Abstract

1. Introduction

As expertly described in the accompanying review articles, sex steroid hormones are potent regulators of neuron survival in multiple CNS regions and across a variety of circumstances ranging from normal development to neural injury. A compelling, and as yet largely unrealized, promise of sex steroid hormones is the translation of their neuroprotective properties into efficacious strategies for the treatment and or prevention of age-related neurodegenerative disorders such as Alzheimer’s disease (AD). Despite this unfulfilled therapeutic potential, abundant experimental, epidemiological and clinical evidence suggest that neural actions androgens, estrogens, and perhaps even progestogens can reduce the risk for AD.


AD is an age-related neurological disease that is the leading cause of dementia. Neuropathologically, AD is characterized by brain region-specific deposition of β-amyloid protein (Aβ) which creates senile plaques, hyperphosphorylation of the cytoskeletal protein tau that forms lesions called neurofibrillary tangles and neuropil threads, glial activation which is associated with inflammatory responses, and both synaptic and neuronal loss [4, 38, 140, 141, 197]. Although the mechanisms of AD pathogenesis remain to be fully resolved, the leading hypothesis posits that the disease is initiated and driven by prolonged elevation of Aβ levels [141]. Aβ is a proteolytic byproduct of the metabolism of amyloid precursor protein, a widely expressed protein with numerous functions ranging from axonal transport to gene transcription [336]. As a consequence of amyloid precursor protein expression, Aβ is normally found as a soluble protein at low levels in fluids and tissues throughout the body. In theory, alterations in either the production or clearance of Aβ that sway Aβ homeostasis towards increased neural levels will promote the development of AD [141]. The accumulation of Aβ encourages its abnormal assembly into oligomeric species that exhibit an altered structural conformation and can induce a range of neurodegenerative effects [137]. Consequently, enormous effort has been expended on identifying factors that regulate Aβ accumulation and or affect its neurodegenerative properties. One such class of factors is sex steroid hormones.


In this review, we will discuss the neuroprotective properties of sex steroid hormones as they relate to AD pathogenesis, focusing largely on their effects on Aβ accumulation and its associated neurodegeneration. Estrogens are the most thoroughly studied steroid hormones in terms of AD. We will cover the epidemiological and clinical evidence that suggests depletion of ovarian hormones at menopause increases the risk of AD in postmenopausal women, a danger some studies suggest may be mitigated by estrogen-based hormone therapy (HT). Consistent with a protective role against AD, experimental studies demonstrate that estrogens not only reduce neuron loss induced by AD-related insults but also act to reduce Aβ levels. However, in the Women’s Health Initiative (WHI) trial, the most exhaustive clinical evaluation of HT thus far, HT was associated with increased rather than decreased risk of AD. Analysis of experimental work reveals several key limitations of estrogen’s neuroprotective actions that may contribute to this clinical observation, including loss of neural estrogen responsiveness with age and interactions with progestogens that may limit estrogen neuroprotection. A more recent and still emerging literature suggests a parallel relationship in men with their primary sex steroid hormone, testosterone. That is, normal age-related testosterone is associated with increased risk of AD in aging men. Like estrogens, androgens also exert neuroprotective properties relevant to AD, including promotion of survival in neurons challenged with AD-related insults and reduction of Aβ levels. Finally, we consider future directions in this field, emphasizing the clinical potential of sex steroid hormones in prevention rather than treatment of AD and the emerging promise of selective estrogen receptor and androgen receptor modulators.


2. Menopause, hormone therapy, and Alzheimer’s disease

Converging lines of evidence indicate a potentially important role of estrogens in regulating AD pathogenesis. Preliminary clues suggesting this possibility stemmed from reports of sex differences in AD risk, with women showing higher prevalence and incidence. Although sex differences in AD are difficult to interpret due to gender differences in life expectancy, many studies of various cohorts indicate that women are at greater risk of AD [10, 18, 40, 94, 102, 138, 181, 222, 281, 290]. Further, there is some evidence that AD pathogenesis may be more severe in women as indicated by sex differences in cognitive deficits and neuropathology [22, 50, 67, 150], although other studies indicate men have higher levels of tau pathology [293, 294]. Further, there is a stronger association between the apolipoprotein E ε4 allele and sporadic AD in women compared to men [67, 85, 166] and the ε4 allele has been shown to be associated with greater hippocampal atrophy and memory impairments in women compared to men [99]. When considered together, these epidemiological and neuropathological studies indicate sex differences in AD, suggesting that women may be more vulnerable to AD than men.


Several transgenic mouse models exhibit sex differences in AD-like neuropathology that appears to parallel that observed in human AD cases. For example, at both 15 and 19 mo of age, female Tg2576 mice display a higher plaque load burden and higher levels of both soluble and insoluble Aβ40 and Aβ42 than age-matched males [52]. Similarly, female APPswexPS1 transgenic mice have higher Aβ load burden and plaque number than age-matched males [349]. The same pattern of greater Aβ deposition in female versus male mice is also observed in the 3xTg-AD triple transgenic mouse [157]. These studies in transgenic mouse models of AD suggest that the female brain may be more vulnerable to AD pathogenesis.


The increased risk of AD in women is presumed to be associated with the precipitous loss of estrogens and progesterone at menopause. Consistent with this position, plasma levels of the estrogen 17β-estradiol (E2) [213] are reported to be lower in women with AD in comparison to age-matched controls. If the depletion of ovarian hormones at menopause contributes to women’s increased risk of AD, then one would predict that estrogen-based hormone therapy (HT) would be effective in the prevention and or treatment of AD. This critical issue remains unresolved with persuasive arguments both for and against the use of HT for AD.An early study of this issue found that AD risk was lower in women who used HT relative to nonusers and this risk decreased significantly as both dose and duration of HT use increased [247]. Similarly, findings from several other case control and prospective studies suggest that postmenopausal women with a history of HT use were at reduced risk of AD [53, 155, 185, 248, 329, 351, 382]. Further, a meta-analysis of studies found that HT was associated with decreased risk of cognitive dysfunction [159, 198, 242, 306]. Collectively, these studies suggest a potential protective role of estrogen against the development of AD.


Despite indications of benefits, the potential protective of HT against AD remains controversial. Arguing against a protective role, several studies found that HT use was not associated with reduced risk of AD (reviewed in [147]) or failed to yield significant cognitive benefits [8, 23, 35, 41, 120, 154, 229, 266, 350, 370]. One possible explanation for this discrepancy is suggested by findings from the Cache County Study, which demonstrated that the association between HT use and reduced risk of AD was strengthened in long-term HT users [382]. Interpretations of these findings include the concept that HT may have a largely preventive role against AD and or the hypothesis that early initiation of HT is essential as women who took HT for longer periods likely began treatment nearer the time of menopause.


The notion that estrogen-based HT can effectively reduce the risk of AD and improve age-related deficits in cognition has been challenged by findings from the Women’s Health Initiative Memory Study (WHIMS). WHIMS was a randomized, multi-center, double-blind, placebo-controlled study of ∼4500 women between 65–79 yrs of age that evaluated effects of HT consisting of conjugated equine estrogen (CEE) alone or CEE with the progestin medroxyprogesterone acetate (MPA). This study reported that neither CEE nor CEE+MPA significantly improved cognition versus placebo in women showing cognitive decline associated with normal aging [92, 273] or dementia [311, 312]. In the CEE alone arm, there was no significant difference in dementia incidence between HT or placebo groups although there was a non-significant trend towards increased risk in the HT group [92, 311]. In the CEE+MPA arm, they found that women receiving HT had a higher risk of probable dementia [312]. HT use was also associated with increased incidence of stroke and breast cancer, suggesting that the risks of HT may outweigh its benefits. Further, the important differences between the CEE alone arm and the CEE+MPA arm raise several issues about the inclusion of a progestin component.


Although the WHIMS findings raise serious concerns over HT use, many challenge the interpretation that these data dismiss the potential efficacy of HT in reducing the risk of AD [76, 119, 151, 210, 265, 276]. A variety of issues have been identified that may have affected HT outcomes in the WHIMS compared to the many observational studies such as differences in methodological techniques, outcome measures, hormone exposures, menopausal symptoms, and the timing of hormone use [151]. In particular, neural sensitivity to sex steroid hormones may diminish during the menopause transition, resulting in a critical window in which to initiate HT in order to realize benefits (reviewed in [76]). Because HT was initiated many years after menopause in the WHIMS study, the study’s design may have inadvertently focused on an age group in which estrogen is minimally active in brain and thus was unlikely to detect potential cognitive benefits. In addition, as suggested by prior epidemiological and clinical findings, estrogen may be most effective in preventing rather than treating AD. In this case, the relatively advanced age of WHIMS subjects would also be biased against positive outcomes. Additional issues include HT formulation, route of administration, and treatment regime (e.g., cyclic versus continuous hormone delivery). In order to unravel this conundrum, a greater understanding of the neuroprotective actions of estrogens and progestogens are needed, as well as their limitations in the context of aging.

In the following sections, we examine the neuroprotective effects of estrogen, focusing on its abilities to increase neuronal resilience against AD insults and to antagonize AD pathogenesis by reducing Aβ accumulation. Importantly, we also discuss experimental evidence that addresses how these protective actions of estrogen are affected by the concerns raised by WHIMS.


3. Estrogen neuroprotection and Alzheimer’s disease insults

An established neural action of estrogens that may contribute to a protective role against AD is promotion of neuron viability. Estrogen is neuroprotective against a variety of insults in several cell culture and rodent paradigms of injury and neurodegenerative disease. Of particular interest is the ability of estrogen to protect against neuronal loss induced by Aβ, which is thought to be primary neurodegenerative agent in AD. Reports from several groups demonstrate that estrogen can protect cultured neurons and neural cell lines from Aβ mediated toxicity [28, 124, 134, 136, 223, 259].


Estrogen may potentially protect against Aβ-induced neurotoxicity at several steps in the degenerative process. The leading theory of Aβ toxicity posits a pathologic assembly of Aβ involving adoption of a β-sheet conformation, resulting in a change in protein structure that is associated with a toxic gain of function [262], (reviewed [347]). Consistent with this working hypothesis of Aβ neurotoxicity, our prior work has shown that Aβ is toxic only in an assembled state [261, 264]. Assembled Aβ in the form of soluble oligomers and insoluble fibrils can induce neuronal death [74, 348, 373] degeneration of neurites [170, 262] and synaptic dysruption [146, 183, 288, 348] leading to impaired learning and memory [62, 199]. Interaction of Aβ assemblies with neurons initiates a cascade of upstream signaling mechanisms associated with cell death, including calcium dysregulation [217, 356], oxidative stress [26, 125, 215], and activation of pro-inflammatory pathways promoting chronic gliosis [90, 263]. Most evidence suggests that the plethora of upstream signaling cascades elicited by Aβ ultimately mediate neurotoxicity by downstream activation of neuronal apoptosis pathways [72, 73]. In particular, we find that Aβ-induced neuronal apoptosis involves activation of JNK signaling and consequent dysregulation of the Bcl-2 family of apoptosis-related proteins [374].


Since neuronal apoptosis is an important downstream mediator of Aβ neurotoxicity, regulation of apoptosis is predicted to be a key mechanism of estrogen protection from Aβ. Consistent with this hypothesis, estrogen has been implicated in the regulation of Bcl-2 family members in neurons [86, 111, 192, 236, 254, 259, 316, 320]. The Bcl-2 family includes both proteins that promote cell survival (e.g., Bcl-2, Bcl-xL, and Bcl-w) and others that antagonize it (e.g., Bax, Bad, Bak, Bik, Bid, BNIP3, and Bim) (reviewed in [13, 71]). We have found that physiological levels of E2 inhibit neuronal apoptosis at least in part by increasing expression of anti-apoptotic Bcl-xL [259, 320] and Bcl-w [375] while down-regulating expression of the pro-apoptotic Bim [375]. The observed effects of E2 on Aβ mediated apoptosis were found to be mediated ER dependent mechanisms, since the anti-apoptotic effects of E2 were blocked by pre-treatment with an ER antagonist [259, Yao, 2007 #3161, 375] Supporting this, it has been previously demonstrated that both ERα and ERβ are crucial in regulating Bcl-2 expression and neuronal survival [389] and recently it has been shown that E2can also increase Bcl-2 through Akt-dependent CREB activation [381].


Interestingly, estrogen dependent regulation of the Bcl-2 family of proteins has also been implicated in neuroprotection against excitotoxicity, a form of neuronal injury implicated in AD neurodegeneration (reviewed in [30, 280]). Evidence suggests that in AD, Aβ toxicity and glutamate excitotoxicity may cooperatively activate pathways leading neuronal death. Glutamate-induced excitotoxic injury is potentiated by Aβ. In cell culture paradigms, the combination of sub-lethal concentrations of glutamate combined with sub-lethal levels Aβ yields robust neuronal loss [190, 216]. Such degenerative interactions are predicted to occur in AD because the AD brain exhibits both Aβ accumulation and evidence of glutamate injury. Whether upstream and or downstream pathways of Aβ and glutamate action are responsible for their synergistic toxic effects is unclear. Glutamate excitotoxicity leads to calcium dysregulation and oxidative stress, and excitotoxic neuron death is mediated in part by apoptosis (reviewed in [65]).


Several studies have demonstrated that estrogen reduces excitotoxic neuronal death induced by glutamate agonists in cell culture [44, 235, 275, 315, 317, 319]. For example, Dorsa and colleagues found that estrogen inhibited neuronal death in murine cortical cultures following excitotoxic insult, an effect that could be pharmacologically blocked with the ER antagonist, tamoxifen [317]. Further, Brinton and collegues found estrogen to promote intracellular Ca2+ accumulation in neuronal cultures treated glutamate at physiological doses, while inhibiting intracellular Ca2+ accumulation following treatment with excitotoxic glutamate doses [235].Similar observations of estrogen neuroprotection following excitotoxic challenge have been reported by several groups [124, 275, 354].


Estrogen has also been shown to regulate the extent of excitotoxic injury in rodent models [15, 16, 56, 287]. For example, administration of exogenous E2 to ovariectomized (OVX) rats has been reported to protect against kainate-induced neuronal loss [16]. Additionally, depletion of endogenous estrogen levels may increase susceptibility to excitotoxicity, with pronounced kainate-induced neural loss observed in intact rats during proestrus or following OVX [15]. Similarly, we observe neuroprotection against kainate lesion following administration of estrogen to OVX rats [56, 287].


Interestingly, estrogen neuroprotection against excitotoxic injury shares mechanistic similarities with protection against Aβ-induced apoptosis. That is, estrogen regulation of the Bcl-2 family is implicated in protective actions against glutamate-related injury [234, 236, 316, 389].Brinton and colleagues found that estrogen mediated neuroprotection against glutamate excitotoxicity by promoting mitochondrial Ca2+sequestration, and this was associated with increased expression of Bcl-2 [234]. Estrogen dependent modulation Bcl-2 following excitotoxicity may be mediated by rapid, non-genomic ER-dependent signaling mechanisms [387, 389]. Estrogen activates an ER-dependent Src/ERK/CREB signaling pathway that leads to upreguation of Bcl-2 [363]. In contrast, others suggest that estrogen may regulate Bcl-2 family expression through a direct genomic mechanism. Supporting this we have described an estrogen responsive element (ERE) on the Bcl-x gene [259], while others describe EREs on Bcl-2 [84, 256]. These pathways are summarized in Figure 1.


Estrogens activate neuroprotective pathways that may attenuate Alzheimer’s disease. Estrogens including 17β-estradiol (E2) reduce neuronal apoptosis by (i) non-genomic signaling cascades, including activation of PI3K, protein kinase C ...


In addition to regulation of the Bcl-2 family of proteins, estrogen has also been implicated in neuroprotection against many other prominent features of the AD-neurodegenerative cascade including inflammation and oxidative stress. Many studies indicate that Aβ may be contribute to AD-related oxidative stress [26, 125, 215] and deposition of Aβ may activate microglia and promote inflammation [291, 338]. Abundant evidence indicates that estrogen is a potent inhibitor of oxidative damage [340] and hydrogen peroxide mediated neuronal death [27, 28, 124]. Estrogen mediates these antioxidant effects through ER-independent mechanisms, acting as a free radical scavenger owing to its phenolic structure [278, 279, 324, 325]. However, these neuroprotective effects are only elicited at supraphysiological estrogen dosages [27, 203], thereby limiting the clinical relevance of these neuroprotective actions of estrogens. In contrast, the anti-inflammatory actions of estrogen are mediated by ER-dependent mechanisms and may provide preventative or therapeutic benefit to a range of neurodegenerative diseases where neuroinflammation is a major degenerative process. Pluripotent effects of estrogen have been described on glial function and activation, Estrogen may both suppress reactive gliotic responses associated with traumatic injury and neurodegeneration [110, 112], while also promoting the neuroprotective properties of astrocytes by increasing arborization and promoting synaptogenesis [81, 323]. Interestingly, estrogen has been found to attenuate microglial activation only when estrogen treatment was given to the cultures prior to inflammatory insult, indicating estrogen does not have the capacity to modulate inflammatory reactions once microglial activation has been initiated [341]. This suggests the anti-inflammatory benefits of estrogen may be limited to prevention of AD rather than treatment.


Another potentially significant neuroprotective action of estrogen that is highly relevant to AD and related neurodegenerative disorders is inhibition of pathological tau hyperphosphorylation. Both estrogen and progesterone can modulate activities of kinases and phosphatases involved in regulating levels of tau phosphorylation. Specifically, E2 and P4 regulate tau phosphorylation through the glycogen synthase kinase-3β (GSK-3β) pathway [9, 123]. Estrogen can reduce GSK-3β activity [123], and progesterone can decrease expression of both tau and GSK-3β. Estrogen can also lower tau hyperphosphorylation through the wnt signaling pathway and a gene called dickkopf-1 [384] as well as through the protein kinase A pathway [205]. These results demonstrate some of the first insights into the mechanism behind estrogen neuroprotection in tau-related disorders.


The described neuroprotective actions of estrogen against AD-related insults are largely mediated by activation of estrogen receptors (ER). It has been well established that both ER subtypes are widely distributed in the brain, including in brain regions affected in AD such as the hippocampus, frontal cortex, and amygdala [145, 309, 310]. Recently, changes in the subcellular distribution of ERs in hippocampal neurons have been implicated in AD pathogenesis [207]. Specifically, the shift of ERα from the nucleus to the cytoplasm may decrease the development of AD pathology in humans [156] and transgenic mice [182]. In addition, recent studies have demonstrated that ERα levels in the frontal cortex correlated with mini-mental state examination scores in women with end-stage AD [186] and that specific allele differences in ERα are correlated with an increased risk for AD in women with Down syndrome [296]. Also, ERβ immunoreactivity is reportedly increased in the hippocampus compared to age-matched controls [292]. Taken together, these studies suggest that the expression of ERα/β in the AD brain may play an integral role in the neuroprotective actions of E2. These effects are discussed in several recent reviews [39, 241].


Both ER subtypes ERα and ERβ are implicated in mediating estrogen neuroprotection, although their relative contributions have been incompletely defined. Selective expression of ERα versus ERβ in neural cell lines has suggested a more important role of ERα in mediating neuroprotection in some studies [188] but significant contributions from both ERα and ERβ in others [98, 204, 219]. Cell culture studies utilizing selective ERα (propylpyrazole triol, PPT), and ERβ agonists (diarylpropionitrile; DPN) to study estrogen neuroprotection typically report similar levels of protection from both agonists, although some evidence suggests greater activity of the ERα agonist PPT [29]. Our studies in primary neuron culture indicate comparable levels of neuroprotection against Aβ from E2, PPT, and DPN, suggesting potential contributions from both ERα and ERβ in estrogen neuroprotection [70]. However, our data also suggested potential differences between ER subtypes in terms of protective mechanisms, with PPT but not DPN inducing PKC-dependent neuroprotection [70]. Similarly, Brinton and colleagues find that PPT and DPN closely mimic E2protection from glutamate excitotoxicity, including activation of ERK signaling and upregulation of Bcl-2 expression [387, 389]. They also reported differences between the agonists in which DPN effects showed greater calcium dependence [387]. Collectively, available evidence suggests that both ERα and ERβ likely contribute to neuroprotection against AD-related insults but that each may mediate protection by preferentially activate different signaling pathways.


4. Estrogen regulation of β-amyloid accumulation

In addition to increasing neuronal resistance to AD-related insults, estrogen may also protect against AD by preventing the key initiator of AD pathogenesis, accumulation of Aβ. Steady state levels of Aβ are influenced by opposing pathways of Aβ production and Aβ clearance, both of which appear to be regulated by estrogen. Estrogen regulation of Aβ was first suggested by cell culture experiments focused on Aβ production. Early studies demonstrated that estrogen modulates processing of amyloid precursor protein (APP), the transmembrane parent protein of Aβ [108].

The majority of APP is metabolized by two competing pathways, the amyloidogenic and non-amyloidogenic pathways. In the amyloidogenic pathway, thought to occur following endocytosis of cell-surface APP, APP is first cleaved by β-secretase (BACE) to liberate β-APPs. The C-terminal fragment (C99/β-CTF) is left embedded in the membrane and is cleaved by the γ-secretase enzyme liberating the Aβ40/Aβ42 peptides. It is thought that another fragment is also released termed the APP intracellular domain (AICD), which can translocate to the nucleus and activate gene transcription. In the non-amyloidogenic pathway, which is the predominant pathway, APP is cleaved within the Aβ domain by a-secretase to liberate a neuroprotective, secreted form of APP (α-APPs). A C-terminal fragment (C83/αCTF) is left embedded in the membrane for further cleavage into non-amyloidogenic fragments [297, 343].


Estrogen appears to regulate Aβ levels at least in part by promoting the non-amyloidogenic cleavage of APP, precluding production of the Aβ peptide. In the human kidney 293 cell line, E2 has been shown to reduce the level of Aβ peptide in a concentration-dependent manner [60]. Further, some preliminary results suggest that in a clinical setting, short-term E2 treatment is able to reduce plasma levels of Aβ in postmenopausal women naïve to HT [21], although the significance of plasma Aβ in terms of both AD pathogenesis and diagnostic value remains controversial.


How estrogen activates regulates APP processing is not clear, although multiple pathways have been implicated. First, most data suggest that estrogen increases the α secretase pathway of APP processing [172, 214, 366, 385]. There is evidence that estrogen can promote the α-secretase pathway via activation of extracellular-regulated kinase 1 & 2 (ERK1 & ERK2) signaling [214], a well-established estrogen signaling pathway [319, 335, 353]. The action of estrogen on the ERK components of mitogen-activated protein kinase (MAPK) signaling pathway and APP generation are rapid and may be ER independent [214]. Estrogen may also regulate APP processing through protein kinase C (PKC)-dependent pathways. PKC signaling is a strong activator of non-amyloidogenic APP processing [75, 191, 243]. Further, estrogen is a significant activator of PKC in both neuron culture [68, 69] and in brain [12, 268, 299]. Consistent with this possibility, a recent cell culture study has shown that estrogen activation of α-secretase APP processing is blocked by PKC inhibitors [385]. However, not all studies demonstrate such straightforward results. For example, an in vitro study demonstrated that E2 is capable of increasing the production of sAPPα but not reducing the release of Aβ in cortical neurons over-expressing APPswe [344].


Some evidence also suggests that estrogen may promote non-amyloidogenic APP processing by altering APP trafficking. Specifically, Greenfield et. al. reported that estrogen promoted the secretion of APP containing vesicles from the primary site of amyloidogenic APP processing, the trans golgi nucleus, thereby decreasing available APP substrate for Aβ formation [135]. In addition to promoting the non-amyloidogenic APP processing pathway, estrogen has also been implicated in the modulation of APP levels through the regulation of alternative splicing [333] and APP over-expression post-injury [307], thereby altering substrate for APP processing and subsequent Aβ production.


Importantly, subsequent studies demonstrated that estrogen also functions as a regulator of Aβ in animal models. For example, the depletion of endogenous estrogen by OVX in guinea pigs increased the levels of soluble Aβ in brain, an effect partially reversed by E2 treatment [258]. This estrogen action may be sex-dependent since androgen but not E2 treatment reduced elevated Aβ levels resulting from orchiectomy (ORX) of adult male rats [271]. A similar pattern of E2 regulation of Aβ has been observed in several transgenic mouse models of AD. That is, OVX is associated with increased Aβ and E2 treatment with reduced Aβ levels in Tg2576 [380, 390], APPswe [200], Tg2576xPS1 [367, 390] and 3xTg-AD [54, 55] mice. The mechanism by which estrogen regulates Aβ in vivo has yet to be elucidated. In the APPswetransgenic mouse model of AD, E2 treatment was associated with increased sAPPα, indicating increased α-secretase APP processing [200]. However, in wild type guinea pigs, experimental manipulation of estrogen status was not associated with corresponding changes in sAPPα levels [258]. Increased levels and activity of BACE observed in aromatase knock out mice also suggests a potential role for estrogen in the regulation of secretase expression and/or activity [380]. Further work will be needed to elucidate whether estrogen regulation of Aβ in animals involves APP processing and, if so, to define the relevant upstream signaling components (e.g., MAPK, PKC).


Curiously, estrogen levels are associated with Aβ accumulation only in some but not all transgenic mouse models of AD [121, 133, 149, 380]. In the Tg2576, PDAPP, and APPswexPS1 mouse models, OVX was not associated with increased Aβ levels and E2 treatment did not reduce Aβ levels. Discrepancies in the effects of estrogen on Aβ levels across transgenic mouse models may reflect several differences, ranging from molecular design of the transgenic lines to variability in methodological parameters such as the timing and dosing of hormone manipulations. One potentially important methodological difference across the studies is their varying techniques for Aβ quantification, each of which preferentially measures different pools of Aβ ranging from soluble monomeric Aβ to oligomeric and deposited forms. Whether estrogen differentially regulates these various Aβ pools is currently unknown. In our laboratory, we found that estrogen decreases Aβ accumulation in the 3xTg-AD mouse model as assessed by the immunoreactive load method [54, 55], which detects relatively insoluble intra- and extra-cellular Aβ. Future studies will be needed to determine exactly which Aβ pool(s) estrogen is capable of regulating and whether this contributes to observed differences in estrogen actions across models.


Discrepancies between studies on the role of estrogen as a regulator of Aβ accumulation also may indicate differences in brain levels of estrogen. Recent work suggests that OVX has limitations as a strategy to fully deplete brain estrogens. For example, in the estrogen-responsive element-luciferase mouse model, which was engineered to express the non-mammalian luciferase protein in response to classical ER activation, OVX resulted in relatively high brain estrogen activity in comparison to other body regions [66]. In evaluating the role of brain estrogens in Aβ regulation, Yue et al. [380] found that in the APP23 transgenic mouse model OVX alone was not sufficient to remove all brain estrogen and did not result in elevated Aβ levels. However, they reported elevated Aβ levels after preventing E2 formation in brain by crossing the APP23 mice with aromatase knock-out mice [380]. Thus, brain levels of sex steroid hormones, which are affected not only by gonadal hormone production but also by de novo steroid hormone synthesis in brain (i.e., neurosteroidogenesis), may be the critical factor in regulation of brain Aβ accumulation. This hypothesized importance of brain hormone levels is supported by the finding that brain levels of E2 are lower in female AD patients in comparison to age-matched control cases [380], a finding we have recently replicated [284]. In addition, it has recently been shown that long-term OVX in female mice significantly lowers E2 levels in the hippocampus while increasing serum levels of Aβ [106]. Thus, estrogen regulation of Aβ may depend primarily upon brain estrogen levels, which may be depend on both ovarian and brain steroid production.


In addition to modulating the production of Aβ, estrogen may also promote Aβ clearance. One mechanism of Aβ clearance is through microglial degradation [282]. Estrogen has been shown to promote microglial phagocytosis [48, 267] and E2 treatment increases microglial internalization of Aβ in microglia of both murine [144] and human origin [202]. Estrogen treatment has also been found to reduce Aβ accumulation in rats following intracerebroventricular Aβ injection [144]. Correspondingly, increased Aβ burden and impaired microglial Aβ clearance has been reported in an estrogen deficient transgenic mouse model of AD [380]. Estrogen has also been implicated in the regulation of levels of two major Aβ degrading enzymes, insulin degrading enzyme and neprilysin. Both of these enzymes are significant regulators of Aβ levels and their regulation and activities are implicated in AD pathogenesis [331, 337]. A few recent studies suggest that estrogen depletion by OVX can decrease neprilysin activity in female rat brain, an effect reversed by E2replacement [164]. Thus, estrogens may influence Aβ clearance through regulation of neprilysin, although this pathway may involve an androgen responsive element on the neprilysin gene [365]. Estrogen pathways of Aβ-lowering are illustrated in Figure 1. Given the significance of Aβ accumulation to AD pathogenesis, future studies must clearly define the role of estrogen in regulating both Aβ production and clearance pathways and how they are affected by differences in E2 brain levels.


5. Aging effects on estrogen responsiveness

Whether the described neuroprotective effects of estrogen prove to have therapeutic relevance to age-related neural diseases including AD will depend in part on the brain’s responsiveness to estrogen with advancing age. One of the primary criticisms of WHIMS and other clinical studies of estrogen-based HT is that the intervention may have been initiated in beyond a critical window of opportunity [119, 265, 276]. This notion refers to the possibility that the aging brain age may lose responsiveness to sex steroid hormones after an extensive period of low hormone levels, such as occurs following menopause. According to this argument, HT may exert estrogenic effects only if begun near the time of menopause. Since most participants in the WHIMS study were many years beyond menopause, the critical window hypothesis could explain in part the absence of beneficial neural actions of HT. Consistent with this position, several clinical studies have noted that HT is associated with positive neural effects in women showing menopause symptoms (e.g., flushing), suggesting retained estrogen responsiveness [152, 153, 158, 305, 306]. Although this issue remains to be fully evaluated, results from the Multi-Institutional Research on Alzheimer Genetic Epidemiology Study show that HT was associated with reduced AD risk only in postmenopausal women that initiated treatment at a relatively younger age [152].


The critical window notion of an age-related loss in brain responsiveness to estrogen is supported by studies in animal models. One experimental approach to address this issue is the “gap paradigm” in which OVX animals are treated with E2 replacement after short versus long periods of time, a design that assesses the effects of prolonged hormone deprivation on subsequent hormone exposure. In general, results from gap studies indicate diminished estrogen responses following many months of hormone absence. For example, E2and E2 plus progesteorne replacement was associated with improved spatial memory on the delayed match task when administered within 3 months post-OVX, but not when hormone treatment was delayed 10 months post-OVX [116]. Further, E2 given immediately after OVX in rats improved spatial memory performance on radial arm maze while E2 given after 5 months of OVX did not [79]. Recently, it was reported that E2 given immediately after OVX but not after a 5 months delay was able to increase hippocampal ChAT protein levels in middle-aged OVX wild type rats [36]. Although this topic requires further work to elucidate the key factors and underlying mechanism(s), the data generated thus far suggest that in laboratory animals, the brain can show reduced hormone responsiveness after an extended period of hormone depletion.


Another approach to evaluate the role of aging in neural estrogen responsiveness has been to simply compare the effects of estrogen in young adult versus aging female rodents. This type of study is critical in determining the efficacy of female sex steroid hormones before and after the onset of reproductive senescence, the result of normal reproductive aging in female rodents (reviewed in [59]). While reproductive aging in rodents does not mimic menopause, rodents do experience a pattern of estrus cycle irregularity followed by reproductive senescence that is similar in some respects with the perimenopause period in women [96]. Age-related reproductive changes in female rodents typically become apparent between 9–11 months of age, depending upon species and strain. For example, the age-related estrus cycle changes have been well characterized in the C57Bl6 mouse strain [96]. Like women, these female mice demonstrate a large range of variability in cycle irregularity and hormone levels during middle and old age. These mice experience cycle cessation between 11–16 mo of age during which a substantial proportion enter a period of persistent vaginal cornification lasting 2–4 months. After this variable period, all mice enter an irreversible final stage of permanent diestrus characterized by low E2 and progesterone levels and elevated luteinizing hormone levels [115], ovarian follicle depletion, and loss of reproductive capability [127].


In estrogen neuroprotection studies, middle-aged female rodents undergoing reproductive senescence show diminished effects in some studies but retained protection in others. First, several reports suggest that the neural effects of OVX and E2 treatment are diminished in aging female rodents. Studies by Sohrabji and colleagues suggest that reproductive senescence in middle-aged female rats reduces protective estrogen actions. For example, in assessing estrogen regulation of neurotrophin expression, they found that E2treatment in OVX young adult female rats (age 3 mo) increased levels of BDNF, trkA, and trkB in olfactory bulb and diagonal band of Broca, whereas E2 treatment of middle-aged (17 months) OVX reproductively senescent, female rats showed either no increase or decreased expression of neurotrophins [176]. Similar studies by this research group found that, in comparison to young adult female rats, reproductively senescent female rats show several alterations in estrogen-mediated effects including cytokine and growth factor responses following injury [177, 238] and blood-brain-barrier permeability [20]. In another paradigm, Finch and colleagues reported differences between young adult (3 mo) and middle-aged (18 mo) female rats on estrogen regulation of compensatory neuronal sprouting following entorhinal cortex lesion. In comparison to young rats, older rats no longer exhibited an OVX-induced decrease in sprouting and showed differences in regulation of GFAP mRNA [321]. However, some neural actions of estrogen appear to remain relatively robust during aging. For example, E2 treatment in middle-aged OVX rats enhanced performance on hippocampal-dependent spatial memory tasks [79] as well as altered the hippocampal expression of several genes [2].


In neural injury paradigms, there is evidence of both retained and altered estrogen neuroprotection. In reproductively senescent female rats, both E2 and progesterone reduced infarct size in a model of stroke injury [6]. Wise and colleagues reported similar neuroprotective effects of E2 treatment in OVX young (3–4 mo) and middle-aged (9–12 mo) rats in the middle cerebral artery occlusion (MCAO) model of stroke [87, 360]. In our laboratory, we have found that reproductively senescent female rats show evidence of altered but not completely diminished estrogen neuroprotection. In young adult (3 mo) female rats, an excitotoxic lesion to hippocampus induced by systemic application of the glutamate agonist kainate is significantly worsened in animals OVX for 2 weeks prior to the lesion. In a similar paradigm, we find that E2 treatment, initiated at the time of OVX, significantly increases the number of viable hippocampal neurons following kainate lesion [287]. However, in reproductively senescent (14 mo) rats, OVX was not associated with further cell loss, although E2 replacement still resulted in a significant increase in neuron survival [56]. In a model of spinal cord injury, the effects of E2 treatment in young (2 mo) versus middle-aged (12 mo) sham-OVX and OVX female rats were investigated. Treatment with E2 protected against several indices of injury in both young and aging OVX rats, however in ovary-intact rats E2 neuroprotection was lost in middle-aged rats [61].


Interestingly, emerging data suggest that patterns of reproductive aging in female rats may contribute to altered estrogen responsiveness with age. A recent study compared estrogen protection using the MCAO stroke model in middle-aged female rats stratified by their stage of reproductive aging: reproductively senescent rats that had entered a persistent acyclic state, and rats with normal but lengthened cycles. In comparison to the middle-aged cycling rats, the reproductively senescent rats exhibited larger lesions and no longer showed reduced lesion size following E2 treatment [298]. Our recent data also suggest that responsiveness to estrogen protection vary according to status of reproductive aging. In the kainate lesion model, we observed that E2 was most protective in rats showing an acyclic state of persistent vaginal cornification compared to rats with present, albeit irregular cycles [56]. Thus, although additional studies are necessary to define the relationships, it is reasonable to hypothesize that patterns of reproductive aging affect estrogen neuroprotection. Such findings are consistent with the “critical window” hypothesis of HT and have important implications for the future of clinical use of estrogen-based therapies.

Why the brain shows is less responsive to estrogen with age is unclear, but age-related decreases in ER expression as well as E2 binding to ERs in aged rat brain have been reported [59, 289, 361]. Estrogen actions also show age-related changes in other estrogen responsive tissues, including uterus, bone and heart. For example, the uterus becomes less responsive to estrogen with increasing age, showing smaller OVX-induced decreases in uterine weight [368] and uterotrophic effects of estrogen only when treated soon after OVX [79]. Further, while most studies demonstrate that the trophic effects of E2 on bone extend through middle-age, some studies suggest that as aging progresses, this effect is altered from an ERα/β-mediated effect to only an ERα mediated one [169]. Interestingly, increased ERα predominance has also been suggested to underlie age changes in neural estrogen responsiveness [19].


Taken together, available experimental research suggests that although estrogen neuroprotection is often observed in aging rats, it can also be significantly diminished. Still uncertain is how the many AD-related facets of estrogen neuroprotection may be impacted by aging. Additional studies are necessary to determine the extent to which observed age-related changes in estrogen responsiveness may be delayed, prevented and or reversed. This phenomenon has important implications for the future of HT in postmenopausal women. Ongoing clinical and animal studies promise to shed new insight into this issue in the next few years.


6. Progesterone interactions with estrogen in regulation of Alzheimer’s disease

Estrogen actions must also be considered with respect to the second major class of ovarian hormones, progestogens. Progesterone has long been recognized as a regulator of estrogen, particularly in the female reproductive system (reviewed in [131]), where it often antagonizes estrogen action. Clinically, progestogens are typically a key element of HT that are thought to minimize deleterious effects of estrogen. Perhaps most importantly, in experimental paradigms progesterone can inhibit human endometrial cancer cell growth [78, 80] and in clinical studies progestogens are associated with reduced risk of endometrial cancer [130, 257].


In clinical studies of AD and dementia, most results indicate similar outcomes with both estrogens alone (ie., CEE) and estrogens in combination with progestogens (ie., CEE+MPA). In the WHIMS trial, a comparison between the CEE alone and CEE+MPA arms of the study raised the question that the clinical efficacy of HT may be dependent upon the hormone constituents within. Both the CEE and CEE+MPA arms failed to demonstrate a protective effect and both actually increased the risk of dementia compared to women receiving placebo [312]. Notably, the CEE alone arm had a negative impact on global cognitive function, however this negative impact was worsened when pooled with the data from the CEE+MPA arm [92]. Interestingly, short-term HT treatment in women with existing AD was associated with some benefits on psychiatric symptoms in the CEE group but not in the CEE+MPA group [163]. These and related issues suggest that the inclusion of a progestogens may influence the effects of CEE alone on cognitive outcomes and risk of dementia is post-menopausal women [76].


Despite the common use of the progestogen MPA in current HT paradigms, researchers have only relatively recently begun to investigate the effects of progesterone on the CNS in regards to aging and neurodegenerative diseases. Although compelling experimental evidence indicates numerous protective actions of estrogen and progesterone when delivered independently [45, 295], comparatively less is known about interactions between estrogen and progesterone when they are administered together. Interestingly, accumulating observations indicate that progesterone often antagonizes rather than synergizes with estrogen-mediated neuroprotective actions. This incomplete understanding highlights the need to determine the interactive effects of these hormones in the brain.


Findings from an increasing body of research have begun to provide insight into how beneficial neural actions of estrogen are affected by interactions with progesterone. Findings from both in vitro and in vivo paradigms suggest that progesterone treatment can antagonize estrogen neuroprotection. For example, long-term progesterone treatment in aged (23–24 mo) OVX female rats blocked estrogen upregulation of brain derived neurotrophic factor, nerve growth factor, and neurotrophin 3 [34]. In a similar paradigm of hormone treatment in middle-aged OVX rats, estrogen-induced improvement in spatial memory performance was blocked by co-administration of progesterone [33]. More specific to neuron viability though are recent studies from our laboratory demonstrating that progesterone blocks estrogen neuroprotection from excitotoxic injury in female rats. In both young adult (3 mo) [287] and middle-aged (14 mo) [56] OVX rats, continuous E2treatment reduced kainate-induced neuron loss in hippocampus CA2/3 whereas continuous progesterone did not significantly affect neuron viability. Importantly, when progesterone was included with E2, estrogen neuroprotection was no longer observed [56, 287]. Similarly, Brinton and colleagues recently reported that E2 and progesterone administered independently to wild type rats enhanced several markers of brain mitochondrial function but, when replaced in combination, the hormones showed attenuated rather than enhanced responses [168].


While the mechanisms behind this antagonistic property remain unknown, our laboratory has begun to investigate the possibility that progesterone modulates estrogen function in part by regulating ER expression. We have observed in primary neuron culture that low physiological concentrations of progesterone rapidily downregulate mRNA levels of both ERα and ERβ [175]. Functionally, this progesterone-mediated decrease in ER expression was associated with inhibition of both ERE-mediated transcriptional activity and estrogen neuroprotection against Aβ [175]. Although progesterone treatment by itself did not significantly affect Aβ toxicity in this paradigm, abundant evidence demonstrates that progesterone can activate several neuroprotective cell signaling pathways, including Akt [318] and ERK [233, 234] signaling and upregulation of the anti-apoptotic protein Bcl-2 [234]. Further, as reviewed in an accompanying article, progesterone can significantly protect neurons against numerous insults [45, 295]. Thus, independently estrogen and progesterone can exert protective actions, but in combination they can inhibit each other and thus fail to protect. Perhaps not unexpectedly, the interactive neuroprotective effects of the two female sex steroid hormones are not quite so straightforward. Besides antagonistic effects, progesterone can also synergistically interact with estrogen to promote beneficial neural effects including increased spine density [101, 313]. In OVX female rats, acute progesterone treatment (2–6 hours) was observed to augment the estrogen-induced increase in spine density, whereas prolonged progesterone treatment (18 hours) blocked the estrogen effect [362]. Thus, a key factor in understanding estrogen neuroprotection and perhaps its relevance to HT in postmenopausal women is elucidation of the interactions between estrogen and progesterone.


A similar progesterone regulatory relationship also appears to affect estrogen protection from indices of AD-like neuropathology. In recent studies, our laboratory has begun to investigate how progesterone interacts with estrogen in regulating Aβ accumulation in the 3xTg-AD transgenic mouse model of AD. As discussed above, we found that 3 months following OVX of young adult female 3xTg-AD mice, there were robust increases in Aβ accumulation and tau phosphorylation and impaired performance in spontaneous alternation behavior in comparison to sham OVX 3xTg-AD mice [55]. Continuous E2 treatment during the 3 month OVX period largely prevented the worsening of AD-like neuropathology and behavioral performance, however continuous progesterone by itself did not affect OVX-induced changes in either Aβ accumulation or behavior. Consistent with an antagonistic role of progesterone, we observed that co-treatment with progesterone antagonized the beneficial effect of estrogen in lowering Aβ accumulation [55]. However, estrogen and progesterone co-treatment did show a beneficial effect in reducing levels of tau hyperphosphorylation, suggesting positive estrogen-progesterone interactions. Consistent with a beneficial role of progesterone, Frye and colleagues report that long-term progesterone treatment following OVX in another transgenic mouse model of AD was associated with some cognitive benefits [105]. Taken together, these studies provide the first information to date on the interactive effects of estrogen and progesterone on AD-like neuropathology and demonstrated the potential for both positive and negative outcomes in terms of protection from AD-related neuropathology.

Despite evidence of antagonistic neural effects of progesterone on estrogen neuroprotective actions, progestogens are still deemed a necessary component of HT in women with a uterus. Therefore, HT may need to be optimized to maximize the benefits and minimize the unwanted consequences associated with estrogen-progestogen interactions. One possible strategy is the use of cyclic hormone delivery rather than the continuous, combined treatment that is currently common to HT. Initial clinical evaluation of cyclic progestogen exposure has been completed and more is underway. Several clinical HT trials have incorporated a comparison between continuous versus cyclic progestogen in postmenopausal women on osteoporosis and cognitive function. For example, two completed studies have both demonstrated that long term HT with a cyclic progestogen dose was able to increase bone mineral density in post-menopausal women [1, 57]. Similarly, a continuous estrogen plus cyclic progestogen paradigm is currently employed in the KEEPS (Kronos Early Estrogen Prevention Study) Cognitive and Affective Study, a randomized, placebo-controlled, double-blind study investigating the effects of HT in postmenopausal women who are within 36 months of their final menstrual period [357]. This and similar new trials promise to provide important insight on the efficacy of cyclic hormone delivery and the hypothesized importance of a critical window of hormone intervention.


Experimental studies in animal models lend support for the use of cyclic rather than continuous progesterone to optimize estrogen neuroprotection. For example, Gibbs and colleagues have demonstrated that short-term treatment with E2 and progesterone can improve cholinergic function [117], with maximal benefit resulting from cyclic administration of estradiol and progesterone and the least benefit from continuous combined treatment [116]. In our laboratory, we have begun investigating the potential utility of cyclic progesterone against AD neuropathology by comparing cyclic versus continuous progesterone in the presence and absence of continuous E2 in OVX 3xTg-AD mice. Our results show that whereas continuous progesterone largely inhibits estrogen protection from AD-related neuropathology, cyclic progesterone appears to significantly increase estrogen protection against Aβ accumulation, tau phosphorylation, and working memory deficits (unpublished observations). These exciting new findings support the hypothesis that estrogen-progesterone interactions can yield additive neuroprotection and that cyclic hormone delivery may be a critical parameter.


7. Age-related androgen depletion and Alzheimer’s disease

In parallel to the relationships between age-related estrogen loss in women and increased AD risk, testosterone is depleted as a normal consequence of aging in men and is linked with elevated risk of AD. As discussed above, a significant biological event in women that contributes to the role of aging in AD is menopause and the resultant loss of the sex steroid hormones estrogen and progesterone. Although men do not experience menopause per se (i.e., a cessation of reproductive ability, nearly complete loss of sex steroid hormones), men do experience a somewhat similar process termed androgen deficiency in aging males (ADAM).ADAM refers to normal, age-related depletion of testosterone and the corresponding constellation of symptoms that reflect dysfunction and vulnerability to disease in androgen-responsive tissues including brain [25, 51, 97, 126, 180, 184, 218, 225, 302]. The decline in testosterone levels begins in the 3rd decade and continues at an annual rate of 0.2 – 1% for total testosterone and 2 – 3% for bioavailable testosterone [95, 132, 228]. Unlike menopause, aging men do not experience comparable levels of andropause. That is, although all men exhibit significant age-related testosterone loss typically beginning in the third decade of life, men vary in the extent of testosterone loss and the corresponding severity of clinical manifestations [226, 326]. It is estimated that 30% – 70% of men aged 70 and older are hypogonadal, resulting in at least 5 million aging men in the U.S. suffering the consequences of andropause and only a small minority of those receive hormone treatment [142, 227]. ADAM is associated with increased risk of sarcopenia, osteoporosis, falls, frailty, and all cause mortality.

The brain is a highly androgen responsive tissue where androgens induce several beneficial actions. For example, androgens have been shown to improve mood and promote select aspects of cognition, including spatial abilities [128, 174] and verbal fluency [5]. For example, men with a higher free testosterone index have been found to perform better on visual and verbal memory and exhibited better long-term memory [24], while those with a low free testosterone index can show decreased visual memory, visuomotor scanning, verbal memory, and visuospatial processing [220]. ADAM has been associated with impaired cognitive performance in some but not all studies [143, 220].


One recently established consequence of ADAM is an increased risk for the development of AD. Several [160–162, 274, 352] but not all [255] studies have identified a relationship between low circulating levels of testosterone and a clinical diagnosis of AD. In these studies, the relationship between testosterone and AD appears to be strongest when circulating levels of free rather than total testosterone are examined, and when mean ages are under 80 years [160–162, 249]. The relationship between testosterone and AD may be influenced by the presence of at least one apolipoprotein ε4 allele, a genetic risk factor for AD [322]. Specifically, men with at least one ε4 allele had lower levels of testosterone than men without an ε4 allele [161]. Animal studies support a link between apolipoprotein E, testosterone, and AD [269, 270].


Although the majority of studies have identified a relationship between low testosterone and increased AD risk in men, most were unable to determine whether low testosterone contributes to the disease process or is merely a result of it. However, two complementary studies suggest low testosterone occurs prior to or in the early stages of AD pathogenesis, and thus likely acts a risk factor. The first study compared clinical diagnosis of dementia with blood levels of testosterone in the prospective Baltimore Longitudinal Study on Aging [221]. Male subjects were followed for 4 to 37 years with a mean of 19 years per subject and were diagnosed as clinically normal at the time of their first testosterone measurement. Subjects that eventually received a clinical diagnosis of AD showed lower circulating levels of free testosterone. Interestingly, in those men with AD, testosterone levels were reduced at checkups 5 to 10 years prior to diagnosis [221], suggesting androgen loss occurred well before clinical manifestations of the disease. Consistent with this study are findings from a study by our laboratory in which we linked low brain levels of testosterone with increased risk of AD in men [285]. First, using human postmortem brain tissue from neuropathologically normal men, we found that levels of testosterone but not E2 show a significant age-related decline. When we examined changes in brain levels of hormones across cases stratified by neuropathological status, we found significantly decreased brain levels of testosterone in AD cases as compared with neuropathologically normal cases even after controlling for the age-related hormone loss [285]. We also measured brain levels of androgens in cases with mild neuropathological changes, consistent with the earliest stages of AD. These cases also exhibited low testosterone levels [285], again indicating that testosterone loss occurs prior to robust pathology and thus may contribute to the development of AD. Taken together, these results suggest age-related testosterone depletion in men is a risk factor for AD.


Unlike the numerous clinical studies that have evaluated the efficacy of HT use in treating and or preventing AD in postmenopausal women, comparatively few studies have examined testosterone therapy in men for protective roles against age-related cognitive decline and development of dementia. Androgen therapies have been approved and used for the treatment of hypogonadism in men and are typically associated with reduced fat mass and improved muscle mass and bone density as well as increased mood, libido, and overall quality of life [31, 327, 332]. However, clinical evaluation of beneficial cognitive effects of testosterone therapy have been mixed. For example, a study of hypogonadal and eugonadal men found that testosterone increased verbal fluency [5]. Others report improvements in spatial cognition and working memory following testosterone treatment [63, 173, 174]. In contrast, some studies did not find significant changes in cognition following testosterone therapy [32, 91, 143, 211, 339].


A few studies have evaluated the effects of testosterone therapy in subjects with AD. In a small clinical study of men recently diagnosed with AD, testosterone treatment for up to one year contributed to improvement in both overall cognitive ability and visual spatial skills [328]. In contrast, other studies have not reported significant benefits of testosterone therapy in men with mild cognitive impairment and AD [64, 206]. As with the observed inconsistencies in the literature of HT use in women, there are likely several factors that contribute to the observed differences between testosterone studies, including cognitive domains, treatment type and duration, and the age and other characteristics of the subjects.


8. Testosterone neuroprotection and Alzheimer’s disease insults

One beneficial action of androgens that is hypothesized to contribute to a role in reducing risk of AD is neuroprotection. Androgens are established promoters of neuron viability during neural development as well as in adult brain following mechanical injury and disease-related toxicity. One target of androgen neuroprotection is motorneurons following axotomy [178]. In this paradigm, testosterone treatment accelerates the rate of nerve regeneration and attenuates neuron loss [165, 178, 179, 193, 194, 196, 330, 378, 379]. Similarly, following facial nerve crush in male hamsters, testosterone increased the rate of axonal growth and functional recovery [193, 195]. These effects are true not only for testosterone but also its potent androgen metabolite dihydrotestosterone (DHT) [377–379]. In addition, the anti-androgen flutamide was able to block testosterone’s neuroprotective effects on motor neurons [196], corroborating the role of androgen versus estrogen pathways. The mechanism behind this neuroprotective action appears to be through androgen regulation of trophic factors [377, 379]. Recently, studies have found that in addition to long-term treatment, short-term testosterone, DHT, and estrogen treatments are protective and suggest a more direct mechanism of hormone action [165].


In addition to neuroprotective effects on motor neurons, androgens have also been found to promote neuron survival in brain regions vulnerable to neurodegenerative diseases such as Alzheimer’s disease. These areas include the hippocampus and cortical regions, which are both affected in AD and rich in androgen receptors [314]. In a study by Garcia Segura and colleagues, acute testosterone treatment attenuated neuron loss in the hilus of the dentate gyrus following excitotoxic lesion in ORX male mice [17]. Interestingly, acute treatment of E2 was also protective while DHT treatment did not protect against neuron loss in androgen depleted mice. Furthermore, the protective effect of testosterone was blocked by an aromatase inhibitor, suggesting that in this model of acute hormone treatment, estrogen is responsible for testosterone neuroprotection [17]. A study from our laboratory investigated the effect of long-term hormone replacement on neuronal death induced by excitotoxic lesion. In this study, depletion of endogenous androgens as a consequence of ORX resulted in increased hippocampal neuron loss following kainate lesion in comparison with sham ORX rats. However, in ORX rats treated for two weeks with DHT, this increase in cell loss was blocked [272]. In contrast to the results of Garcia Segura and colleagues, we found that rats treated with E2 did not exhibit decreases cell loss following kainate lesion. This suggests that androgen regulation of neuroprotection in this paradigm is a result of androgen rather than estrogen pathways. Behavioral responses to seizures were measured and no differences were observed between groups suggesting that the observed androgen neuroprotection was not a result of decreased seizure severity [272]. Although we did not observe an androgen-mediated effect on seizure severity, other studies have found a significant effect of androgens on seizure severity. In studies by Frye and colleagues, testosterone decreased neuron loss by inhibiting seizure activity [103, 104]. Subsequent studies found that protection was due to DHT metabolism to 5α-androstane-3α, 17α-diol, an androgen that acts on GABAA receptors. GABA activation redcues excitatory signaling, which in turn attenuates seizure activity thereby minimizing lesion severity [88, 89, 277].


While in vivo models provide valuable insight into androgen neuroprotection, neuron culture models of toxicity have proven valuable in defining the underlying molecular mechanisms. Cell culture models of neural injury have demonstrated testosterone protection against serum deprivation [46, 139], Aβ toxicity [231, 260, 386], and oxidative damage [3]. Testosterone neuroprotection against serum deprivation-induced apoptosis requires activation of an androgen receptor (AR) dependent mechanism [139]. Specifically, the anti-androgen flutamide attenuated protection while an aromatase inhibitor had no effect on neuron viability [139]. Consistent with this androgen-mediated mechanism of androgen neuroprotection is an early study from our laboratory, which found that testosterone neuroprotection against toxicity induced by extracellular Aβ results from DHT not E2 [260]. DHT treatment in this paradigm was equally as protective as testosterone, but use of an anti-estrogen droloxifene failed to block protection, suggesting androgen pathways are responsible for neuroprotection [260]. Further, we observe androgen neuroprotection in PC12 cells transfected with AR but not in either untransfected PC12 cells or those transfected with empty vector [231].


There are several potential mediators of androgen neuroprotection downstream of AR. Some evidence suggests androgen neuroprotection may be mediated through attenuation of oxidative stress [3]. Another potential mechanism involves a classic genomic mechanism, increased expression heat shock proteins. Androgen attenuation of Aβ induced toxicity was associated with elevated levels of heat shock protein 70, which is known to participate in protective responses against cellular stress and neurodegeneration [386]. Recent work from our lab identified a non-genomic, AR-dependent mechanism involving activation of a mitogen-activated protein kinase (MAPK) / extracellular signal regulated kinase (ERK) pathway [231]. We observed that physiological levels of testosterone and DHT rapidly and transiently activated MAPK/ERK signaling in cultured hippocampal neurons. Downstream of ERK, we found that androgens activated p90kDa ribosomal S6 kinase (Rsk), which in turn phosphorylates the pro-apoptotic protein Bad. Phosphorylation of Bad results in its inactivation of Bad, thereby tilting the balance of apoptosis towards increased cell viability. Pharmacological inhibition at any step of this pathway prevents both phosphorylation of Bad and androgen neuroprotection [231]. Confirming the non-genomic nature of this pathway, we found that in this neuron culture paradigm the classic anti-androgens flutamide and cyproterone acetate mimicked neuroprotection against apoptotic insults afforded by testosterone and DHT even though they also blocked classic genomic actions of androgens [232]. Further, protective actions of flutamide and cyproterone acetate were only observed in AR-containing cell lines. Whether activation of protective androgen pathways requires membrane AR, intracellular AR, or both is unclear. We observed that testosterone conjugated to albumin – which is designed to prevent cell entry and thus permit only activation of membrane receptors – was ineffective in activating protective MAPK/ERK androgen signaling [231]. Similarly, Gatson and colleagues reported that while DHT phosphorylates both ERK and Akt, albumin-conjugated DHT resulted in dose-dependent suppression of ERK signaling in glioma cells expressing AR [113]. Pathways of androgen neuroprotection are summarized in Figure 2.


Figure 2

Androgens activate neuroprotective pathways that may attenuate Alzheimer’s disease. First, testosterone (T) is aromatized in brain to 17β-estradiol (E2), which activates estrogen-mediated neuroprotective pathways (summarized in Figure ...

In contrast to these observations of androgen neuroprotection, there appear to be circumstances in which androgens fail to protect against neural injury and can even exacerbate insults. For example, Dluzen and colleagues found that E2 but not testosterone reduces methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) nigrostriatal dopaminergic toxicity [83]. Further, testosterone has been found to exacerbate neuron loss following middle cerebral artery occlusion (MCAO) [371, 372]. In this model removal of testosterone 6 hours prior to MCAO reduces lesion volume by as much as half [371, 372]. It is not clear whether the absence of androgen neuroprotection in these paradigms reflects roles of brain region, insult, and or other factors. One important issue in at least some studies may be androgen concentration. That is, at physiological low nanomolar levels testosterone can protect against excitoxicity in cultured neurons, but worsens cell death when present at supraphysiological micromolar levels [244]. Similarly, micromolar but not nanomolar concentrations of testosterone were associated with increased calcium signaling and neuronal apoptosis [93, 114]. Thus, although there is ample evidence of androgen neuroprotection against AD-related insults, androgen effects on neuron viability may depend on a number of factors, including brain region, type of insult, and hormone concentration.


In addition to direct neuroprotection, androgens also protect against another form of neuropathology directly relevant to AD, hyperphosphorylation of tau. Abnormal, excessive phosphorylation of the cytoskeletal protein tau in the form of neuropil threads and neurofibrillary tangles is a defining neuropathological characteristic of AD and several other neurodegenerative disorders [167, 355, 358]. Although relatively little research has examined the relationship between androgens and tau hyperphosphorylation, available evidence does indicate that androgens are protective against this pathology. Papasozomenos and colleagues have examined the effects of testosterone and E2 on tau hyperphosphorylation using a model of heat shock-induced phosphorylation. In this paradigm, testosterone but not E2 prevents tau hyperphosphorylation [250–252]. In addition to affecting phosphorylation of tau, recent evidence suggests that androgens also regulate tau cleavage [253]. Specifically, testosterone prevented calpain-mediated tau cleavage preventing the generation of the 17-kDa tau fragment [253].


9. Testosterone regulation of β-amyloid accumulation

In addition to classic neuroprotective actions, androgens may also protect the brain from AD by regulating accumulation of Aβ. Initial work suggesting a relationship between androgens and Aβ came from a small study evaluating men treated with anti-androgen therapies for prostate cancer. Gandy and colleagues found that within several weeks following initiation of anti-androgen therapy (consisting of leuprolide and flutamide), circulating levels of testosterone and E2 were largely depleted whereas plasma levels of Aβ were significantly elevated [107].Martins and colleagues similarly reported an association between androgen depletion and elevated Aβ levels in males receiving anti-androgen therapy for the treatment of prostate cancer [7] and in older males suffering from memory loss or dementia [118]. To what extent these observations reflect effects of androgen versus estrogen pathways or perhaps reflect associated changes in gonadotropins remains incompletely resolved [14, 286].


Consistent with the observations in aging men, experimental work in rodents also indicates that androgens function as endogenous negative regulators of Aβ. Early studies in our laboratory demonstrated that androgens but not estrogens reduce brain levels of soluble Aβ in male rats [271]. Specifically, we found that ORX-induced depletion of endogenous androgens was associated with a modest but significant increase in soluble Aβ from hemi-brain lysates.Further, DHT treatment in ORX rats resulted in a significant reduction in brain Aβ levels, although E2 treatment had no effect [271]. We observed similar findings on the relationship between androgen levels and Aβ accumulation in the 3xTg-AD mouse model of AD. Male 3xTg-AD mice were depleted of androgens by ORX at age 3 mo, a time prior to the development of significant AD-like pathology [240]. Mice were exposed immediately to either DHT or vehicle, treatments that were maintained continuously for the next 4 months. After the treatment period, we observed significant intracellular accumulation of presumably insoluble Aβ in gonadally intact, sham ORX 3xTg-AD mice (age 7 mo) in the CA1 hippocampus, subiculum, and amygdala [283]. In comparison to the sham ORX animals, the ORX mice showed significantly higher levels of Aβ in all three brain regions as well as significantly poorer behavioral performance on a spontaneous alternation task [283]. The elevated pathology and exacerbated behavioral impairments in the ORX group were blocked in ORX mice treated with DHT, suggesting a preventive effect of androgens in regulation of AD-related pathology. Confirming our findings in animal models, a recent study reported that levels of Aβ in plasma and cerebrospinal fluid were significantly elevated following ORX in guinea pigs, an effect prevented by testosterone replacement 1 week following ORX [346].


There are likely several mechanisms that contribute to androgen regulation of Aβ (Figure 2). One obvious possibility is that aromatase-mediated conversion of testosterone to E2 in brain allows activation of the several estrogen pathways of Aβ regulation discussed above. In fact, cell culture studies are consistent with indirect testosterone activation of estrogen-mediated regulation of APP processing. The first study to examine testosterone regulation of APP and Aβ in culture found that prolonged testosterone treatment was associated with increased α-secretase cleavage of APP and reduced Aβ [129]. Unclear was whether these findings involved direct androgen pathways, indirect activation of estrogen pathways, or both. In a subsequent culture study, testosterone was also observed to promote proteolysis of APP by α-secretase, however this effect was blocked in the presence of aromatase inhibitors suggesting estrogen dependence [122]. In our study of male rats in which androgens were found to regulate brain levels of soluble Aβ, we did not observe detectable differences in either full-length APP or APPα across androgen treatment groups [271], perhaps indicating that alterations in APP processing are not the only mechanism by which androgens affect Aβ levels.


In addition to regulating Aβ generation via effects on APP proteolysis, androgens may also decrease Aβ levels by promoting endogenous clearance pathways. Recent evidence from our group demonstrates that androgens reduce Aβ levels as a consequence of regulating expression of the Aβ-catabolizing enzyme neprilysin [376], a critical enzyme in homeostasis of brain Aβ [171]. In neural cell cultures, we found that androgens robustly increase protein levels of neprilysin. A classic, AR-dependent genomic mechanism is implicated since (i) the neprilysin gene contains androgen response elements [304], (ii) neprilysin regulation was blocked by the anti-androgens flutamide and cyproterone acetate, and (iii) androgens only increased neprilysin in AR-containing cells [376]. Overexpression of human APP in cultured cells resulted in elevated levels of soluble Aβ that were reduced by testosterone and DHT. Further, we found that pharmacological inhibition of either AR or neprilysin blocked the ability of androgens to reduce Aβ levels, suggesting that the Aβ-lowering actions of androgens is mediated by AR-dependent regulation of neprilysin expression [376]. Importantly, these observations were replicated in male rats. We found that ORX-induced androgen depletion resulted in decreased levels of NEP and elevated Aβ, and DHT replacement restored NEP and Aβ to levels observed in sham GDX animals [376].


10. Emerging strategies of hormone-related therapies in Alzheimer’s disease

As reviewed in this article, there are numerous neuroprotective actions of estrogens and androgens that have direct relevance to AD pathogenesis and compelling potential to prevent and possibly treat the disease. However, the promise of estrogen-based and androgen based HTs in reducing AD risk have yet to be realized. As findings and directions from clinical and basic science research become increasingly integrated, it is anticipated that critical parameters affecting HT efficacy will be optimized, including age of HT initiation as well as the formulation, regimen, and delivery of HT. As ongoing research continues to address these crucial and immediate concerns, an emerging area of investigation is the development of natural and synthetic hormone mimetics that will preferentially activate estrogen and androgen neuroprotective mechanisms while minimizing deleterious consequences in other tissues.


Estrogen compounds that show tissue-selective agonist actions are termed selective estrogen receptor modulators (SERMs). Currently the most studied and clinically relevant SERMs are tamoxifen and raloxifene, synthetic compounds that exhibit tissue-dependent ER agonist and antagonist actions. As a potent antagonist of estrogen action in breast tissue, tamoxifen is best recognized as an antiestrogen used to treat breast cancer although it can exert agonist ER effects on bone and lipids [49, 300]. Interestingly, low concentrations of tamoxifen can protect cultured neurons from toxicity due to Aβ and glutamate [239], suggesting the potential for a protective role against AD. However, tamoxifen has also been observed to block E2–mediated protection in cultured neurons [58, 383]. Potential benefits of tamoxifen use in postmenopausal women for prevention or treatment of AD has not been well studied. Some studies indicate increased risk of cognitive deficits in tamoxifen users [246, 308], whereas another study suggested tamoxifen may reduce AD risk [42]. Like tamoxifen, raloxifene antagonizes estrogen actions in breast but has agonist actions on bone [49, 82]. In brain, raloxifene mimics some but not all protective estrogen actions. Raloxifene increases choline acetyltransferase in hippocampus of OVX rats [364] and in cultured neurons it increases neurite outgrowth [237] and can reduce Aβ toxicity [239]. Conversely, E2 but not raloxifene was effective in attenuating Aβ-induced inflammatory reaction in OVX rats [334]. In postmenopausal women, raloxifene use has been linked with reduced risk of cognitive impairment and development of AD [369].


Currently, effort is being focused on next generation SERMs that exhibit more robust and specific neuroprotective actions [43, 303]. For example, Brinton and colleagues recently developed a synthetic SERM with both estrogenic and antioxidant potential that protects cultured neurons from cell death [388]. Evaluation of such compounds in animal models of AD is only beginning. Towards this end, our laboratory has begun to investigate two particular SERMs, propylpyrazole triol (PPT) and diarylpropionitrile (DPN), which show relative specificity for ER? and ERβ, respectively. Although we observe that both compounds protect cultured neurons from Aβ toxicity [70], PPT but not DPN treatments effectively mimicked E2 in reducing Aβ accumulation and improving behavioral deficits in OVX 3xTg-AD mice [54].


In parallel to the ongoing research on SERMs, there has been a recent growth in the interest and development of tissue-specific SARMs [109, 148, 187]. Given the prevalence of ADAM and its widespread deleterious consequences, androgen based therapy is of considerable interest. However, prostate cancer, which is the second leading cancer among aging men in terms of both prevalence and cause of death [100], is androgen-dependent and typically treated by androgen deprivation therapy. Consequently, the use of testosterone therapy with its potential to increase risk and or progression of prostate tumorigenesis has been controversial [184] and promoted the development of SARMs that lack significant androgen action in prostate but exert agonist effects in select androgen-responsive tissues of interest, including brain, muscle, and bone.


Several strategies of SARM design are currently been pursued [47, 109, 359]. One strategy is to develop novel steroidal compounds that are not substrates for the enzyme 5α-reductase or yield reduced metabolites with minimal androgencity. Testosterone is converted to DHT by the actions of 5α-reductase, an enzyme localized in specific target tissues such as prostate. Prostate growth depends largely on the actions of DHT rather than T because DHT exhibits ∼10-fold greater net potency, which reflects both a higher binding affinity for AR and a slower dissociation rate from AR [359]. SARMs that are not 5α-reductase substrates or form DHT-like derivatives with weak androgenic activity have low androgen action in prostate [109, 189, 201, 212, 245, 301, 342]. A promising SARM in this category is 7α-methyl-19-nortestosterone, commonly called MENT [224, 301, 342], which was developed by the Population Council and is currently in clinical trials as an androgen therapy for hypogonadal men [11, 345]. MENT shows low androgen activity in prostate but is more potent than T in other peripheral androgen-responsive tissues including bone [77, 301, 342] and muscle [77, 301, 342]. The effects of MENT on neural function are virtually unknown, but it has been shown to mimic the ability of testosterone to induce sexual behavior in castrated rats [224] and is a robust regulator of the hypothalamic-pituitary-gonadal axis, suggesting neural efficacy.


Another SARM design strategy is the development of non-steroidal synthetic AR ligands.One such SARM is flutamide, which mimics the abilities of T and DHT to increase hippocampal spine density in both male and female rats [208, 209]. In cultured neurons, we have found that flutamide antagonizes classic genomic actions of T and DHT but mimics rather than blocks the nongenomic neuroprotective actions of these androgens [230]. Of particular interest are novel compounds that bind AR but have altered interactions with AR binding pocket side chains that underlie tissue specificity. Recent research has determined that although synthetic SARMs must closely mimic the rigid backbone core structure of the AR ligand binding domain, there can be extensive variation in how they interact with amino side chains in the binding pocket [359]. Thus, SARMs are predicted to exert tissue-specific effects dependent in part upon how they interact with amino acid sidechains in the AR binding pocket [37, 47, 201]. Recent research indicates success in developing SARMs with desired tissue specificity [245]. As with SERMs, the development of brain-specific SARMs that exert neuroprotective actions associated with reduction of AD pathogenesis is a key topic of ongoing research.


Acknowledgment


This work was supported by NIH grants AG026572 and AG23739. JCC was supported by NIH grant AG032233.


Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.


References

1. Effects of estrogen or estrogen/progestin regimens on heart disease risk factors in postmenopausal women. The Postmenopausal Estrogen/Progestin Interventions (PEPI) Trial. The Writing Group for the PEPI Trial. Jama. 1995;273:199–208.

2. Aenlle KK, Kumar A, Cui L, Jackson TC, Foster TC. Estrogen effects on cognition and hippocampal transcription in middle-aged mice. Neurobiol Aging. 2007

3. Ahlbom E, Prins GS, Ceccatelli S. Testosterone protects cerebellar granule cells from oxidative stress-induced cell death through a receptor mediated mechanism. Brain Research. 2001;892:255–262.

4. Akiyama H, Barger S, Barnum S, Bradt B, Bauer J, Cole GM, Cooper NR, Eikelenboom P, Emmerling M, Fiebich BL. Inflammation and Alzheimer's disease. Neurobiology of Aging. 2000;21:383–421.

5. Alexander GM, Swerdloff RS, Wang C, Davidson T, McDonald V, Steiner B, Hines M. Androgen-behavior correlations in hypogonadal men and eugonadal men. II.Cognitive abilities. Horm Behav. 1998;33:85–94.

6. Alkayed NJ, Murphy SJ, Traystman RJ, Hurn PD, Miller VM. Neuroprotective effects of female gonadal steroids in reproductively senescent female rats. Stroke. 2000;31:161–168.

7. Almeida OP, Flicker L. Testosterone and dementia: too much ado about too little data. J Br Menopause Soc. 2003;9:107–110.

8. Almeida OP, Lautenschlager NT, Vasikaran S, Leedman P, Gelavis A, Flicker L. A 20-week randomized controlled trial of estradiol replacement therapy for women aged 70 years and older: effect on mood, cognition and quality of life. Neurobiol Aging. 2006;27:141–149.

9. Alvarez-De-La-Rosa M, Silva I, Nilsen J, Perez MM, Garcia-Segura LM, Avila J, Naftolin F. Estradiol prevents neural Tau hyperphosphorylation characteristic of Alzheimer's disease. Annals of the New York Academy of Sciences. 2005;1052:210–224.

10. Andersen K, Launer LJ, Dewey ME, Letenneur L, Ott A, Copeland JRM, Dartigues J-F, Kragh-Sorensen P, Baldereschi M, Brayne C, Lobo A, Martinez-Lage JM, Stijnen T, Hofman A. Gender differences in the incidence of AD and vascular dementia:The EURODEM Studies. Neurology. 1999;53:1992–1997.

11. Anderson RA, Martin CW, Kung AW, Everington D, Pun TC, Tan KC, Bancroft J, Sundaram K, Moo-Young AJ, Baird DT. 7α-methyl-19-nortestosterone maintains sexual behavior and mood in hypogonadal men. J Clin Endocrinol Metab. 1999;84:3556–3562.

12. Ansonoff MA, Etgen AM. Estradiol elevates protein kinase C catalytic activity in the preoptic area of female rats. Endocrinology. 1998;139:3050–3056. [PubMed]

13. Antonsson B, Martinou JC. The Bcl-2 protein family. Exp Cell Res. 2000;256:50–57.

14. Atwood CS, Meethal SV, Liu T, Wilson AC, Gallego M, Smith MA, Bowen RL. Dysregulation of the Hypothalamic-Pituitary-Gonadal Axis with Menopause and Andropause Promotes Neurodegenerative Senescence. Journal of Neuropathology and Experimental Neurology. 2005;64:93–103.

15. Azcoitia I, Sierra A, Garcia_Segura LM. Neuroprotective effects of estradiol in the adult rat hippocampus: interaction with insulin-like growth factor-I signalling. Journal of Neuroscience Research. 1999;58:815–822.

16. Azcoitia I, Sierra A, Garcia-Segura LM. Estradiol prevents kainic acid-induced neuronal loss in the rat dentate gyrus. Neuroreport. 1998;9:3075–3079.

17. Azcoitia I, Sierra A, Veiga S, Honda S, Harada N, Garcia-Segura LM. Brain aromatase is neuroprotective. Journal Of Neurobiology. 2001;47:318–329.

18. Bachman DL, Wolf PA, Linn R, Knoefel JE, Cobb J, Belanger A, D'Agostino RB, White LR. Prevalence of dementia and probable senile dementia of the Alzheimer type in the Framingham Study. Neurology. 1992;42:115–119.

19. Bake S, Ma L, Sohrabji F. Estrogen receptor-alpha overexpression suppresses 17beta-estradiol-mediated vascular endothelial growth factor expression and activation of survival kinases. Endocrinology. 2008;149:3881–3889.

20. Bake S, Sohrabji F. 17beta-estradiol differentially regulates blood-brain barrier permeability in young and aging female rats. Endocrinology. 2004;145:5471–5475.

21. Baker LD, Sambamurti K, Craft S, Cherrier M, Raskind MA, Stanczyk FZ, Plymate SR, Asthana S. 17beta-estradiol reduces plasma Abeta40 for HRT-naive postmenopausal women with Alzheimer disease: a preliminary study. Am J Geriatr Psychiatry. 2003;11:239–244.

22. Barnes LL, Wilson RS, Bienias JL, Schneider JA, Evans DA, Bennett DA. Sex differences in the clinical manifestations of Alzheimer disease pathology. Arch Gen Psychiatry. 2005;62:685–691.

23. Barret-Connor E, Kritz-Silverstein D. Estrogen Replacement Therapy and Cognitive Function in Older Women. The Journal of the American Medical Association. 1993;269:2637–2641.

24. Barrett-Connor E, Goodman-Gruen D, Patay B. Endogenous sex hormones and cognitive function in older men. The Journal of Clinical Endocrinology and Metabolism. 1999;84:3681–3685.

25. Baumgartner RN, Waters DL, Gallagher D, Morley JE, Garry PJ. Predictors of skeletal muscle mass in elderly men and women. Mech Ageing Dev. 1999;107:123–136.

26. Behl C, Davis JB, Lesley R, Schubert D. Hydrogen peroxide mediates amyloid b protein toxicity. Cell. 1994;77:817–827.

27. Behl C, Trapp T, Skutella T, Holsboer F. Protection against oxidative stress-induced neuronal cell death--a novel role for RU486. The European Journal of Neuroscience. 1997;9:912–920.

28. Behl C, Widmann M, Trapp T, Holsboer F. 17-beta estradiol protects neurons from oxidative stress-induced cell death in vitro. Biochem Biophys Res Commun. 1995;216:473–482.

29. Benvenuti S, Luciani P, Vannelli GB, Gelmini S, Franceschi E, Serio M, Peri A. Estrogen and selective estrogen receptor modulators exert neuroprotective effects and stimulate the expression of selective Alzheimer's disease indicator-1, a recently discovered antiapoptotic gene, in human neuroblast long-term cell cultures. J Clin Endocrinol Metab. 2005;90:1775–1782.

30. Bezprozvanny I, Mattson MP. Neuronal calcium mishandling and the pathogenesis of Alzheimer's disease. Trends Neurosci. 2008;31:454–463.

31. Bhasin S, Cunningham GR, Hayes FJ, Matsumoto AM, Snyder PJ, Swerdloff RS, Montori VM. Testosterone therapy in adult men with androgen deficiency syndromes:an endocrine society clinical practice guideline. J Clin Endocrinol Metab. 2006;91:1995–2010.

32. Bhasin S, Woodhouse L, Casaburi R, Singh AB, Bhasin D, Berman N, Chen X, Yarasheski KE, Magliano L, Dzekov C, Dzekov J, Bross R, Phillips J, Sinha-Hikim I, Shen R, Storer TW. Testosterone dose-response relationships in healthy young men. Am J Physiol Endocrinol Metab. 2001;281:E1172–E1181.

33. Bimonte-Nelson HA, Francis KR, Umphlet CD, Granholm AC. Progesterone reverses the spatial memory enhancements initiated by tonic and cyclic oestrogen therapy in middle-aged ovariectomized female rats. Eur J Neurosci. 2006;24:229–242.

34. Bimonte-Nelson HA, Nelson ME, Granholm AC. Progesterone counteracts estrogen-induced increases in neurotrophins in the aged female rat brain. Neuroreport. 2004;15:2659–2663.

35. Binder EF, Schechtman KB, Birge SJ, Williams DB, Kohrt WM. Effects of hormone replacement therapy on cognitive performance in elderly women. Maturitas. 2001;38:137–146.

36. Bohacek J, Bearl AM, Daniel JM. Long-term ovarian hormone deprivation alters the ability of subsequent oestradiol replacement to regulate choline acetyltransferase protein levels in the hippocampus and prefrontal cortex of middle-aged rats. J Neuroendocrinol. 2008;20:1023–1027.

37. Bohl CE, Miller DD, Chen J, Bell CE, Dalton JT. Structural basis for accommodation of nonsteroidal ligands in the androgen receptor. J Biol Chem. 2005;280:37747–37754.

38. Braak H, Braak E. Alzheimer's disease: striatal amyloid deposits and neurofibrillary changes. J Neuropathol Exp Neurol. 1990;49:215–224.

39. Brann DW, Dhandapani K, Wakade C, Mahesh VB, Khan MM. Neurotrophic and neuroprotective actions of estrogen: Basic mechanisms and clinical implications. Steroids. 2007;72:381–405.

40. Brayne C, Gill C, Huppert FA, Barkley C, Gehlhaar E, Girling DM, O'Connor DW, Paykel ES. Incidence of clinically diagnosed subtypes of dementia in an elderly population. Cambridge Project for Later Life. Br J Psychiatry. 1995;167:255–262.

41. Brenner DE, Kukull WA, Stergachis A, Belle van G, Bowen JD, McCormick WC, Teri L, Larson EB. Postmenopausal estrogen replacement therapy and the risk of Alzheimer's disease: a population-based case-control study. Am J Epidemiol. 1994;140:262–267.

42. Breuer B, Anderson R. The relationship of tamoxifen with dementia, depression, and dependence in activities of daily living in elderly nursing home residents. Women Health. 2000;31:71–85.

43. Brinton RD. Requirements of a brain selective estrogen: advances and remaining challenges for developing a NeuroSERM. J Alzheimers Dis. 2004;6:S27–S35.

44. Brinton RD, Chen S, Montoya M, Hsieh D, Minaya J. The estrogen replacement therapy of the Women's Health Initiative promotes the cellular mechanisms of memory and neuronal survival in neurons vulnerable to Alzheimer's disease. Maturitas 34. 2000;2 Suppl:S35–S52.

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

46. Brooks BP, Merry DE, Paulson HL, Lieberman AP, Kolson DL, Fischbeck KH. A cell culture model for androgen effects in motor neurons. J Neurochem. 1998;70:1054–1060.

47. Brown TR. Nonsteroidal selective androgen receptors modulators (SARMs): designer androgens with flexible structures provide clinical promise. Endocrinology. 2004;145:5417–5419.

48. Bruce-Keller AJ, Keeling JL, Keller JN, Huang FF, Camondola S, Mattson MP. Antiinflammatory Effects of Estrogen on Microglial Activation. Endocrinology. 2000;141:3646–3656.

49. Bryant HU, Dere WH. Selective estrogen receptor modulators: an alternative to hormone replacement therapy. Proc Soc Exp Biol Med. 1998;217:45–52.

50. Buckwalter JG, Sobel E, Dunn ME, Diz MM, Henderson VW. Gender differences on a brief measure of cognitive functioning in Alzheimer's disease. Archives of Neurology. 1993;50:757–760.

51. Burger H, de Laet CE, van Daele PL, Weel AE, Witteman JC, Hofman A, Pols HA. Risk factors for increased bone loss in an elderly population: the Rotterdam Study. Am J Epidemiol. 1998;147:871–879.

52. Callahan MJ, Lipinski WJ, Bian F, Durham RA, Pack A, Walker LC. Augmented senile plaque load in aged female beta-amyloid precursor protein-transgenic mice. Am J Pathol. 2001;158:1173–1177.

53. Carlson MC, Zandi PP, Plassman BL, Tschanz JT, Welsh-Bohmer KA, Steffens DC, Bastian LA, Mehta KM, Breitner JC. Hormone replacement therapy and reduced cognitive decline in older women: the Cache County Study. Neurology. 2001;57:2210–2216.

54. Carroll JC, Pike CJ. Selective estrogen receptor modulators differentially regulate Alzheimer-like changes in female 3xTg-AD mice. Endocrinology. 2008;149:2607–2611.

55. Carroll JC, Rosario ER, Chang L, Stanczyk FZ, Oddo S, LaFerla FM, Pike CJ. Progesterone and estrogen regulate Alzheimer-like neuropathology in female 3xTg-AD mice. J Neurosci. 2007;27:13357–13365.

56. Carroll JC, Rosario ER, Pike CJ. Progesterone blocks estrogen neuroprotection from kainate in middle-aged female rats. Neurosci Lett. 2008;445:229–232.

57. Castelo-Branco C, Figueras F, Sanjuan A, Pons F, Vicente JJ, Vanrell JA. Long-term postmenopausal hormone replacement therapy effects on bone mass: differences between surgical and spontaneous patients. Eur J Obstet Gynecol Reprod Biol. 1999;83:207–211.

58. Chae HS, Bach JH, Lee MW, Kim HS, Kim YS, Kim KY, Choo KY, Choi SH, Park CH, Lee SH, Suh YH, Kim SS, Lee WB. Estrogen attenuates cell death induced by carboxy-terminal fragment of amyloid precursor protein in PC12 through a receptor-dependent pathway. J Neurosci Res. 2001;65:403–407.

59. Chakraborty TR, Gore AC. Aging-related changes in ovarian hormones, their receptors, and neuroendocrine function. Exp Biol Med (Maywood) 2004;229:977–987.

60. Chang D, Kwan J, Timiras PS. Estrogens influence growth, maturation, and amyloid beta-peptide production in neuroblastoma cells and in a beta-APP transfected kidney 293 cell line. Adv Exp Med and Biol. 1997;429:261–271.

61. Chaovipoch P, Jelks KA, Gerhold LM, West EJ, Chongthammakun S, Floyd CL. 17beta-estradiol is protective in spinal cord injury in post- and pre-menopausal rats. J Neurotrauma. 2006;23:830–852.

62. Cheng IH, Scearce-Levie K, Legleiter J, Palop JJ, Gerstein H, Bien-Ly N, Puolivali J, Lesne S, Ashe KH, Muchowski PJ, Mucke L. Accelerating amyloid-beta fibrillization reduces oligomer levels and functional deficits in Alzheimer disease mouse models. J Biol Chem. 2007;282:23818–23828.

63. Cherrier MM, Asthana S, Plymate S, Baker L, Matsumoto AM, Peskind E, Raskind MA, Brodkin K, Bremner W, Petrova A, LaTendresse S, Craft S. Testosterone supplementation improves spatial and verbal memory in healthy older men. Neurology. 2001;57:80–88.

64. Cherrier MM, Matsumoto AM, Amory JK, Asthana S, Bremner W, Peskind ER, Raskind MA, Craft S. Testosterone improves spatial memory in men with Alzheimer disease and mild cognitive impairment. Neurology. 2005;64:2063–2068.

65. Choi DW. Ionic dependence of glutamate neurotoxicity. J.Neurosci. 1987;7:369–379.

66. Ciana P, Di Luccio G, Belcredito S, Pollio G, Vegeto E, Tatangelo L, Tiveron C, Maggi A. Engineering of a Mouse for the in Vivo Profiling of Estrogen Receptor Activity. Molecular Endocrinology. 2001;15:1104–1113.

67. Corder EH, Ghebremedhin E, Taylor MG, Thal DR, Ohm TG, Braak H. The biphasic relationship between regional brain senile plaque and neurofibrillary tangle distributions: modification by age, sex, and APOE polymorphism. Ann N Y Acad Sci. 2004;1019:24–28.

68. Cordey M, Gundimeda U, Gopalakrishna R, Pike CJ. Estrogen activates protein kinase C in neurons: role in neuroprotection. Journal of Neurochemistry. 2003;84:1340–1348.

69. Cordey M, Pike CJ. Conventional protein kinase C isoforms mediate neuroprotection induced by phorbol ester and estrogen. J Neurochem. 2006;96:204–217.

70. Cordey M, Pike CJ. Neuroprotective properties of selective estrogen receptor agonists in cultured neurons. Brain Res. 2005;1045:217–223.

71. Cory S, Adams JM. The Bcl2 family: regulators of the cellular life-or-death switch. Nat Rev Cancer. 2002;2:647–656.

72. Cotman CW, Anderson AJ. A potential role for apoptosis in neurodegeneration and Alzheimer's disease. Mol Neurobiol. 1995;10:19–45.

73. Cotman CW, Cribbs DH, Pike CJ, Ivins KJ. Cell death in Alzheimer's death. When Cells Die: A Comprehensive Evaluation of Apoptosis and Programmed Cell Death. 1998:385–409.

74. Cotman CW, Pike CJ, Copani A. beta-Amyloid neurotoxicity: a discussion of in vitro findings. Neurobiol Aging. 1992;13:587–590.

75. Coughlan CM, Breen KC. Factors influencing the processing and function of the amyloid beta precursor protein — a potential therapeutic target in Alzheimer's disease? Pharmacol. Ther. 2000;86:111–144.

76. Craig MC, Maki PM, Murphy DG. The Women's Health Initiative Memory Study:findings and implications for treatment. Lancet Neurol. 2005;4:190–194.

77. Cummings DE, Kumar N, Bardin CW, Sundaram K, Bremner WJ. Prostate-sparing effects in primates of the potent androgen 7alpha-methyl-19-nortestosterone: a potential alternative to testosterone for androgen replacement and male contraception. J Clin Endocrinol Metab. 1998;83:4212–4219.

78. Dai D, Wolf DM, Litman ES, White MJ, Leslie KK. Progesterone inhibits human endometrial cancer cell growth and invasiveness: down-regulation of cellular adhesion molecules through progesterone B receptors. Cancer Res. 2002;62:881–886.

79. Daniel JM, Hulst JL, Berbling JL. Estradiol replacement enhances working memory in middle-aged rats when initiated immediately after ovariectomy but not after a long-term period of ovarian hormone deprivation. Endocrinology. 2006;147:607–614.

80. Davies S, Dai D, Wolf DM, Leslie KK. Immunomodulatory and transcriptional effects of progesterone through progesterone A and B receptors in Hec50co poorly differentiated endometrial cancer cells. J Soc Gynecol Investig. 2004;11:494–499.

81. Del Cerro S, Garcia-Estrada J, Garcia-Segura LM. Neuroactive steroids regulate astroglia morphology in hippocampal cultures from adult rats. Glia. 1995;14:65–71.

82. Delmas PD, Bjarnason NH, Mitlak BH, Ravoux AC, Shah AS, Huster WJ, Draper M, Christiansen C. Effects of raloxifene on bone mineral density, serum cholesterol concentrations, and uterine endometrium in postmenopausal women. N Engl J Med. 1997;337:1641–1647.

83. Dluzen DE. Effects of testosterone upon MPTP-induced neurotoxicity of the nigrostriatal dopaminergic system of C57/B1 mice. Brain Research. 1996;715:113–118.

84. Dong L, Wang W, Wang F, Stoner M, Reed JC, Harigai M, Samudio I, Kladde MP, Vyhlidal C, Safe S. Mechanisms of transcriptional activation of bcl-2 gene expression by 17beta-estradiol in breast cancer cells. J Biol Chem. 1999;274:32099–32107.

85. Duara R, Barker WW, Lopez-Alberola R, Loewenstein DA, Grau LB, Gilchrist D, Sevush S, St George-Hyslop S. Alzheimer's disease: interaction of apolipoprotein Egenotype, family history of dementia, gender, education, ethnicity, and age of onset. Neurology. 1996;46:1575–1579.

86. Dubal DB, Shughrue PJ, Wilson ME, Merchenthaler I, Wise PM. Estradiol modulates bcl-2 in cerebral ischemia: a potential role for estrogen receptors. J.Neurosci. 1999;19:6385–6393.

87. Dubal DB, Wise PM. Neuroprotective effects of estradiol in middle-aged female rats. Endocrinology. 2001;142:43–48.

88. Edinger KL, Frye CA. Testosterone's analgesic, anxiolytic, and cognitive-enhancing effects may be due in part to actions of its 5alpha-reduced metabolites in the hippocampus. Behav Neurosci. 2004;118:1352–1364.

89. Edinger KL, Frye CA. Testosterone's anti-anxiety and analgesic effects may be due in part to actions of its 5alpha-reduced metabolites in the hippocampus. Psychoneuroendocrinology. 2005;30:418–430.

90. El Khoury J, Hickman SE, Thomas CA, Cao L, Silverstein SC, Loike JD. Scavenger receptor-mediated adhesion of microglia to beta-amyloid fibrils. Nature. 1996;382:716–719.

91. Emmelot-Vonk MH, Verhaar HJ, Nakhai Pour HR, Aleman A, Lock TM, Bosch JL, Grobbee DE, van der Schouw YT. Effect of testosterone supplementation on functional mobility, cognition, and other parameters in older men: a randomized controlled trial. Jama. 2008;299:39–52.

92. Espeland MA, Rapp SR, Shumaker SA, Brunner R, Manson JE, Sherwin BB, Hsia J, Margolis KL, Hogan PE, Wallace R, Dailey M, Freeman R, Hays J. Conjugated equine estrogens and global cognitive function in postmenopausal women: Women's Health Initiative Memory Study. Jama. 2004;291:2959–2968.

93. Estrada M, Uhlen P, Ehrlich BE. Ca2+ oscillations induced by testosterone enhance neurite outgrowth. J Cell Sci. 2006;119:733–743.

94. Farrer LA, Cupples LA, Haines JL, Hyman B, Kukull WA, Mayeux R, Myers RH, Pericak-Vance MA, Risch N, van Duijn CM. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease. A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. Jama. 1997;278:1349–1356.

95. Feldman HA, Longcope C, Derby CA, Johannes CB, Araujo AB, Coviello AD, Bremner WJ, McKinlay JB. Age Trends in the Level of Serum Testosterone and Other Hormones in Middle-Aged Men: Longitudinal Results from the Massachusetts Male Aging Study. Journal of clinical endocrinology and metabolism. 2002;87:589–598.

96. Felicio LS, Nelson JF, Finch CE. Longitudinal studies of estrous cyclicity in aging C57BL/6J mice: II. Cessation of cyclicity and the duration of persistent vaginal cornification. Biol Reprod. 1984;31:446–453.

97. Ferrando AA, Sheffield-Moore M, Yeckel CW, Gilkison C, Jiang J, Achacosa A, Lieberman SA, Tipton K, Wolfe RR, Urban RJ. Testosterone administration to older men improves muscle function: molecular and physiological mechanisms. Am J Physiol Endocrinol Metab. 2002;282:E601–E607.

98. Fitzpatrick JL, Mize AL, Wade CB, Harris JA, Shapiro RA, Dorsa DM. Estrogen-mediated neuroprotection against beta-amyloid toxicity requires expression of estrogen receptor alpha or beta and activation of the MAPK pathway. Journal of Neurochemistry. 2002;82:674–682.

 99. Fleisher A, Grundman M, Jack CR, Jr, Petersen RC, Taylor C, Kim HT, Schillfer DH, Bagwell V, Sencakova D, Weiner MF, DeCarli C, DeKosky ST, van Dyck CH, Thal LJ. Sex, apolipoprotein E epsilon 4 status, and hippocampal volume in mild cognitive impairment. Arch Neurol. 2005;62:953–957.

100. Fleshner N, Zlotta AR. Prostate cancer prevention: past, present, and future. Cancer. 2007;110:1889–1899.

101. Frankfurt M, Gould E, Woolley CS, McEwen BS. Gonadal steroids modify dendritic spine density in ventromedial hypothalamic neurons: a Golgi study in the adult rat. Neuroendocrinology. 1990;51:530–535.

102. Fratiglioni L, Viitanen M, von Strauss E, Tontodonati V, Herlitz A, Winblad B. Very old women at highest risk of dementia and Alzheimer's disease: incidence data from the Kungsholmen Project, Stockholm. Neurology. 1997;48:132–138.

103. Frye CA, Reed TA. Androgenic neurosteroids: anti-seizure effects in an animal model of epilepsy. Psychoneuroendocrinology. 1998;23:385–399.

104. Frye CA, Seliga AM. Testosterone increases analgesia, anxiolysis, and cognitive performance of male rats. Cogn Affect Behav Neurosci. 2001;1:371–381.

105. Frye CA, Walf AA. Effects of progesterone administration and APPswe+PSEN1Deltae9 mutation for cognitive performance of mid-aged mice. Neurobiol Learn Mem. 2008;89:17–26.

106. Fukuzaki E, Takuma K, Funatsu Y, Himeno Y, Kitahara Y, Gu B, Mizoguchi H, Ibi D, Koike K, Inoue M, Yan SD, Yamada K. Ovariectomy increases neuronal amyloid-beta binding alcohol dehydrogenase level in the mouse hippocampus. Neurochem Int. 2008;52:1358–1364.

107. Gandy S, Almeida OP, Fonte J, Lim D, Waterrus A, Spry N, Flicker L, Martins RN. Chemical andropause and amyloid-beta peptide. Journal of the American Medical Association. 2001;285:2195–2196.

108. Gandy S, Petanceska S. Regulation of alzheimer beta-amyloid precursor trafficking and metabolism. Adv Exp Med Biol. 2001;487:85–100.

109. Gao W, Reiser PJ, Coss CC, Phelps MA, Kearbey JD, Miller DD, Dalton JT. Selective androgen receptor modulator treatment improves muscle strength and body composition and prevents bone loss in orchidectomized rats. Endocrinology. 2005;146:4887–4897.

110. Garcia-Estrada J, Del Rio JA, Luquin S, Soriano E, Garcia-Segura LM. Gonadal hormones down-regulate reactive gliosis and astrocyte proliferation after a penetrating brain injury. Brain Research. 1993;628:271–278.

111. Garcia-Segura LM, Cardona-Gomez P, Naftolin F, Chowen JA. Estradiol upregulates Bcl-2 expression in adult brain neurons. Neuroreport. 1998;9:593–597.

112. Garcia-Segura LM, Wozniak A, Azcoitia I, Rodriguez JR, Hutchison RE, Hutchison JB. Aromatase expression by astrocytes after brain injury: implications for local estrogen formation in brain repair. Neuroscience. 1999;89:567–578.

113. Gatson JW, Kaur P, Singh M. Dihydrotestosterone differentially modulates the mitogen-activated protein kinase and the phosphoinositide 3-kinase/Akt pathways through the nuclear and novel membrane androgen receptor in C6 cells. Endocrinology. 2006;147:2028–2034.

114. Gatson JW, Singh M. Activation of a membrane-associated androgen receptor promotes cell death in primary cortical astrocytes. Endocrinology. 2007;148:2458–2464.

115. Gee DM, Flurkey K, Finch CE. Aging and the regulation of luteinizing hormone in C57BL/6J mice: impaired elevations after ovariectomy and spontaneous elevations at advanced ages. Biol Reprod. 1983;28:598–607.

116. Gibbs RB. Effects of gonadal hormone replacement on measures of basal forebrain cholinergic function. Neuroscience. 2000;101:931–938.

117. Gibbs RB. Oestrogen and the cholinergic hypothesis: implications for oestrogen replacement therapy in postmenopausal women. Novartis Found Symp. 2000;230:94–107. discussion 107–111.

118. Gillett MJ, Martins RN, Clarnette RM, Chubb SA, Bruce DG, Yeap BB. Relationship between testosterone, sex hormone binding globulin and plasma amyloid beta peptide 40 in older men with subjective memory loss or dementia. J Alzheimers Dis. 2003;5:267–269.

119. Gleason CE, Carlsson CM, Johnson S, Atwood C, Asthana S. Clinical pharmacology and differential cognitive efficacy of estrogen preparations. Ann N Y Acad Sci. 2005;1052:93–115.

120. Goebel JA, Birge SJ, Price SC, Hanson JM, Fishel DG. Estrogen replacement therapy and postural stability in the elderly. Am J Otol. 1995;16:470–474.

121. Golub MS, Germann SL, Mercer M, Gordon MN, Morgan DG, Mayer LP, Hoyer PB. Behavioral consequences of ovarian atrophy and estrogen replacement in the APPswe mouse. Neurobiology of Aging. 2007;29:1512–1523.

122. Goodenough S, Engert S, Behl C. Testosterone stimulates rapid secretory amyloid precursor protein release from rat hypothalamic cells via the activation of the mitogen-activated protein kinase pathway. Neuroscience Letters. 2000;296:49–52.

123. Goodenough S, Schleusner D, Pietrzik C, Skutella T, Behl C. Glycogen synthase kinase 3beta links neuroprotection by 17beta-estradiol to key Alzheimer processes. Neuroscience. 2005;132:581–589.

124. Goodman Y, Bruce AJ, Cheng B, Mattson MP. Estrogens attenuate and corticosterone exacerbates excitotoxicity, oxidative injury, and amyloid beta-peptide toxicity in hippocampal neurons. J Neurochem. 1996;66:1836–1844.

125. Goodman Y, Mattson MP. Secreted Forms of b-Amyloid Precursor Protein Protect Hippocampal Neurons against Amyloid b-Peptide-Induced Oxidative Injury. Experimental Neurology. 1994;128:1–12.

126. Gooren L. Androgen deficiency in the aging male: benefits and risks of androgen supplementation. J Steroid Biochem Mol Biol. 2003;85:349–355.

127. Gosden RG, Laing SC, Felicio LS, Nelson JF, Finch CE. Imminent oocyte exhaustion and reduced follicular recruitment mark the transition to acyclicity in aging C57BL/6J mice. Biol Reprod. 1983;28:255–260.

128. Gouchie C, Kimura D. The relationship between testosterone levels and cognitive ability patterns. Psychoneuroendocrinology. 1991;16:323–334.

129. Gouras GK, Xu H, Gross RS, Greenfield JP, Hai B, Wang R, Greengard P. Testosterone reduces neuronal secretion of Alzheimer's beta-amyloid peptides. PNAS. 2000;97:1202–1205.

130. Grady D, Gebretsadik T, Kerlikowske K, Ernster V, Petitti D. Hormone replacement therapy and endometrial cancer risk: a meta-analysis. Obstet Gynecol. 1995;85:304–313.

131. Graham JD, Clarke CL. Physiological action of progesterone in target tissues. Endocr Rev. 1997;18:502–519.

132. Gray A, Feldman H, McKinlay J, Longcope C. Age disease and changing sex hormone levels in middle-aged men: results of the Massachusetts Male Aging Study. Journal of clinical endocrinology and metabolism. 1991;73:1016–1025.

133. Green PS, Bales K, Paul S, Bu GJ. Estrogen therapy fails to alter amyloid deposition in the PDAPP model of Alzheimer's disease. Endocrinology. 2005;146:2774–2781.

134. Green PS, Gridley KE, Simpkins JW. Estradiol protects against beta-amyloid (25–35)-induced toxicity in SK-N-SH human neuroblastoma cells. Neuroscience Letters. 1996;218:165–168.

135. Greenfield JP, Leung LW, Cai D, Kaasik K, Gross RS, Rodriguez_Boulan E, Greengard P, Xu H. Estrogen lowers Alzheimer beta-amyloid generation by stimulating trans-Golgi network vesicle biogenesis. J Biol Chem. 2002;277:12128–12136.

136. Gridley KE, Green PS, Simpkins JW. Low concentrations of estradiol reduce b-amyloid (25–35)-induced toxicity, lipid peroxidation and glucose utilization in human SK-N-SH neuroblastoma cells. Brain Research. 1997;778:158–165.

137. Haass C, Selkoe DJ. Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer–s amyloid beta-peptide. Nat Rev Mol Cell Biol. 2007;8:101–112.

138. Hagnell O, Ojesjo L, Rorsman B. Incidence of dementia in the Lundby Study. Neuroepidemiology. 1992;11 Suppl 1:61–66.

139. Hammond J, Le Q, Goodyer C, Gelfand M, Trifiro M, LeBlanc A. Testosterone-mediated neuroprotection through the androgen receptor in human primary neurons. Journal of Neurochemistry. 2001;77:1319–1326.

140. Hardy J. The Alzheimer family of diseases: many etiologies, one pathogenesis? Proc Natl Acad Sci U S A. 1997;94:2095–2097.

141. Hardy JA, Higgins GA. Alzheimer' disease: the amyloid cascade hypothesis. Science. 1992;256:184–185.

142. Haren MT, Kim MJ, Tariq SH, Wittert GA, Morley JE. Andropause: a quality-of-life issue in older males. Med Clin North Am. 2006;90:1005–1023.

143. Haren MT, Wittert GA, Chapman IM, Coates P, Morley JE. Effect of oral testosterone undecanoate on visuospatial cognition, mood and quality of life in elderly men with low-normal gonadal status. Maturitas. 2005;50:124–133.

144. Harris-White ME, Chu T, Miller SA, Simmons M, Teter B, Nash D, Cole GM, Frautschy SA. Estrogen (E2) and glucocorticoid (Gc) effects on microglia and Ab clearance in vitro and in vivo. Neurochemistry International. 2001;39:435–448.

145. Hart SA, Patton JD, Woolley CS. Quantitative analysis of ERalpha and GAD colocalization in the hippocampus of the adult female rat. The Journal of Comparative Neurology. 2001;440:144–155.

146. Hartley DM, Walsh DM, Ye CP, Diehl T, Vasquez S, Vassilev PM, Teplow DB, Selkoe DJ. Protofibrillar intermediates of amyloid beta-protein induce acute electrophysiological changes and progressive neurotoxicity in cortical neurons. J Neurosci. 1999;19:8876–8884.

147. Haskell SG, Richardson ED, Horwitz RI. The effect of estrogen replacement therapy on cognitive function in women: A critical review of the literature. Journal of Clinical Epidemiology. 1997;50:1249–1264.

148. He B, Minges JT, Lee LW, Wilson EM. The FXXLF motif mediates androgen receptor-specific interactions with coregulators. J Biol Chem. 2002;277:10226–10235.

149. Heikkinen T, Kalesnykas G, Rissanen A, Tapiola T, Iivonen S, Wang J, Chaudhuri J, Tanila H, Miettinen R, Puolivali J. Estrogen treatment improves spatial learning in APP + PS1 mice but does not affect beta amyloid accumulation and plaque formation. Exp Neurol. 2004;187:105–117.

150. Henderson V, Buckwalter J. Cognitive deficits of men and women with Alzheimer's disease. Neurology. 1994;44:90–96.

151. Henderson VW. Estrogen-containing hormone therapy and Alzheimer's disease risk: understanding discrepant inferences from observational and experimental research. Neuroscience. 2006;138:1031–1039.

152. Henderson VW, Benke KS, Green RC, Cupples LA, Farrer LA. Postmenopausal hormone therapy and Alzheimer's disease risk: interaction with age. J Neurol Neurosurg Psychiatry. 2005;76:103–105.

153. Henderson VW, Guthrie JR, Dudley EC, Burger HG, Dennerstein L. Estrogen exposures and memory at midlife: a population-based study of women. Neurology. 2003;60:1369–1371.

154. Henderson VW, Paganini_Hill A, Miller BL, Elble RJ, Reyes PF, Shoupe D, McCleary CA, Klein RA, Hake AM, Farlow MR. Estrogen for Alzheimer's disease in women: randomized, double-blind, placebo-controlled trial. Neurology. 2000;54:295–301.

155. Henderson VW, Paganini-Hill A, Emanuel CK, Dunn ME, Buckwalter JG. Estrogen replacement therapy in older women. Comparisons between Alzheimer's disease cases and nondemented control subjects. Arch Neurol. 1994;51:896–900.

156. Hestiantoro A, Swaab DF. Changes in estrogen receptor-alpha and -beta in the infundibular nucleus of the human hypothalamus are related to the occurrence of Alzheimer's disease neuropathology. J Clin Endocrinol Metab. 2004;89:1912–1925.

157. Hirata-Fukae C, Li HF, Hoe HS, Gray AJ, Minami SS, Hamada K, Niikura T, Hua F, Tsukagoshi-Nagai H, Horikoshi-Sakuraba Y, Mughal M, Rebeck GW, LaFerla FM, Mattson MP, Iwata N, Saido TC, Klein WL, Duff KE, Aisen PS, Matsuoka Y. Females exhibit more extensive amyloid, but not tau, pathology in an Alzheimer transgenic model. Brain Res. 2008;1216:92–103.

158. Hlatky MA, Boothroyd D, Vittinghoff E, Sharp P, Whooley MA. Quality-of-life and depressive symptoms in postmenopausal women after receiving hormone therapy: results from the Heart and Estrogen/Progestin Replacement Study (HERS) trial. Jama. 2002;287:591–597.

159. Hogervorst E, Boshuisen M, Riedel W, Willeken C, Jolles J. 1998 Curt P. Richter Award. The effect of hormone replacement therapy on cognitive function in elderly women. Psychoneuroendocrinology. 1999;24:43–68.

160. Hogervorst E, Combrinck M, Smith AD. Testosterone and gonadotropin levels in men with dementia. Neuroendocrinology Letters. 2003;24:203–208.

161. Hogervorst E, Lehmann DJ, Warden DR, McBroom J, Smith AD. Apolipoprotein E epsilon4 and testosterone interact in the risk of Alzheimer's disease in men. Int. J. Ger.Psych. 2002;17:938–940.

162. Hogervorst E, Williams J, Biudge M, Barnetson L, Combrinck M, Smith AD. Serum total testosterone is lower in men with Alzheimer's disease. Neuro Endocrinol Lett. 2001;22:163–168.

163. Honjo H, Iwasa K, Kawata M, Fushiki S, Hosoda T, Tatsumi H, Oida N, Mihara M, Hirasugi Y, Yamamoto H, Kikuchi N, Kitawaki J. Progestins and estrogens and Alzheimer's disease. J Steroid Biochem Mol Biol. 2005;93:305–308.

164. Huang J, Guan H, Booze RM, Eckman CB, Hersh LB. Estrogen regulates neprilysin activity in rat brain. Neurosci Lett. 2004;367:85–87.

165. Huppenbauer CB, Tanzer L, DonCarlos LL, Jones KJ. Gonadal steroid attenuation of developing hamster facial motoneuron loss by axotomy: equal efficacy of testosterone, dihydrotestosterone, and 17-beta estradiol. J Neurosci. 2005;25:4004–4013.

166. Hyman BT, Gomez-Isla T, Briggs M, Chung H, Nichols S, Kohout F, Wallace R. Apolipoprotein E and cognitive change in an elderly population. Ann Neurol. 1996;40:55–66.

167. Iqbal K, Alonso Adel C, Chen S, Chohan MO, El-Akkad E, Gong CX, Khatoon S, Li B, Liu F, Rahman A, Tanimukai H, Grundke-Iqbal I. Tau pathology in Alzheimer disease and other tauopathies. Biochim Biophys Acta. 2005;1739:198–210.

168. Irwin RW, Yao J, Hamilton RT, Cadenas E, Brinton RD, Nilsen J. Progesterone and estrogen regulate oxidative metabolism in brain mitochondria. Endocrinology. 2008;149:3167–3175.

169. Islander U, Erlandsson MC, Hasseus B, Jonsson CA, Ohlsson C, Gustafsson JA, Dahlgren U, Carlsten H. Influence of oestrogen receptor alpha and beta on the immune system in aged female mice. Immunology. 2003;110:149–157.

170. Ivins KJ, Bui ET, Cotman CW. Beta-amyloid induces local neurite degeneration in cultured hippocampal neurons: evidence for neuritic apoptosis. Neurobiol Dis. 1998;5:365–378.

171. Iwata N, Tsubuki S, Takaki Y, Watanabe K, Sekiguchi M, Hosoki E, Kawashima-Morishima M, Lee HJ, Hama E, Sekine-Aizawa Y, Saido TC. Identification of the major Abeta1-42-degrading catabolic pathway in brain parenchyma: suppression leads to biochemical and pathological deposition. Nat Med. 2000;6:143–150.

172. Jaffe AB, Toran_Allerand CD, Greengard P, Gandy SE. Estrogen regulates metabolism of Alzheimer amyloid beta precursor protein. J Biol Chem. 1994;269:13065–13068.

173. Janowsky JS, Chavez B, Orwoll E. Sex steroids modify working memory. Journal of Cognitive Neuroscience. 2000;12:407–414.

174. Janowsky JS, Oviatt SK, Orwoll ES. Testosterone Influences Spatial Cognition in Older Men. Behavioral Neuroscience. 1994;108:325–332.

175. Jayaraman A, Pike CJ. Progesterone attenuates oestrogen neuroprotection via downregulation of oestrogen receptor expression in cultured neurones. J Neuroendocrinol. 2009;21:77–81.

176. Jezierski MK, Sohrabji F. Neurotrophin expression in the reproductively senescent forebrain is refractory to estrogen stimulation. Neurobiol. Aging. 2001;22:309–319.

177. Johnson AB, Bake S, Lewis DK, Sohrabji F. Temporal expression of IL-1beta protein and mRNA in the brain after systemic LPS injection is affected by age and estrogen. J Neuroimmunol. 2006;174:82–91.

178. Jones KJ, Brown TJ, Damaser M. Neuroprotective effects of gonadal steroids on regenerating peripheral motoneurons. Brain Res Brain Res Rev. 2001;37:372–382.

179. Jones KJ, Coers S, Storer PD, Tanzer L, Kinderman NB. Androgenic regulation of the central glia response following nerve damage. Journal of Neurobiology. 1999;40:560–573.

180. Jones RD, Pugh PJ, Hall J, Channer KS, Jones TH. Altered circulating hormone levels, endothelial function and vascular reactivity in the testicular feminised mouse. European Journal of Endocrinology / European Federation of Endocrine Societies. 2003;148:111–120.

181. Jorm AF, Korten AE, Henderson AS. The prevalence of dementia: a quantitative integration of the literature. Acta Psychiatr Scand. 1987;76:465–479.

182. Kalesnykas G, Roschier U, Puolivali J, Wang J, Miettinen R. The effect of aging on the subcellular distribution of estrogen receptor-alpha in the cholinergic neurons of transgenic and wild-type mice. Eur J Neurosci. 2005;21:1437–1442.

183. Kamenetz F, Tomita T, Hsieh H, Seabrook G, Borchelt D, Iwatsubo T, Sisodia S, Malinow R. APP processing and synaptic function. Neuron. 2003;37:925–937.

184. Kaufman JM, Vermeulen A. The decline of androgen levels in elderly men and its clinical and therapeutic implications. Endocr Rev. 2005

185. Kawas C, Resnick S, Morrison A, Brookmeyer R, Corrada M, Zonderman A, Bacal C, Donnell Lingle D, Metter E. A prospective study of estrogen replacement therapy and the risk of developing Alzheimer's disease: The Baltimore Longitudinal Study of Aging. Neurology. 1997;48:1517–1521.

186. Kelly JF, Bienias JL, Shah A, Meeke KA, Schneider JA, Soriano E, Bennett DA. Levels of estrogen receptors alpha and beta in frontal cortex of patients with Alzheimer's disease: relationship to Mini-Mental State Examination scores. Curr Alzheimer Res. 2008;5:45–51.

187. Kilbourne EJ, Moore WJ, Freedman LP, Nagpal S. Selective androgen receptor modulators for frailty and osteoporosis. Curr Opin Investig Drugs. 2007;8:821–829.

188. Kim H, Bang OY, Jung MW, Ha SD, Hong HS, Huh K, Kim SU, Mook-Jung I. Neuroprotective effects of estrogen against beta-amyloid toxicity are mediated by estrogen receptors in cultured neuronal cells. Neurosci Lett. 2001;302:58–62.

189. Kim J, Wu D, Hwang DJ, Miller DD, Dalton JT. The para substituent of S-3-(phenoxy)-2-hydroxy-2-methyl-N-(4-nitro-3-trifluoromethyl-phenyl)-prop ionamides is a major structural determinant of in vivo disposition and activity of selective androgen receptor modulators. J Pharmacol Exp Ther. 2005;315:230–239.

190. Koh JY, Yang LL, Cotman CW. Beta-amyloid protein increases the vulnerability of cultured cortical neurons to excitotoxic damage. Brain Research. 1990;533:315–320.

191. Koo E. Phorbol esters affect multiple steps in beta-amyloid precursor protein trafficking and amyloid beta-protein production. Mol Med. 1997;3:204–211.

192. Koski CL, Hila S, Hoffman GE. Regulation of cytokine-induced neuron death by ovarian hormones: involvement of antiapoptotic protein expression and c-JUN N-terminal kinase-mediated proapoptotic signaling. Endocrinology. 2004;145:95–103.

193. Kujawa KA, Emeric E, Jones KJ. Testosterone differentially regulates the regenerative properties of injured hamster facial motoneurons. J Neurosci. 1991;11:3898–3906.

194. Kujawa KA, Jacob JM, Jones KJ. Testosterone regulation of the regenerative properties of injured rat sciatic motor neurons. J Neurosci Res. 1993;35:268–273.

195. Kujawa KA, Kinderman NB, Jones KJ. Testosterone-induced acceleration of recovery from facial paralysis following crush axotomy of the facial nerve in male hamsters. Exp Neurol. 1989;105:80–85.

196. Kujawa KA, Tanzer L, Jones KJ. Inhibition of the accelerative effects of testosterone on hamster facial nerve regeneration by the antiandrogen flutamide. Exp Neurol. 1995;133:138–143.

197. LaFerla FM, Oddo S. Alzheimer's disease: Abeta, tau and synaptic dysfunction. Trends Mol Med. 2005;11:170–176.

198. LeBlanc ES, Janowsky J, Chan BKS, Nelson HD. Hormone replacement therapy and cognition: Systematic review and meta-analysis. Journal of the American Medical Association. 2001;285:1489.

199. Lesne S, Koh MT, Kotilinek L, Kayed R, Glabe CG, Yang A, Gallagher M, Ashe KH. A specific amyloid-beta protein assembly in the brain impairs memory. Nature. 2006;440:352–357.

200. Levin-Allerhand JA, Lominska CE, Wang J, Smith JD. 17Alpha-estradiol and 17beta-estradiol treatments are effective in lowering cerebral amyloid-beta levels in AbetaPPSWE transgenic mice. J Alzheimers Dis. 2002;4:449–457.

201. Li JJ, Sutton JC, Nirschl A, Zou Y, Wang H, Sun C, Pi Z, Johnson R, Krystek SR, Jr, Seethala R, Golla R, Sleph PG, Beehler BC, Grover GJ, Fura A, Vyas VP, Li CY, Gougoutas JZ, Galella MA, Zahler R, Ostrowski J, Hamann LG. Discovery of Potent and Muscle Selective Androgen Receptor Modulators through Scaffold Modifications. J Med Chem. 2007;50:3015–3025.

202. Li R, Shen Y, Yang LB, Lue LF, Finch C, Rogers J. Estrogen enhances uptake of amyloid beta-protein by microglia derived from the human cortex. Journal of Neurochemistry. 2000;75:1447–1454.

203. Liehr JG, Roy D. Free radical generation by redox cycling of estrogens. Free Radical Biology and Medicine. 1990;8:415–423.

204. Linford N, Wade C, Dorsa D. The rapid effects of estrogen are implicated in estrogen-mediated neuroprotection. J Neurocytol. 2000;29:367–374.

205. Liu XA, Zhu LQ, Zhang Q, Shi HR, Wang SH, Wang Q, Wang JZ. Estradiol attenuates tau hyperphosphorylation induced by upregulation of protein kinase-A. Neurochem Res. 2008;33:1811–1820.

206. Lu PH, Masterman DA, Mulnard R, Cotman C, Miller B, Yaffe K, Reback E, Porter V, Swerdloff R, Cummings JL. Effects of testosterone on cognition and mood in male patients with mild Alzheimer disease and healthy elderly men. Arch Neurol. 2006;63:177–185.

207. Lu YP, Zeng M, Swaab DF, Ravid R, Zhou JN. Colocalization and alteration of estrogen receptor-alpha and -beta in the hippocampus in Alzheimer's disease. Hum Pathol. 2004;35:275–280.

208. MacLusky NJ, Hajszan T, Leranth C. Effects of dehydroepiandrosterone and flutamide on hippocampal CA1 spine synapse density in male and female rats: implications for the role of androgens in maintenance of hippocampal structure. Endocrinology. 2004;145:4154–4161.

209. MacLusky NJ, Hajszan T, Prange-Kiel J, Leranth C. Androgen modulation of hippocampal synaptic plasticity. Neuroscience. 2006;138:957–965.

210. Maki P. Hormone therapy and risk for dementia: where do we go from here? Gynecological Endocrinology. 2003;19:354.

211. Maki PM, Ernst M, London ED, Mordecai KL, Perschler P, Durso SC, Brandt J, Dobs A, Resnick SM. Intramuscular testosterone treatment in elderly men: evidence of memory decline and altered brain function. J Clin Endocrinol Metab. 2007;92:4107–4114.

212. Manfredi MC, Bi Y, Nirschl AA, Sutton JC, Seethala R, Golla R, Beehler BC, Sleph PG, Grover GJ, Ostrowski J, Hamann LG. Synthesis and SAR of tetrahydropyrrolo[1,2-b][1,2,5]thiadiazol-2(3H)-one 1,1-dioxide analogues as highly potent selective androgen receptor modulators. Bioorg Med Chem Lett. 2007;17:4487–4490.

213. Manly JJ, Merchant CA, Jacobs DM, Small S, Bell K, Ferin M, Mayeux R. Endogenous estrogen levels and Alzheimer's disease among postmenopausal women. Neurology. 2000;54:833–837.

214. Manthey D, Heck S, Engert S, Behl C. Estrogen induces a rapid secretion of amyloid b precursor protein via the mitogen-activated protein kinase pathway. European Journal of Biochemistry. 2001;268:4285–4291.

215. Mark R, Hensley K, Butterfield D, Mattson M. Amyloid beta-peptide impairs ion-motive ATPase activities: evidence for a role in loss of neuronal Ca2+ homeostasis and cell death. Journal of Neuroscience. 1995;15:6239–6249.

216. Mattson MP, Lovell MA, Furukawa K, Markesbery WR. Neurotrophic factors attenuate glutamate-induced accumulation of peroxides, elevation of intracellular Ca2+ concentration, and neurotoxicity and increase antioxidant enzyme activities in hippocampal neurons. J.Neurochem. 1995;65:1740–1751.

217. Mattson MP, Partin J, Begley JG. Amyloid beta-peptide induces apoptosis-related events in synapses and dendrites. Brain Res. 1998;807:167–176.

218. Meier DE, Orwoll ES, Keenan EJ, Fagerstrom RM. Marked decline in trabecular bone mineral content in healthy men with age: lack of association with sex steroid levels. J Am Geriatr Soc. 1987;35:189–197.

219. Mize AL, Young LJ, Alper RH. Uncoupling of 5-HT1A receptors in the brain by estrogens: regional variations in antagonism by ICI 182,780. Neuropharmacology. 2003;44:584–591.

220. Moffat SD, Zonderman AB, Metter EJ, Blackman MR, Harman SM, Resnick SM. Longitudinal assessment of serum free testosterone concentration predicts memory performance and cognitive status in elderly men. J Clin Endo Metab. 2002;87:5001–5007.

221. Moffat SD, Zonderman AB, Metter EJ, Kawas C, Blackman MR, Harman SM, Resnick SM. Free testosterone and risk for Alzheimer disease in older men. Neurology. 2004;62:188–193.

222. Molsa PK, Marttila RJ, Rinne UK. Epidemiology of dementia in a Finnish population. Acta Neurol Scand. 1982;65:541–552.

223. Mook-Jung I, Joo I, Sohn S, Kwon HJ, Huh K, Jung MW. Estrogen blocks neurotoxic effects of beta-amyloid (1–42) and induces neurite extension on B103 cells. Neurosci Lett. 1997;235:101–104.

224. Morali G, Lemus AE, Munguia R, Arteaga M, Perez-Palacios G, Sundaram K, Kumar N, Bardin CW. Induction of male sexual behavior in the rat by 7 alpha-methyl-19-nortestosterone, an androgen that does not undergo 5 alpha-reduction. Biol Reprod. 1993;49:577–581.

225. Morley JE. Androgens and aging. Maturitas. 2001;38:61–71. discussion 71-63.

226. Morley JE, Kaiser FE, Perry HM, Patrick P, Morley PM, Stauber PM, Vellas B, Baumgartner RN, Garry PJ. Longitudinal changes in testosterone, luteinizing hormone, and follicle-stimulating hormone in healthy older men. Metabolism: Clinical and Experimental. 1997;46:410–413.

227. Morley JE, Perry HM. Androgen treatment of male hypogonadism in older males. The Journal of Steroid Biochemistry and Molecular Biology. 2003;85:367–373.

228. Muller M, den Tonkelaar I, Thijssen JH, Grobbee DE, van der Schouw YT. Endogenous sex hormones in men aged 40–80 years. Eur J Endocrinol. 2003;149:583–589.

229. Mulnard RA, Cotman CW, Kawas C, van Dyck CH, Sano M, Doody R, Koss E, Pfeiffer E, Jin S, Gamst A, Grundman M, Thomas R, Thal LJ. Estrogen replacement therapy for treatment of mild to moderate Alzheimer disease: A randomized controlled trial. The Journal of the American Medical Association. 2000;283:1007.

230. Nguyen ST, Prakash R, Anderson CJ, Frydenberg M, Haynes JM. Sex steroids modulate alpha 1-adrenoceptor-stimulated Ca2+ elevation in human cultured prostatic stromal cells. Prostate. 2007;67:74–82.

231. Nguyen T-VV, Yao M, Pike CJ. Androgens activate mitogen-activated protein kinase signaling: Role in neuroprotection. Journal of Neurochemistry. 2005;94:1639–1651.

232. Nguyen TV, Yao M, Pike CJ. Flutamide and cyproterone acetate exert agonist effects: induction of androgen receptor-dependent neuroprotection. Endocrinology. 2007;148:2936–2943.

233. Nilsen J, Brinton RD. Divergent impact of progesterone and medroxyprogesterone acetate (Provera) on nuclear mitogen-activated protein kinase signaling. Proc Natl Acad Sci U S A. 2003;100:10506–10511.

234. Nilsen J, Brinton RD. Impact of progestins on estradiol potentiation of the glutamate calcium response. Neuroreport. 2002;13:825–830.

235. Nilsen J, Chen S, Brinton RD. 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.

236. Nilsen J, Diaz_Brinton R. Mechanism of estrogen-mediated neuroprotection: regulation of mitochondrial calcium and Bcl-2 expression. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:2842–2847.

237. Nilsen J, Mor G, Naftolin F. Raloxifene induces neurite outgrowth in estrogen receptor positive PC12 cells. Menopause. 1998;5:211–216.

238. Nordell VL, Scarborough MM, Buchanan AK, Sohrabji F. Differential effects of estrogen in the injured forebrain of young adult and reproductive senescent animals. Neurobiol.Aging. 2003;24:733–743.

239. O'Neill K, Chen S, Brinton RD. Impact of the selective estrogen receptor modulator, raloxifene, on neuronal survival and outgrowth following toxic insults associated with aging and Alzheimer's disease. Exp Neurol. 2004;185:63–80.

240. Oddo S, Caccamo A, Kitazawa M, Tseng BP, LaFerla FM. Amyloid deposition precedes tangle formation in a triple transgenic model of Alzheimer's disease. Neurobiol Aging. 2003;24:1063–1070.

241. Ogiue-Ikeda M, Tanabe N, Mukai H, Hojo Y, Murakami G, Tsurugizawa T, Takata N, Kimoto T, Kawato S. Rapid modulation of synaptic plasticity by estrogens as well as endocrine disrupters in hippocampal neurons. Brain Res Rev. 2008;57:363–375.

242. Ohkura T, Isse K, Akazawa K, Hamamoto M, Yaoi Y, Hagino N. Evaluation of estrogen treatment in female patients with dementia of the Alzheimer type. Endocr J. 1994;41:361–371.

243. Olariu A, Yamada K, Nabeshima T. Amyloid pathology and protein kinase C (PKC): possible therapeutics effects of PKC activators. J Pharmacol Sci. 2005;97:1–5.

244. Orlando R, Caruso A, Molinaro G, Motolese M, Matrisciano F, Togna G, Melchiorri D, Nicoletti F, Bruno V. Nanomolar concentrations of anabolic-androgenic steroids amplify excitotoxic neuronal death in mixed mouse cortical cultures. Brain Res. 2007;1165:21–29.

245. Ostrowski J, Kuhns JE, Lupisella JA, Manfredi MC, Beehler BC, Krystek SR, Jr, Bi Y, Sun C, Seethala R, Golla R, Sleph PG, Fura A, An Y, Kish KF, Sack JS, Mookhtiar KA, Grover GJ, Hamann LG. Pharmacological and x-ray structural characterization of a novel selective androgen receptor modulator: potent hyperanabolic stimulation of skeletal muscle with hypostimulation of prostate in rats. Endocrinology. 2007;148:4–12.

246. Paganini-Hill A, Clark LJ. Preliminary assessment of cognitive function in breast cancer patients treated with tamoxifen. Breast Cancer Res Treat. 2000;64:165–176.

247. Paganini-Hill A, Henderson VW. Estrogen deficiency and risk of Alzheimer's disease in women. American Journal of Epidemiology. 1994;140:256–261.

248. Paganini-Hill A, Henderson VW. Estrogen replacement therapy and risk of Alzheimer disease. Arch Intern Med. 1996;156:2213–2217.

249. Paoletti AM, Congia S, Lello S, Tedde D, Orru M, Pistis M, Pilloni M, Zedda P, Loddo A, Melis GB. Low androgenization index in elderly women and elderly men with Alzheimer's disease. Neurology. 2004;62:301–303.

250. Papasozomenos S, Shanavas A. Testosterone prevents the heat shock-induced overactivation of glycogen synthase kinase-3 beta but not of cyclin-dependent kinase 5 and c-Jun NH2-terminal kinase and concomitantly abolishes hyperphosphorylation of tau: implications for Alzheimer's disease. Proc Natl Acad Sci U S A. 2002;99:1140–1145.

251. Papasozomenos SC. The heat shock-induced hyperphosphorylation of tau is estrogen-independent and prevented by androgens: implications for Alzheimer disease. Proc Natl Acad Sci U S A. 1997;94:6612–6617.

252. Papasozomenos SC, Papasozomenos T. Androgens Prevent the Heat Shock-Induced Hyperphosphorylation but not Dephosphorylation of t in Female Rats. Implications for Alzheimer's Disease. J Alzheimers Dis. 1999;1:147–153.

253. Park SY, Tournell C, Sinjoanu RC, Ferreira A. Caspase-3- and calpain-mediated tau cleavage are differentially prevented by estrogen and testosterone in beta-amyloid-treated hippocampal neurons. Neuroscience. 2007;144:119–127.

254. Patrone C, Andersson S, Korhonen L, Lindholm D. Estrogen receptor-dependent regulation of sensory neuron survival in developing dorsal root ganglion. Proc Natl Acad Sci U S A. 1999;96:10905–10910.

255. Pennanen C, Kivipelto M, Tuomainen S, Hartikainen P, Hanninen T, Laakso MP, Hallikainen M, Vanhanen M, Nissinen A, Helkala EL, Vainio P, Vanninen R, Partanen K, Soininen H. Hippocampus and entorhinal cortex in mild cognitive impairment and early AD. Neurobiol Aging. 2004;25:303–310.

256. Perillo B, Sasso A, Abbondanza C, Palumbo G. 17beta-estradiol inhibits apoptosis in MCF-7 cells, inducing bcl-2 expression via two estrogen-responsive elements present in the coding sequence. Mol Cell Biol. 2000;20:2890–2901.

257. Persson I, Adami HO, Bergkvist L, Lindgren A, Pettersson B, Hoover R, Schairer C. Risk of endometrial cancer after treatment with oestrogens alone or in conjunction with progestogens: results of a prospective study. Bmj. 1989;298:147–151.

258. Petanceska SS, Nagy V, Frail D, Gandy S. Ovariectomy and 17beta-estradiol modulate the levels of Alzheimer's amyloid beta peptides in brain. Exp Gerontol. 2000;35:1317–1325.

259. Pike CJ. Estrogen modulates neuronal Bcl-xL expression and beta-amyloid-induced apoptosis: relevance to Alzheimer's disease. J Neurochem. 1999;72:1552–1563.

260. Pike CJ. Testosterone attenuates beta-amyloid toxicity in cultured hippocampal neurons. Brain Research. 2001;919:160–165.

261. Pike CJ, Burdick D, Walencewicz AJ, Glabe CG, Cotman CW. Neurodegeneration induced by beta-amyloid peptides in vitro: the role of peptide assembly state. J Neurosci. 1993;13:1676–1687.

262. Pike CJ, Cummings BJ, Cotman CW. beta-Amyloid induces neuritic dystrophy in vitro: similarities with Alzheimer pathology. Neuroreport. 1992;3:769–772.

263. Pike CJ, Cummings BJ, Monzavi R, Cotman CW. Beta-amyloid-induced changes in cultured astrocytes parallel reactive astrocytosis associated with senile plaques in Alzheimer's disease. Neurosci. 1994;63:517–531.

264. Pike CJ, Walencewicz AJ, Glabe CG, Cotman CW. Aggregation-related toxicity of synthetic beta-amyloid protein in hippocampal cultures. Eur J Pharmacol. 1991;207:367–368.

265. Pinkerton JV, Henderson VW. Estrogen and cognition, with a focus on Alzheimer's disease. Semin Reprod Med. 2005;23:172–179.

266. Polo-Kantola P, Portin R, Polo O, Helenius H, Irjala K, Erkkola R. The effect of short-term estrogen replacement therapy on cognition: a randomized, double-blind, cross-over trial in postmenopausal women. Obstet Gynecol. 1998;91:459–466.

267. Pow DV, Perry VH, Morris JF, Gordon S. Microglia in the neurohypophysis associate with and endocytose terminal portions of neurosecretory neurons. Neuroscience. 1989;33:567–578.

268. Qiu J, Bosch MA, Tobias SC, Grandy DK, Scanlan TS, Ronnekleiv OK, Kelly MJ. Rapid Signaling of Estrogen in Hypothalamic Neurons Involves a Novel G-Protein-Coupled Estrogen Receptor that Activates Protein Kinase C. J. Neurosci. 2003;23:9529–9540.

269. Raber J. Androgens, apoE, and Alzheimer's disease. 2004:re2.

270. Raber J. AR, apoE, and cognitive function. Horm Behav. 2008;53:706–715.

271. Ramsden M, Nyborg AC, Murphy MP, Chang L, Stanczyk FZ, Golde TE, Pike CJ. Androgens modulate beta-amyloid levels in male rat brain. J Neurochem. 2003;87:1052–1055.

272. Ramsden M, Shin TM, Pike CJ. Androgens modulate neuronal vulnerability to kainite lesion. Neuroscience. 2003;122:573–578.

273. Rapp SR, Espeland MA, Shumaker SA, Henderson VW, Brunner RL, Manson JE, Gass ML, Stefanick ML, Lane DS, Hays J, Johnson KC, Coker LH, Dailey M, Bowen D. Effect of estrogen plus progestin on global cognitive function in postmenopausal women: the Women's Health Initiative Memory Study: a randomized controlled trial. Jama : the Journal of the American Medical Association. 2003;289:2663–2672.

274. Rasmuson S, Nasman B, Carlstrom K, Olsson T. Increased levels of adrenocortical and gonadal hormones in mild to moderate Alzheimer's disease. Dementia and Geriatric Cognitive Disorders. 2002;13:74–79.

275. Regan RF, Guo Y. Estrogens attenuate neuronal injury due to hemoglobin, chemical hypoxia, and excitatory amino acids in murine cortical cultures. Brain Research. 1997;764:133–140.

276. Resnick SM, Henderson VW. Hormone therapy and risk of Alzheimer disease: a critical time. Jama. 2002;288:2170–2172.

277. Rhodes ME, Frye CA. Androgens in the hippocampus can alter, and be altered by, ictal activity. Pharmacol Biochem Behav. 2004;78:483–493.

278. Rice-Evans C, Miller N, Paganga G. Antioxidant properties of phenolic compounds. Trends in Plant Science. 1997;2:152–159.

279. Rice-Evans CA, Miller NJ, Paganga G. Structure-antioxidant activity relationships of flavonoids and phenolic acids. Free Radical Biology and Medicine. 1996;20:933–956.

280. Riederer P, Hoyer S. From benefit to damage. Glutamate and advanced glycation end products in Alzheimer brain. J Neural Transm. 2006;113:1671–1677.

281. Rocca WA, Amaducci LA, Schoenberg BS. Epidemiology of clinically diagnosed Alzheimer's disease. Ann Neurol. 1986;19:415–424.

282. Rogers J, Strohmeyer R, Kovelowski C, Li RG. Microglia and inflammatory mechanisms in the clearance of amyloid beta peptide. Glia. 2002;40:260–269.

283. Rosario ER, Carroll JC, Oddo S, LaFerla FM, Pike JC. Androgens regulate the development of neuropathology in a triple transgenic mouse model of Alzheimer's disease. J Neurosci. 2006;26:13384–13389.

284. Rosario ER, Chang L, Head EH, Stanczyk FZ, Pike CJ. Brain levels of sex steroid hormones in men and women during normal aging and in Alzheimer's disease. Neurobiol Aging. 2009 In Press.

285. Rosario ER, Chang L, Stanczyk FZ, Pike CJ. Age-related testosterone depletion and the development of Alzheimer disease. Jama. 2004;292:1431–1432.

286. Rosario ER, Pike CJ. Androgen regulation of beta-amyloid protein and the risk of Alzheimer's disease. Brain Res Rev. 2008;57:444–453.

287. Rosario ER, Ramsden M, Pike CJ. Progestins inhibit the neuroprotective effects of estrogen in rat hippocampus. Brain Res. 2006;1099:206–210.

288. Rowan MJ, Klyubin I, Wang Q, Hu NW, Anwyl R. Synaptic memory mechanisms: Alzheimer's disease amyloid beta-peptide-induced dysfunction. Biochem Soc Trans. 2007;35:1219–1223.

289. Rubin BS, Fox TO, Bridges RS. Estrogen binding in nuclear and cytosolic extracts from brain and pituitary of middle-aged female rats. Brain Res. 1986;383:60–67.

290. Ruitenberg A, Ott A, van Swieten JC, Hofman A, Breteler MMB. Incidence of dementia: does gender make a difference? Neurobiology of Aging. 2001;22:575–580.

291. Sastre M, Walter J, Gentleman SM. Interactions between APP secretases and inflammatory mediators. J Neuroinflammation. 2008;5:25.

292. Savaskan E, Olivieri G, Meier F, Ravid R, Muller-Spahn F. Hippocampal estrogen beta-receptor immunoreactivity is increased in Alzheimer's disease. Brain Res. 2001;908:113–119.

293. Schultz C, Braak H, Braak E. A sex difference in neurodegeneration of the human hypothalamus. Neurosci Lett. 1996;212:103–106.

294. Schultz C, Ghebremedhin E, Braak E, Braak H. Sex-dependent cytoskeletal changes of the human hypothalamus develop independently of Alzheimer's disease. Exp Neurol. 1999;160:186–193.

295. Schumacher M, Sitruk-Ware R, De Nicola AF. Progesterone and progestins:neuroprotection and myelin repair. Curr Opin Pharmacol. 2008;8:740–746.

296. Schupf N, Lee JH, Wei M, Pang D, Chace C, Cheng R, Zigman WB, Tycko B, Silverman W. Estrogen receptor-alpha variants increase risk of Alzheimer's disease in women with Down syndrome. Dement Geriatr Cogn Disord. 2008;25:476–482.

297. Selkoe DJ, Yamazaki T, Citron M, Podlisny MB, Koo EH, Teplow DB, Haass C. The role of APP processing and trafficking pathways in the formation of amyloid beta-protein. Ann NY Acad Sci. 1996;777:57–64.

298. Selvamani A, Sohrabji F. Reproductive age modulates the impact of focal ischemia on the forebrain as well as the effects of estrogen treatment in female rats. Neurobiol Aging. 2008

299. Sétáló G, Singh M, Nethrapalli IS, Toran-Allerand CD. Protein Kinase C Activity is Necessary for Estrogen-Induced Erk Phosphorylation in Neocortical Explants. Journal Neurochemical Research. 2005;30:779–790.

300. Shang Y, Brown M. Molecular determinants for the tissue specificity of SERMs. Science. 2002;295:2465–2468.

301. Shao TC, Li HL, Kasper S, Matusik R, Ittmann M, Cunningham GR. Comparison of the growth-promoting effects of testosterone and 7-alpha-methyl-19-nor-testosterone MENT) on the prostate and levator ani muscle of LPB-tag transgenic mice. Prostate. 2006;66:369–376.

302. Sheffield-Moore M, Urban RJ. An overview of the endocrinology of skeletal muscle. Trends Endocrinol Metab. 2004;15:110–115.

303. Sheldahl LC, Marriott LK, Bryant DM, Shapiro RA, Dorsa DM. Neuroprotective effects of estrogen and selective estrogen receptor modulators begin at the plasma membrane. Minerva Endocrinol. 2007;32:87–94.

304. Shen R, Sumitomo M, Dai J, Hardy DO, Navarro D, Usmani B, Papandreou CN, Hersh LB, Shipp MA, Freedman LP, Nanus DM. Identification and characterization of two androgen response regions in the human neutral endopeptidase gene. Mol Cell Endocrinol. 2000;170:131–142.

305. Sherwin BB. Estrogen and Cognitive Functioning in Women. Endocrine Reviews. 2003;24:133–151.

306. Sherwin BB. Estrogen and/or androgen replacement therapy and cognitive functioning in surgically menopausal women. Psychoneuroendocrinology. 1988;13:345–357.

307. Shi J, Panickar KS, Yang S-H, Rabbani O, Day AL, Simpkins JW. Estrogen attenuates over-expression of b-amyloid precursor protein messager RNA in an animal model of focal ischemia. Brain Research. 1998;810:87–92.

308. Shilling V, Jenkins V, Fallowfield L, Howell T. The effects of hormone therapy on cognition in breast cancer. J Steroid Biochem Mol Biol. 2003;86:405–412.

309. Shughrue PJ, Lane MV, Merchenthaler I. Comparative distribution of estrogen receptor-alpha and -beta mRNA in the rat central nervous system. J Comp Neurol. 1997;388:507–525.

310. Shughrue PJ, Merchenthaler I. Estrogen is More Than just a “Sex Hormone”: Novel Sites for Estrogen Action in the Hippocampus and Cerebral Cortex. Frontiers in Neuroendocrinology. 2000;21:95–101.

311. Shumaker SA, Legault C, Kuller L, Rapp SR, Thal L, Lane DS, Fillit H, Stefanick ML, Hendrix SL, Lewis CE, Masaki K, Coker LH. Conjugated equine estrogens and incidence of probable dementia and mild cognitive impairment in postmenopausal women: Women's Health Initiative Memory Study. Jama. 2004;291:2947–2958.

312. Shumaker SA, Legault C, Rapp SR, Thal L, Wallace RB, Ockene JK, Hendrix SL, Jones BN, Assaf AR, Jackson RD, Kotchen JM, Wassertheil_Smoller S, Wactawski_Wende J. Estrogen plus progestin and the incidence of dementia and mild cognitive impairment in postmenopausal women: the Women's Health Initiative Memory Study: a randomized controlled trial. Jama : the Journal of the American Medical Association. 2003;289:2651–2662.

313. Silva I, Mello LE, Freymuller E, Haidar MA, Baracat EC. Estrogen, progestogen and tamoxifen increase synaptic density of the hippocampus of ovariectomized rats. Neurosci Lett. 2000;291:183–186.

314. Simerly RB, Chang C, Muramatsu M, Swanson LW. Distribution of androgen and estrogen receptor mRNA-containing cells in the rat brain: an in situ hybridization study. The Journal of Comparative Neurology. 1990;294:76–95.

315. Singer CA, Figueroa_Masot XA, Batchelor RH, Dorsa DM. The mitogen-activated protein kinase pathway mediates estrogen neuroprotection after glutamate toxicity in primary cortical neurons. The Journal of Neuroscience : the Official Journal of the Society For Neuroscience. 1999;19:2455–2463.

316. Singer CA, Rogers KL, Dorsa DM. Modulation of Bcl-2 expression: a potential component of estrogen protection in NT2 neurons. Neuroreport. 1998;9:2565–2568.

317. Singer CA, Rogers KL, Strickland TM, Dorsa DM. Estrogen protects primary cortical neurons from glutamate toxicity. Neuroscience Letters. 1996;212:13–16.

318. Singh M. Ovarian hormones elicit phosphorylation of Akt and extracellular-signal regulated kinase in explants of the cerebral cortex. Endocrine. 2001;14:407–415.

319. Singh M, Setalo G, Jr, Guan X, Warren M, Toran-Allerand CD. Estrogen-Induced Activation of Mitogen-Activated Protein Kinase in Cerebral Cortical Explants:Convergence of Estrogen and Neurotrophin Signaling Pathways. J. Neurosci. 1999;19:1179–1188.

320. Stoltzner SE, Berchtold NC, Cotman CW, Pike CJ. Estrogen regulates bcl-x expression in rat hippocampus. Neuroreport. 2001;12:2797–2800.

321. Stone DJ, Rozovsky I, Morgan TE, Anderson CP, Lopez LM, Shick J, Finch CE. Effects of age on gene expression during estrogen-induced synaptic sprouting in the female rat. Exp Neurol. 2000;165:46–57.

322. Strittmatter WJ, Roses AD. Apolipoprotein E and Alzheimer disease. Proc Natl Acad Sci U S A. 1995;92:4725–4727.

323. Struble RG, Nathan BP, Cady C, Cheng X, McAsey M. Estradiol regulation of astroglia and apolipoprotein E: An important role in neuronal regeneration. Experimental Gerontology. 2007;42:54–63.

324. Subbiah MT, Kessel B, Agrawal M, Rajan R, Abplanalp W, Rymaszewski Z. Antioxidant potential of specific estrogens on lipid peroxidation. J Clin Endocrinol Metab. 1993;77:1095–1097.

325. Sugioka K, Shimosegawa Y, Nakano M. Estrogens as natural antioxidants of membrane phospholipid peroxidation. FEBS Letters. 1987;210:37–39.

326. Swerdloff RS, Wang C. Androgen deficiency and aging in men. The Western Journal of Medicine. 1993;159:579–585.

327. Tan RS, Culberson JW. An integrative review on current evidence of testosterone replacement therapy for the andropause. Maturitas. 2003;45:15–27.

328. Tan RS, Pu SJ. A pilot study on the effects of testosterone in hypogonadal aging male patients with Alzheimer's disease. Aging Male. 2003;6:13.

329. Tang MX, Jacobs D, Stern Y, Marder K, Schofield PR, Gurland B, Andrews H, Mayeux R. Effect of oestrogen during menopause on risk and age at onset of Alzheimer's disease. The Lancet. 1996;348:429.

330. Tanzer L, Jones KJ. Gonadal steroid regulation of hamster facial nerve regeneration: effects of dihydrotestosterone and estradiol. Exp Neurol. 1997;146:258–264.

331. Tanzi RE, Moir RD, Wagner SL. Clearance of Alzheimer's A[beta] Peptide: The many Roads to Perdition. Neuron. 2004;43:605–608.

332. Tariq SH, Haren MT, Kim MJ, Morley JE. Andropause: is the emperor wearing any clothes? Rev Endocr Metab Disord. 2005;6:77–84.

333. Thakur MK, Mani ST. Estradiol regulates APP mRNA alternative splicing in the mice brain cortex. Neuroscience Letters. 2005;381:154–157.

334. Thomas T, Bryant M, Clark L, Garces A, Rhodin J. Estrogen and Raloxifene Activities on Amyloid--Induced Inflammatory Reaction. Microvascular Research. 2001;61:28–39.

335. Toran-Allerand CD, Singh M, Sétáló G. Novel Mechanisms of Estrogen Action in the Brain: New Players in an Old Story. Frontiers in Neuroendocrinology. 1999;20:97–121.

336. Turner PR, O'Connor K, Tate WP, Abraham WC. Roles of amyloid precursor protein and its fragments in regulating neural activity, plasticity and memory. Prog Neurobiol. 2003;70:1–32.

337. Vardy ERLC, Catto AJ, Hooper NM. Proteolytic mechanisms in amyloid-[beta] metabolism: therapeutic implications for Alzheimer's disease. Trends in Molecular Medicine. 2005;11:464–472.

338. Vasto S, Candore G, Listi F, Balistreri CR, Colonna-Romano G, Malavolta M, Lio D, Nuzzo D, Mocchegiani E, Di Bona D, Caruso C. Inflammation, genes and zinc in Alzheimer's disease. Brain Res Rev. 2008;58:96–105.

339. Vaughan C, Goldstein FC, Tenover JL. Exogenous testosterone alone or with finasteride does not improve measurements of cognition in healthy older men with low serum testosterone. J Androl. 2007;28:875–882.

340. Vedder H, Anthes N, Stumm G, Wurz C, Behl C, Krieg J-C. Estrogen Hormones Reduce Lipid Peroxidation in Cells and Tissues of the Central Nervous System. Journal of Neurochemistry. 1999;72:2531–2538.

341. Vegeto E, Bonincontro C, Pollio G, Sala A, Viappiani S, Nardi F, Brusadelli A, Viviani B, Ciana P, Maggi A. Estrogen Prevents the Lipopolysaccharide-Induced Inflammatory Response in Microglia. Journal of Neuroscience. 2001;21:1809–1818.

342. Venken K, Boonen S, Van Herck E, Vandenput L, Kumar N, Sitruk-Ware R, Sundaram K, Bouillon R, Vanderschueren D. Bone and muscle protective potential of the prostate-sparing synthetic androgen 7alpha-methyl-19-nortestosterone: evidence from the aged orchidectomized male rat model. Bone. 2005;36:663–670.

343. Verdile G, Fuller S, Atwood CS, Laws SM, Gandy SE, Martins RN. The role of beta amyloid in Alzheimer's disease: still a cause of everything or the only one who got caught? Pharmacological Research. 2004;50:397–409.

344. Vincent B, Smith JD. Effect of estradiol on neuronal Swedish-mutated beta-amyloid precursor protein metabolism: reversal by astrocytic cells. Biochemical and Biophysical Research Communications. 2000;271:82–85.

345. Eckardstein Svon, Noe G, Brache V, Nieschlag E, Croxatto H, Alvarez F, Moo-Young A, Sivin I, Kumar N, Small M, Sundaram K. A clinical trial of 7 alpha-methyl-19-nortestosterone implants for possible use as a long-acting contraceptive for men. J Clin Endocrinol Metab. 2003;88:5232–5239.

346. Wahjoepramono EJ, Wijaya LK, Taddei K, Martins G, Howard M, Ruyck Kd, Bates K, Dhaliwald SS, Verdile G, Carruthers M, Martins RN. Distinct Effects of Testosterone on Plasma and Cerebrospinal Fluid Amyloid-b Levels. Journal of Alzheimer's Disease. 2008;129:129–137.

347. Walsh D, Selkoe D. A beta oligomers - a decade of discovery. Journal of Neurochemistry. 2007;101:1172–1184.

348. Walsh DM, Klyubin I, Fadeeva JV, Rowan MJ, Selkoe DJ. Amyloid-beta oligomers: their production, toxicity and therapeutic inhibition. Biochem Soc Trans. 2002;30:552–557.

349. Wang J, Tanila H, Puolivali J, Kadish I, van Groen T. Gender differences in the amount and deposition of amyloidbeta in APPswe and PS1 double transgenic mice. Neurobiol Dis. 2003;14:318–327.

350. Wang PN, Liao SQ, Liu RS, Liu CY, Chao HT, Lu SR, Yu HY, Wang SJ, Liu HC. Effects of estrogen on cognition, mood, and cerebral blood flow in AD: A controlled study. Neurology. 2000;54:2061–2066.

351. Waring SC, Rocca WA, Petersen RC, O_Brien PC, Tangalos EG, Kokmen E. Postmenopausal estrogen replacement therapy and risk of AD: a population-based study. Neurology. 1999;52:965–970.

352. Watanabe T, Koba S, Kawamura M, Itokawa M, Idei T, Nakagawa Y, Iguchi T, Katagiri T. Small dense low-density lipoprotein and carotid atherosclerosis in relation to vascular dementia. Metabolism. 2004;53:476–482.

353. Watters JJ, Campbell JS, Cunningham MJ, Krebs EG, Dorsa DM. Rapid Membrane Effects of Steroids in Neuroblastoma Cells: Effects of Estrogen on Mitogen Activated Protein Kinase Signalling Cascade and c-fos Immediate Early Gene Transcription. Endocrinology. 1997;138:4030–4033.

354. Weaver CE, Jr, Park-Chung M, Gibbs TT, Farb DH. 17beta-Estradiol protects against NMDA-induced excitotoxicity by direct inhibition of NMDA receptors. Brain Res. 1997;761:338–341.

355. Weaver CL, Espinoza M, Kress Y, Davies P. Conformational change as one of the earliest alterations of tau in Alzheimer's disease. Neurobiol Aging. 2000;21:719–727.

356. Weiss JH, Pike CJ, Cotman CW. Ca2+ channel blockers attenuate beta-amyloid peptide toxicity to cortical neurons in culture. J Neurochem. 1994;62:372–375.

357. Wharton GC, Brinton W, Cedars E, Lobo M, Manson RA, Merriam J, Miller GR, Santoro VM, Taylor N, Harman HS, Asthana S SM. Baseline cognitive and demographic characteristics of women enrolled in the Kronos Early Estrogen Prevention Study (KEEPS) Alzheimer's ' Dementia: The Journal of the Alzheimer's Association. 2008;4:T792.

358. Wilhelmsen KC, Clark LN, Miller BL, Geschwind DH. Tau mutations in frontotemporal dementia. Dement Geriatr Cogn Disord. 1999;10 Suppl 1:88–92.

359. Wilson EM. Muscle-bound? A tissue-selective nonsteroidal androgen receptor modulator. Endocrinology. 2007;148:1–3.

360. Wise PM. Estrogen therapy: does it help or hurt the adult and aging brain? Insights derived from animal models. Neuroscience. 2006;138:831–835.

361. Wise PM, Parsons B. Nuclear estradiol and cytosol progestin receptor concentrations in the brain and the pituitary gland and sexual behavior in ovariectomized estradiol-treated middle-aged rats. Endocrinology. 1984;115:810–816.

362. Woolley CS, McEwen Bruce S. Roles of estradiol and progesterone in regulation of hippocampal dendritic spine density during the estrous cycle in the rat. The Journal of Comparative Neurology. 1993;336:293–306.

363. Wu TW, Wang JM, Chen S, Brinton RD. 17Beta-estradiol induced Ca2+ influx via L-type calcium channels activates the Src/ERK/cyclic-AMP response element binding protein signal pathway and BCL-2 expression in rat hippocampal neurons: a potential initiation mechanism for estrogen-induced neuroprotection. Neuroscience. 2005;135:59–72.

364. Wu X, Glinn MA, Ostrowski NL, Su Y, Ni B, Cole HW, Bryant HU, Paul SM. Raloxifene and estradiol benzoate both fully restore hippocampal choline acetyltransferase activity in ovariectomized rats. Brain Res. 1999;847:98–104.

365. Xiao Z-M, Sun L, Liu Y-M, Zhang J-J, Huang J. Estrogen Regulation of the Neprilysin Gene Through A Hormone-Responsive Element. Journal Journal of Molecular Neuroscience. 2009 In press.

366. Xu H, Gouras GK, Greenfield JP, Vincent B, Naslund J, Mazzarelli L, Fried G, Jovanovic JN, Seeger M, Relkin NR, Liao F, Checler F, Buxbaum JD, Chait BT, Thinakaran G, Sisodia SS, Wang R, Greengard P, Gandy S. Estrogen reduces neuronal generation of Alzheimer beta-amyloid peptides. Nat Med. 1998;4:447–451.

367. Xu H, Wang R, Zhang Y-W, Zhang X. Estrogen, -Amyloid Metabolism/Trafficking, and Alzheimer's Disease. Annals of the New York Academy of Sciences. 2006;1089:324–342.

368. Xu Y, Armstrong SJ, Arenas IA, Pehowich DJ, Davidge ST. Cardioprotection by chronic estrogen or superoxide dismutase mimetic treatment in the aged female rat. Am J Physiol Heart Circ Physiol. 2004;287:H165–H171.

369. Yaffe K, Krueger K, Cummings SR, Blackwell T, Henderson VW, Sarkar S, Ensrud K, Grady D. Effect of raloxifene on prevention of dementia and cognitive impairment in older women: the Multiple Outcomes of Raloxifene Evaluation (MORE) randomized trial. Am J Psychiatry. 2005;162:683–690.

370. Yaffe K, Krueger K, Sarkar S, Grady D, Barrett-Connor E, Cox DA, Nickelsen T. The Multiple Outcomes of Raloxifene Evaluation Investigators, Cognitive Function in Postmenopausal Women Treated with Raloxifene. New England Journal of Medicine. 2001;344:1207–1213.

371. Yang SH, Liu R, Wen Y, Perez E, Cutright J, Brun-Zinkernagel AM, Singh M, Day AL, Simpkins JW. Neuroendocrine mechanism for tolerance to cerebral ischemia-reperfusion injury in male rats. J Neurobiol. 2005;62:341–351.

372. Yang SH, Perez E, Cutright J, Liu R, He Z, Day AL, Simpkins JW. Testosterone increases neurotoxicity of glutamate in vitro and ischemia-reperfusion injury in an animal model. J Appl Physiol. 2002;92:195–201.

373. Yankner BA, Lu T. Amyloid {beta}-Protein Toxicity and the Pathogenesis of Alzheimer Disease. J Biol Chem. 2009;284:4755–4759.

374. Yao M, Nguyen TV, Pike CJ. Beta-amyloid-induced neuronal apoptosis involves c-Jun N-terminal kinase-dependent downregulation of Bcl-w. J Neurosci. 2005;25:1149–1158.

375. Yao M, Nguyen TV, Pike CJ. Estrogen regulates Bcl-w and Bim expression: role in protection against beta-amyloid peptide-induced neuronal death. J Neurosci. 2007;27:1422–1433.

376. Yao M, Nguyen TV, Rosario ER, Ramsden M, Pike CJ. Androgens regulate neprilysin expression: role in reducing beta-amyloid levels. J Neurochem. 2008

377. Yu WH. Administration of testosterone attenuates neuronal loss following axotomy in the brain-stem motor nuclei of female rats. J Neurosci. 1989;9:3908–3914.

378. Yu WH. Effect of testosterone on the regeneration of the hypoglossal nerve in rats. Exp Neurol. 1982;77:129–141.

379. Yu WH, Srinivasan R. Effect of testosterone and 5 alpha-dihydrotestosterone on regeneration of the hypoglossal nerve in rats. Exp Neurol. 1981;71:431–435.

380. Yue X, Lu M, Lancaster T, Cao P, Honda S-I, Staufenbiel M, Harada N, Zhong Z, Shen Y, Li R. Brain estrogen deficiency accelerates Ab plaque formation in an Alzheimer's disease animal model. Proceedings of the National Academy of Sciences. 2005;102:19198–19203.

381. Yune TY, Park HG, Lee JY, Oh TH. Estrogen-induced Bcl-2 expression after spinal cord injury is mediated through phosphoinositide-3-kinase/Akt-dependent CREB activation. J Neurotrauma. 2008;25:1121–1131.

382. Zandi PP, Carlson MC, Plassman BL, Welsh-Bohmer KA, Mayer LS, Steffens DC, Breitner JCS. Hormone Replacement Therapy and Incidence of Alzheimer Disease in Older Women: The Cache County Study. Journal of the American Medical Association. 2002;288:2123–2129.

383. Zhang L, Rubinow DRCA, Xaing G-q, Li B-S, Chang YH, Maric D, Barker JL, Ma W. Estrogen protects against b-amyloid-induced neurotoxicity in rat hippocampal neurons by activation of Akt. Neuroreport. 2001;12:1919–1923.

384. Zhang QG, Wang R, Khan M, Mahesh V, Brann DW. Role of Dickkopf-1, an antagonist of the Wnt/beta-pathway, in estrogen-induced neuroprotection and attenuation of tau phosphorylation. J Neurosci. 2008;28:8430–8441.

385. Zhang S, Huang Y, Zhu YC, Yao T. Estrogen stimulates release of secreted amyloid precursor protein from primary rat cortical neurons via protein kinase C pathway. Acta Pharmacol Sin. 2005;26:171–176.

386. Zhang Y, Champagne N, Beitel LK, Goodyer CG, Trifiro M, LeBlanc A. Estrogen and androgen protection of human neurons against intracellular amyloid beta(1–42) toxicity through heat shock protein 70. Journal of Neuroscience. 2004;24:5315–5321.

387. Zhao L, Brinton RD. Estrogen receptor alpha and beta differentially regulate intracellular Ca(2+) dynamics leading to ERK phosphorylation and estrogen neuroprotection in hippocampal neurons. Brain Res. 2007;1172:48–59.

388. Zhao L, Jin C, Mao Z, Gopinathan MB, Rehder K, Brinton RD. Design, synthesis, and estrogenic activity of a novel estrogen receptor modulator--a hybrid structure of 17beta-estradiol and vitamin E in hippocampal neurons. J Med Chem. 2007;50:4471–4481.

389. Zhao L, Wu TW, Brinton RD. Estrogen receptor subtypes alpha and beta contribute to neuroprotection and increased Bcl-2 expression in primary hippocampal neurons. Brain Res. 2004;1010:22–34.

390. Zheng H, Xu H, Uljon SN, Gross R, Hardy K, Gaynor J, Lafrancois J, Simpkins J, Refolo LM, Petanceska S, Wang R, Duff K. Modulation of A(beta) peptides by estrogen in mouse models. J Neurochem. 2002;80:191–196.