed  Google Scholar 

  111. Clemens JA, Meites J. Neuroendocrine status of old constant-estrous rats. Neuroendocrinology. 1971;7(4):249–56.

    Article  CAS  PubMed  Google Scholar 

  112. Huang HH, et al. Patterns of sex steroid and gonadotropin secretion in aging female rats. Endocrinology. 1978;103(5):1855–9.

    Article  CAS  PubMed  Google Scholar 

  113. Pawluski JL, et al. Effects of steroid hormones on neurogenesis in the hippocampus of the adult female rodent during the estrous cycle, pregnancy, lactation and aging. Front Neuroendocrinol. 2009;30(3):343–57.

    Article  CAS  PubMed  Google Scholar 

  114. Woolley CS, McEwen BS. Estradiol mediates fluctuation in hippocampal synapse density during the estrous cycle in the adult rat. J Neurosci. 1992;12(7):2549–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Chen JR, et al. Gonadal hormones modulate the dendritic spine densities of primary cortical pyramidal neurons in adult female rat. Cereb Cortex. 2009;19(11):2719–27.

    Article  PubMed  Google Scholar 

  116. DiCarlo LM, Vied C, Nowakowski RS. The stability of the transcriptome during the estrous cycle in four regions of the mouse brain. J Comp Neurol. 2017;525(15):3360–87.

    Article  CAS  PubMed  Google Scholar 

  117. Gewirtz JC, McNish KA, Davis M. Is the hippocampus necessary for contextual fear conditioning? Behav Brain Res. 2000;110(1-2):83–95.

    Article  CAS  PubMed  Google Scholar 

  118. Cossio R, et al. Sex differences and estrous cycle effects on foreground contextual fear conditioning. Physiol Behav. 2016;163:305–11.

    Article  CAS  PubMed  Google Scholar 

  119. Milad MR, et al. Estrous cycle phase and gonadal hormones influence conditioned fear extinction. Neuroscience. 2009;164(3):887–95.

    Article  CAS  PubMed  Google Scholar 

  120. Arakawa K, et al. Effects of the estrous cycle and ovarian hormones on central expression of interleukin-1 evoked by stress in female rats. Neuroendocrinology. 2014;100(2-3):162–77.

    Article  CAS  PubMed  Google Scholar 

  121. DiSabato DJ, et al. Interleukin-1 receptor on hippocampal neurons drives social withdrawal and cognitive deficits after chronic social stress. Mol Psychiatry. 2020. https://doi.org/10.1038/s41380-020-0788-3.

  122. Carswell HV, et al. Investigation of estrogen status and increased stroke sensitivity on cerebral blood flow after a focal ischemic insult. J Cereb Blood Flow Metab. 2000;20(6):931–6.

    Article  CAS  PubMed  Google Scholar 

  123. Raval AP, et al. Pretreatment with a single estradiol-17beta bolus activates cyclic-AMP response element binding protein and protects CA1 neurons against global cerebral ischemia. Neuroscience. 2009;160(2):307–18.

    Article  CAS  PubMed  Google Scholar 

  124. Raval AP, et al. Periodic 17beta-estradiol pretreatment protects rat brain from cerebral ischemic damage via estrogen receptor-beta. PLoS One. 2013;8(4):e60716.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. de Rivero Vaccari JP, et al. Estrogen receptor beta signaling alters cellular inflammasomes activity after global cerebral ischemia in reproductively senescence female rats. J Neurochem. 2016;136(3):492–6.

    Article  PubMed  CAS  Google Scholar 

  126. Vegeto E, et al. Estrogen receptor-alpha mediates the brain antiinflammatory activity of estradiol. Proc Natl Acad Sci U S A. 2003;100(16):9614–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Simpkins JW, et al. Estrogens may reduce mortality and ischemic damage caused by middle cerebral artery occlusion in the female rat. J Neurosurg. 1997;87(5):724–30.

    Article  CAS  PubMed  Google Scholar 

  128. Simpkins JW, et al. Neuroprotection and estrogen receptors. Neuroendocrinology. 2012;96(2):119–30.

    Article  CAS  PubMed  Google Scholar 

  129. Selvamani A, Sohrabji F. The neurotoxic effects of estrogen on ischemic stroke in older female rats is associated with age-dependent loss of insulin-like growth factor-1. J Neurosci. 2010;30(20):6852–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Sarvari M, et al. Ovariectomy and subsequent treatment with estrogen receptor agonists tune the innate immune system of the hippocampus in middle-aged female rats. PLoS One. 2014;9(2):e88540.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  131. Mawhinney LJ, et al. Heightened inflammasome activation is linked to age-related cognitive impairment in Fischer 344 rats. BMC Neurosci. 2011;12:123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. de Rivero Vaccari JP, et al. Exosome-mediated inflammasome signaling after central nervous system injury. J Neurochem. 2016;136(Suppl 1):39–48.

    Article  PubMed  CAS  Google Scholar 

  133. Kerr N, et al. Inflammasome proteins in serum and serum-derived extracellular vesicles as biomarkers of stroke. Front Mol Neurosci. 2018;11:309.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Lu D, et al. Activation of G protein-coupled estrogen receptor 1 (GPER-1) ameliorates blood-brain barrier permeability after global cerebral ischemia in ovariectomized rats. Biochem Biophys Res Commun. 2016;477(2):209–14.

    Article  CAS  PubMed  Google Scholar 

  135. Murata T, et al. G protein-coupled estrogen receptor agonist improves cerebral microvascular function after hypoxia/reoxygenation injury in male and female rats. Stroke. 2013;44(3):779–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Noppens RR, et al. Estradiol after cardiac arrest and cardiopulmonary resuscitation is neuroprotective and mediated through estrogen receptor-beta. J Cereb Blood Flow Metab. 2009;29(2):277–86.

    Article  CAS  PubMed  Google Scholar 

  137. Tang H, et al. GPR30 mediates estrogen rapid signaling and neuroprotection. Mol Cell Endocrinol. 2014;387(1-2):52–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Alexander A, Irving AJ, Harvey J. Emerging roles for the novel estrogen-sensing receptor GPER1 in the CNS. Neuropharmacology. 2017;113(Pt B):652–60.

    Article  CAS  PubMed  Google Scholar 

  139. Prossnitz ER, Barton M. The G-protein-coupled estrogen receptor GPER in health and disease. Nat Rev Endocrinol. 2011;7(12):715–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Prossnitz ER, Barton M. Signaling, physiological functions and clinical relevance of the G protein-coupled estrogen receptor GPER. Prostaglandins Other Lipid Mediat. 2009;89(3-4):89–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Thomas P, et al. Identity of an estrogen membrane receptor coupled to a G protein in human breast cancer cells. Endocrinology. 2005;146(2):624–32.

    Article  CAS  PubMed  Google Scholar 

  142. Lebesgue D, et al. Acute administration of non-classical estrogen receptor agonists attenuates ischemia-induced hippocampal neuron loss in middle-aged female rats. PLoS One. 2010;5(1):e8642.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  143. Bai N, et al. G-protein-coupled estrogen receptor activation upregulates interleukin-1 receptor antagonist in the hippocampus after global cerebral ischemia: implications for neuronal self-defense. J Neuroinflammation. 2020;17(1):45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Weiser MJ, Foradori CD, Handa RJ. Estrogen receptor beta in the brain: from form to function. Brain Res Rev. 2008;57(2):309–20.

    Article  CAS  PubMed  Google Scholar 

  145. Kuiper GG, et al. Interaction of estrogenic chemicals and phytoestrogens with estrogen receptor beta. Endocrinology. 1998;139(10):4252–63.

    Article  CAS  PubMed  Google Scholar 

  146. Stauffer SR, et al. Pyrazole ligands: structure-affinity/activity relationships and estrogen receptor-alpha-selective agonists. J Med Chem. 2000;43(26):4934–47.

    Article  CAS  PubMed  Google Scholar 

  147. Meyers MJ, et al. Estrogen receptor-beta potency-selective ligands: structure-activity relationship studies of diarylpropionitriles and their acetylene and polar analogues. J Med Chem. 2001;44(24):4230–51.

    Article  CAS  PubMed  Google Scholar 

  148. Sun J, et al. Molecular basis for the subtype discrimination of the estrogen receptor-beta-selective ligand, diarylpropionitrile. Mol Endocrinol. 2003;17(2):247–58.

    Article  CAS  PubMed  Google Scholar 

  149. Nilsson S, Koehler KF, Gustafsson JA. Development of subtype-selective oestrogen receptor-based therapeutics. Nat Rev Drug Discov. 2011;10(10):778–92.

    Article  CAS  PubMed  Google Scholar 

  150. Waters EM, et al. Estrogen and aging affect the synaptic distribution of estrogen receptor beta-immunoreactivity in the CA1 region of female rat hippocampus. Brain Res. 2011;1379:86–97.

    Article  CAS  PubMed  Google Scholar 

  151. Mirebeau-Prunier D, et al. Estrogen-related receptor alpha and PGC-1-related coactivator constitute a novel complex mediating the biogenesis of functional mitochondria. FEBS J. 2010;227:713–25.

    Article  CAS  Google Scholar 

  152. Simpkins JW, et al. Mitochondrial mechanisms of estrogen neuroprotection. Biochim Biophys Acta. 2010;1800(10):1113–20.

    Article  CAS  PubMed  Google Scholar 

  153. Jayachandran M, et al. Loss of estrogen receptor beta decreases mitochondrial energetic potential and increases thrombogenicity of platelets in aged female mice. Age. 2009;32(1):109–21.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. Klinge CM. Estrogenic control of mitochondrial function and biogenesis. J Cell Biochem. 2008;105(6):1342–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Chen JQ, Yager JD. Estrogen’s effects on mitochondrial gene expression: mechanisms and potential contributions to estrogen carcinogenesis. Ann N Y Acad Sci. 2004;1028:258–72.

    Article  CAS  PubMed  Google Scholar 

  156. Bettini E, Maggi A. Estrogen induction of cytochrome c oxidase subunit III in rat hippocampus. J Neurochem. 1992;58(5):1923–9.

    Article  CAS  PubMed  Google Scholar 

  157. Razmara A, et al. Estrogen suppresses brain mitochondrial oxidative stress in female and male rats. Brain Res. 2007;1176:71–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Moraes CT, et al. Mitochondrial DNA structure and function. Int Rev Neurobiol. 2002;53:3–23.

    Article  CAS  PubMed  Google Scholar 

  159. Diaz F, Moraes CT. Mitochondrial biogenesis and turnover. Cell Calcium. 2008;44(1):24–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Chen JQ, Brown TR, Yager JD. Mechanisms of hormone carcinogenesis: evolution of views, role of mitochondria. Adv Exp Med Biol. 2008;630:1–18.

    Article  CAS  PubMed  Google Scholar 

  161. Cannino G, Di Liegro CM, Rinaldi AM. Nuclear-mitochondrial interaction. Mitochondrion. 2007;7(6):359–66.

    Article  CAS  PubMed  Google Scholar 

  162. Simpkins JW, et al. Estrogen actions on mitochondria--physiological and pathological implications. Mol Cell Endocrinol. 2008;290(1-2):51–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Mattingly KA, et al. Estradiol stimulates transcription of nuclear respiratory factor-1 and increases mitochondrial biogenesis. Mol Endocrinol. 2008;22(3):609–22.

    Article  CAS  PubMed  Google Scholar 

  164. De Rasmo D, et al. cAMP/Ca2+ response element-binding protein plays a central role in the biogenesis of respiratory chain proteins in mammalian cells. IUBMB Life. 2010;62(6):447–52.

    PubMed  Google Scholar 

  165. Lee J, et al. Mitochondrial cyclic AMP response element-binding protein (CREB) mediates mitochondrial gene expression and neuronal survival. J Biol Chem. 2005;280(49):40398–401.

    Article  CAS  PubMed  Google Scholar 

  166. Ryu H, et al. Antioxidants modulate mitochondrial PKA and increase CREB binding to D-loop DNA of the mitochondrial genome in neurons. Proc Natl Acad Sci U S A. 2005;102(39):13915–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Acin-Perez R, et al. Protein phosphorylation and prevention of cytochrome oxidase inhibition by ATP: coupled mechanisms of energy metabolism regulation. Cell Metab. 2011;13(6):712–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Lazarou M, et al. Assembly of nuclear DNA-encoded subunits into mitochondrial complex IV, and their preferential integration into supercomplex forms in patient mitochondria. FEBS J. 2009;276(22):6701–13.

    Article  CAS  PubMed  Google Scholar 

  169. Dave KR, et al. Ischemic preconditioning targets the respiration of synaptic mitochondria via protein kinase C epsilon. J Neurosci. 2008;28(16):4172–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Ogbi M, et al. Cytochrome c oxidase subunit IV as a marker of protein kinase Cepsilon function in neonatal cardiac myocytes: implications for cytochrome c oxidase activity. Biochem J. 2004;382(Pt 3):923–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Guo D, et al. Protein kinase C-epsilon coimmunoprecipitates with cytochrome oxidase subunit IV and is associated with improved cytochrome-c oxidase activity and cardioprotection. Am J Physiol Heart Circ Physiol. 2007;293(4):H2219–30.

    Article  CAS  PubMed  Google Scholar 

  172. Brinton RD. Estrogen regulation of glucose metabolism and mitochondrial function: therapeutic implications for prevention of Alzheimer’s disease. Adv Drug Deliv Rev. 2008;60(13-14):1504–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Gross CJ, et al. K(+) Efflux-independent NLRP3 inflammasome activation by small molecules targeting mitochondria. Immunity. 2016;45(4):761–73.

    Article  CAS  PubMed  Google Scholar 

  174. Rodriguez-Nuevo A, Zorzano A. The sensing of mitochondrial DAMPs by non-immune cells. Cell Stress. 2019;3(6):195–207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Guo H, Callaway JB, Ting JP. Inflammasomes: mechanism of action, role in disease, and therapeutics. Nat Med. 2015;21(7):677–87.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Vanaja SK, Rathinam VA, Fitzgerald KA. Mechanisms of inflammasome activation: recent advances and novel insights. Trends Cell Biol. 2015;25(5):308–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Sutterwala FS, Haasken S, Cassel SL. Mechanism of NLRP3 inflammasome activation. Ann N Y Acad Sci. 2014;1319:82–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Zhou R, et al. A role for mitochondria in NLRP3 inflammasome activation. Nature. 2011;469(7329):221–5.

    Article  CAS  PubMed  Google Scholar 

  179. Minutoli L, et al. ROS-mediated NLRP3 inflammasome activation in brain, heart, kidney, and testis ischemia/reperfusion injury. Oxidative Med Cell Longev. 2016;2016:2183026.

    Article  CAS  Google Scholar 

  180. Yabal M, et al. Stressing out the mitochondria: mechanistic insights into NLRP3 inflammasome activation. J Leukoc Biol. 2019;105(2):377–99.

    Article  CAS  PubMed  Google Scholar 

  181. Yang F, et al. NLRP3 deficiency ameliorates neurovascular damage in experimental ischemic stroke. J Cereb Blood Flow Metab. 2014;34(4):660–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Slowik A, Beyer C. Inflammasomes are neuroprotective targets for sex steroids. J Steroid Biochem Mol Biol. 2015;153:135–43.

    Article  CAS  PubMed  Google Scholar 

  183. Xu Y, et al. NLRP3 inflammasome activation mediates estrogen deficiency-induced depression- and anxiety-like behavior and hippocampal inflammation in mice. Brain Behav Immun. 2016;56:175–86.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We sincerely thank Dr. Thiruma V. Arumugam, Professor of Physiology, Anatomy & Microbiology at the School of Life Sciences, La Trobe University, Bundoora, Victoria, Australia, for the critical review of our manuscript.

Funding

This work was supported by an endowment from Drs. Chantal and Peritz Scheinberg (Ami P. Raval), and by Florida Department of Health # 20K09 (Ami P. Raval).

Author information

Authors and Affiliations

Authors

Contributions

Drs. McCarthy and Raval conceived the scientific idea and discussed outline of the review. Dr. Raval wrote the review and Dr. McCarthy provided input in the writing of the manuscript. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Ami P. Raval.

Ethics declarations

Ethics approval and consent to participate

All institutional and national guidelines for the care and use of laboratory animals were followed. Animal usage and experimentation were approved by the Institutional Animal Care and Use Committee at the University of Miami and was in accordance with the US Public Health Service’s Policy on Humane Care and Use of Laboratory Animals.

Consent for publication

Not applicable

Competing interests

None

Additional information

Publisher’s Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

McCarthy, M., Raval, A.P. The peri-menopause in a woman’s life: a systemic inflammatory phase that enables later neurodegenerative disease. J Neuroinflammation 17, 317 (2020). https://doi.org/10.1186/s12974-020-01998-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12974-020-01998-9

Keywords