Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Nov 24.
Published in final edited form as: Front Biosci. 2008 Jan 1;13:1083–1089. doi: 10.2741/2746

ESTROGENS AND PROGESTERONE AS NEUROPROTECTANTS: WHAT ANIMAL MODELS TEACH US

Meharvan Singh 1, Nathalie Sumien 1, Cheryl Kyser 1, James W Simpkins 1
PMCID: PMC2586167  NIHMSID: NIHMS78915  PMID: 17981614

Abstract

Estradiol and progesterone are two steroid hormones that target a variety of organ systems, including the heart, the bone and the brain. With respect to the latter, a large volume of basic science studies support the neuroprotective role of estradiol and/or progesterone. In fact, the results of such studies prompted the assessment of these hormones as protective agents against such disorders as Alzheimer’s disease, stroke and traumatic brain injury. Interestingly, results from the Women’s Health Initiative (WHI) yielded results that appeared to be inconsistent with the data derived from in vitro and in vivo models. However, we argue that the results from the basic science studies were not inconsistent with the clinical trials, but rather, are consistent with, and may even have predicted, the results from the WHI. To illustrate this point, we review here certain in vivo paradigms that have been used to assess the protective effects of estrogens and progesterone, and describe how the results from these animal models point to the importance of the type of hormone, the age of the subjects and the method of hormone administration, in determining whether or not hormones are neuroprotective.

Keywords: Estrogens, Progestins, Neuroprotection, Alzheimer’s disease, Animal Models, Review

2. INTRODUCTION

The U.S. Census Bureau estimates that by 2010, the population of women between 45 and 64 years old will reach approximately 42 million, a marked increase from the value reported for 2000 (approx. 32 million) (U.S. Census Bureau. Projected population of the United States, by Age and Sex: 200 to 2050. www.census.gov/ipc/www/usinterimproj/Internet release date: March 18, 2004). This increasing number of women will consequently need to make decisions about the use of hormone therapy to treat not only menopausal symptoms, but potentially, to maintain a healthy brain. And though numerous basic science studies, epidemiological studies and some clinical trials have supported the potential benefit of hormone therapy in reducing the incidence of age-associated brain dysfunction (including reducing the risk for Alzheimer’s disease), recent results from the Women’s Health Initiative (WHI) have suggested the contrary and left the field unsettled as to the future of hormone therapy. For example, caveats of the WHI (particularly the WHI memory study, WHIMS) include the possibility that both the age of the subjects and the duration of post-menopausal hormone deprivation diminish the protective brain response to steroid hormones. Additionally, the type of hormone (for example, progesterone versus the synthetic, medroxyprogesterone acetate) and its method of administration may have also influenced the outcomes.

While it is clear that a better understanding of the neurobiology of gonadal steroid hormones and their receptors is needed, it is equally important that we interpret the basic science studies appropriately, and understand their limitations if we are to apply the results from these studies towards the design of the next large-scale clinical trial or in fact, implement safer and more effective ways of treating the menopause and the post-menopausal period. To this end, we review here some of the common animal models that have been used to assess the neuroprotective efficacy of estrogens and progestins, and describe some of the limitations that must be acknowledged in interpreting their results.

3. MODELS USED TO ASSESS THE NEUROPROTECTIVE EFFECTS OF ESTROGENS AND PROGESTINS

3.1. The ovariectomized rodent

In vivo studies that have tested the effects of estrogens and/or progestins on various aspects of neurobiology have often employed the ovariectomized rodent (using either rats or mice) as an experimental model. Indeed, numerous studies have demonstrated that ovariectomy can result in impairment of cognitive performance (13), influence structural plasticity of neurons (4), impair cholinergic function (2, 5), and reduce neurotrophin expression (3, 68). All these changes have been linked to such neurodegenerative diseases such as Alzheimer’s disease, and as such, the ovariectomized rodent has been used to assess if steroid hormones such as estradiol and progesterone may prevent or attenuate some of these deficits. In fact, both estradiol and progesterone have been shown to be effective neuroprotectants in a variety of animal models in which ovariectomy was used to eliminate ovarian steroid hormone production. Examples are provided below:

3.2. The stroke model

The neuroprotective effects of estrogens have been demonstrated in a variety of models of stroke, as induced by causing acute cerebral ischemia. These include transient and permanent middle cerebral artery occlusion models (911), global forebrain ischemia models (12, 13), photothrombotic focal ischemia models (14), and glutamate-induced focal cerebral ischemia models (15). The neuroprotective effects of estrogens have also been demonstrated against subarachnoid hemorrhage, a highly prevalent form of stroke in females (16). These protective effects have been described in multiple species, including rats, mice and gerbils (17, 18). Collectively, these results support the argument that estrogens could be valuable candidates for brain protection in females.

It is important to recognize that ovariectomy causes a dramatic reduction in not only circulating estradiol, but also in circulating progesterone. As such, the structural and functional impairments that are reported to occur following ovariectomy may result from the loss in not only circulating estrogens, but of progesterone as well. Accordingly, it should come as no surprise, that progesterone is also an effective neuroprotectant in animal models of stroke. For example, Jiang et al. illustrated that the administration of progesterone before middle cerebral artery occlusion (MCAO) resulted in a marked reduction in cerebral infarction and reduced impairments that resulted from the occlusion (19). Interestingly, progesterone was also found to be protective even when administered shortly after the ischemic event (20, 21), and resulted in improvements in various functional measures, including the rotarod test, and adhesive-backed somatosensory and neurological scores (22). Further, progesterone also protected against cell death following an acute episode of global ischemia (23), and may be mediated by progesterone’s ability to reduce lipid peroxidation, the generation of isoprostanes (24) and the expression of pro-inflammatory genes (25).

3.3. Traumatic brain injury

Another in vivo paradigm that has taken advantage of the ovariectomized model to assess the effectiveness of steroid hormones as a neuroprotectant is the traumatic brain injury (TBI) model. In particular, studies assessing the effects of progesterone have found that it can significantly reduce cerebral edema, even when administered up to 24 hours after the experimental injury. Mechanistically, the protective effects of progesterone may be attributed to its ability to reduce complement factor C3, glial fibrillary acidic protein (GFAP), and nuclear factor kappa beta (NFκB) in the TBI model (25), and decrease the levels of lipid peroxidation (26).

And in animal models that simulate demyelinating disease, progesterone has also been found to have significant neuroprotective potential, as evidenced by its ability to increase the expression of myelin proteins in the damaged sciatic nerves of young adult rats with nerve crush injuries (27). Furthermore, progesterone has also been shown to promote morphological and functional recovery in the Wobbler mouse, an animal model of spinal cord degeneration (28, 29).

These pieces of evidence that were obtained from animal models certainly supported the neuroprotective potential of these steroid hormones, and offered some insight into the mechanisms by which these hormones may be exerting their protective effects.

However, when this neuroprotective potential of estrogens and progesterone was applied to certain clinical trials, such as the WHIMS, the results were far from expected and were in fact, deemed inconsistent with the data derived from various animal studies. The estrogen formulation, Premarin (30, 31) or the combined estrogen/progestin formulation, PremPro, increased the risk for dementia and stroke (3234) unlike what the animal studies would have predicted. And though the temptation has been to place blame on either the basic science data or alternatively, the clinical studies, we argue that the data resulting from animal models, were on the contrary, consistent with the WHI studies and arguably, predicted the results of the WHI.

4. FACTORS THAT INFLUENCE HORMONE INDUCED NEUROPROTECTION

4.1. The importance of age

The ovariectomized model is an excellent model to study the effect of estrogen and/or progesterone without the potential confound of having endogenous levels of estrogens and progesterone being contributed by the ovary. However, the ovariectomized animal best mimics the surgically menopausal woman, and may not necessarily model the menopause. Thus, the animal studies in which estradiol and/or progesterone administration to a young adult, ovariectomized animal has yielded beneficial effects, such as neuroprotection, should be interpreted as predicting the beneficial effects of a specific form of estrogen (17 β-estradiol), and/or progesterone in younger, premenopausal women (perhaps those who have undergone bilateral oophorectomy). The WHIMS, however, assessed the effects of a particular estrogen formulation on older women who were on average 10 years past the menopause.

Animal studies have, in fact, demonstrated that (reproductive) age influences the protective effects of estradiol. For example, estradiol was found to be effective at protecting against scopolamine-induced cognitive impairment in rats that were beginning to go through the phase of reproductive senescence (as evidenced by a phase of constant estrus). However, in older animals (that were in a state of constant diestrus), estradiol treatment was ineffective (35). Moreover, Nordell and colleagues (36) have demonstrated that in young adults, estradiol administration to an ovariectomized rat attenuates the pro-inflammatory effects of an excitotoxic lesion, whereas estradiol exacerbated the toxicity in older, reproductive senescent animals. Thus, data from animal studies have suggested that neuroprotective efforts with estrogen in older animals is not effective and may actually exacerbate pathological cascades, and as such, are perfectly consistent with the WHI trial in which estrogen treatment of older women was found to lead to an increased risk for dementia and stroke. Furthermore, new reports suggest that, when the WHI data were segregated according to age, younger women were more likely to benefit from hormone therapy, whereas older women who were approximately 10 years beyond the menopause were prone to the negative consequences of hormones therapy (37).

4.2. The type of hormone matters

Another important distinction between those animal studies in which estrogens and/or progesterone has been found to be protective and the clinical trials in which an estrogen or an estrogen and a progestin preparation had negative consequences lies in the choice of hormone used. Basic science studies have, in fact, demonstrated that progesterone is neuroprotective (19, 26, 28, 29, 3840), while the synthetic progestin, MPA, is not (38, 39, 41). For example, in a model of stroke (reversible focal stroke using the intraluminal filament model followed by 22 hours of reperfusion), MPA diminished the protective effects of conjugated equine estrogens (CEE) and MPA diminished estrogen’s ability to reduce stroke damage (41). Interestingly, with regards to the traumatic brain injury model, MPA required a larger dose than P4 to accomplish a comparable reduction in cerebral edema. However regardless of the dose of MPA, MPA did not favor a better behavioral recovery than progesterone (reviewed in (42)).

Therefore, while P4 does not interfere with the beneficial effects of estrogens, MPA appears to have the capacity to prevent estrogen’s beneficial effects. Such data requires us to consider the possibility that some of the negative consequences of hormone therapy observed in the WHI trials may have been a result of the choice of progestin used in the hormone therapy regimen.

4.3. The delivery of estrogens and progesterone

In animal models, several routes of administration and delivery of steroid hormones have been used, including intravenous injection, subcutaneous injection, intranasal administration (43), oral gavage (44, 45), addition of the steroid to the water supply of animals (46), subcutaneous implantation of Silastic® pellets and implantation of the “matrix-driven delivery (MDD)” pellet (IRA, Sarasota, Fl). The latter two methods have been used with the intention of providing continuous delivery of steroid hormones. While the use of Silastic® pellets have been demonstrated be effective in delivering a constant level of 17 β-estradiol (17β-E2) over weeks (10, 47) to months (3) in rats, the delivery of constant and sustained levels of hormone (either 17 β-E2 or P4) to mice has proven to be more problematic. To obviate the problem of sustained steroid hormone delivery to mice, several approaches have been tested, including oral gavage, addition to the water supply of the animals (46) and implantation of the IRA MDD pellet system.

The later method of steroid delivery have become widely used, with IRA reporting 924 published papers using their 17 β-E2 pellets and 142 reports using their P4 pellets (www.innovrsch.com). However, we have discovered that these pellets have not been validated for the kinetic of release of the steroid in mice. Instead, most published reports that utilized the 17 β-E2 pellets from IRA cite either the IRA website (www.innovrsch.com) or other publications to support a continuous release of 17 β-E2. Surprisingly, we found that these citations, in turn, lead to a single ex vivo dissolution study of one of the IRA pellets (48). Indeed, most studies assess levels of E2 at the termination of their study and claim continuous release of the steroid at the reported concentration throughout the duration of treatment. Our data, however, in which we assessed the acute (up to 7 days) release characteristics of these steroid pellets suggest that this is not a valid assumption, and instead, show that both 17 β-E2 and P4 pellets produce a huge initial burst of hormone, followed by a gradual decline that even after 7 days is far exceed the target mid-estrous cycle target levels of 50 pg/ml for 17 β-E2 and 4 ng/ml for P4. (Tables 1 and 2).

Table 1.

Concentration of 17 β-E2 at various times after implantation of IRA “matrix-driven delivery (MDD)” pellets containing 17 β-E2

Hours after implantation Serum 17 β-E2 concentration (pg/ml) Pellet 17 β-E2 Content (mg) Proposed duration of 17 β-E2 release
1 13,044 0.25 60
4 10,563 0.25 60
8 2,536 0.25 60
12 1,113 0.25 60
24 965 0.25 60
Days after implantation
4 972 0.25 60
7 856 0.25 60

17 β-E2: 17 beta estradiol

Table 2.

Measured serum progesterone levels at various times following implantation of the IRA “matrix-driven delivery (MDD)” progesterone pellets

Hours after implantation Serum Progesterone concentration (ng/ml) Pellet Progesterone Content (mg) Proposed duration of Progesterone release
1 133.2 10 60
4 165.2 10 60
8 191.9 10 60
12 97.9 10 60
24 79.6 10 60
Days after implantation
4 37.2 10 60
7 38.6 10 60

The results reported here could also explain the results recently reported by Green et al. (49), who used a similar treatment with a 0.25 mg, 90-day release 17 β-E2 pellets from IRA. They observed that at the end of 90 days, plasma estradiol was elevated 3-fold over ovary-intact controls, uterine weights were 81% higher than ovary-intact controls and pituitary weights were 2.6-fold greater than ovary-intact controls. Additionally, two of the 17β-E2 treated animals had to be euthanized due to vaginal hyperplasia. These findings are consistent with what has been seen with high bolus doses of estrogens, which are known to cause permanent changes in rodent reproductive function (50, 51) including growth of uterine tissue in rodents (52). As such, the initial release of 17 β-E2 following implantation could have contributed to the observed uterine, vaginal and pituitary hyperplasia. Additionally, the few studies using mice that have measured 17 β-E2 concentrations as an end parameter have reported abnormal pharmacological responses (49) or strikingly high 17 β-E2 concentrations (53, 54), even as long as 35 to 90 days after implantation of the pellets (5558).

Most reports using IRA pellets sample for the levels of 17 β-E2 or P4 at the termination of the study [for example, see (49, 5558)], and only a few of these studies report levels of the ovarian hormones within or near physiologically relevant concentrations at that time (49, 57). Regrettably, many reports state their pellets produce physiological levels of 17 β-E2 based upon technical information provided by IRA [for example see (59, 60)].

Based upon our observation of a large initial release of steroid hormones after implantation of these pellets, we urge caution in the interpretation of results obtained from studies that use of these hormone formulations. For example, Theodorsson and Theodorsson reported negative effects of estrogen, when delivered using the IRA pellet (1.5mg), in a transient middle cerebral artery occlusion (MCAO) model of cerebral ischemia (61). This is in sharp contrast to numerous reports that support a benefit of estrogen treatment, where estrogen reduces the lesion size following experimental stroke (11, 16, 62, 63). We argue that this discrepancy was due to differences in the levels of estradiol to which the animal was exposed. The high concentrations may have desensitized the brain to the protective effects of estrogen (potentially through receptor downregulation). Thus, a more accurate conclusion that should have been made is that “high dose” estradiol is not beneficial in preventing damage associated with transient MCAO.

Thus, these animal studies point to the importance of delivering the appropriate levels of estrogen, and as such, question whether persistent delivery of a high dose of estrogen and/or progestin (as is done in most hormone therapy regimens) is the appropriate means of delivering a therapeutic or protective dose of hormone.

4.4. Caveats associated with animal models

Transgenic knockout animals have been valuable in defining potential mechanisms by which estrogens exert their protective effects. However, these models are not without their limitations. For example, in determining the relevant estrogen receptor in the neuroprotective effects of estrogen against experimental stroke, some studies support the role of ER-α(64) while others implicate ER-β in mediating estrogen-induced protection (65). For example, Dubal and colleagues described that the protective effects of low dose estrogen (resulting in plasma levels that are ~ 25 pg/ml) against experimentally-induced stroke was abolished in ER-α knockout (ERαKO) mice (10). However, due to the apparent loss of negative feedback regulation at the level of the hypothalamus, ERαKO mice are exposed to much higher levels of estrogen than their wild-type counterparts (66). As such, the “threshold” for the protective effects of estrogen may have been much higher. Consistent with this idea, administration of higher concentrations of estradiol (~ 200 pg/ml) to ERαKO mice was effective at reducing infarct volume (67, 68). Thus, depending on the region of the brain, either ER-α or ER-β may be involved in mediating estrogen’s protective effects.

5. PERSPECTIVE

The studies described herein strongly support the value of animal models in testing the neuroprotective potential for estrogens and progestins. In fact, numerous in vivo paradigms have supported the ability of estradiol and progesterone to protect against a variety of insults. And though these models have several advantages relative to human clinical trials, at least with regards to the ability to manipulate and control experimental variables, it is imperative that we also recognize the caveats and limitations of each model. Only by recognizing such caveats will we be able to effectively translate these findings to human studies and later, implement them toward the development of novel hormone-based therapies for the menopause and the postmenopausal period, during which the risk for various disorders increases.

Acknowledgments

Some of the work described herein was supported, in part, by NIH grants AG 10485, AG 22550, AG 26672.

References

  • 1.Luine VN, Richards ST, Wu VY, Beck KD. Estradiol enhances learning and memory in a spatial memory task and effects levels of monoaminergic neurotransmitters. Horm Behav. 1998;34:149–62. doi: 10.1006/hbeh.1998.1473. [DOI] [PubMed] [Google Scholar]
  • 2.Singh M, Meyer EM, Millard WJ, Simpkins JW. Ovarian steroid deprivation results in a reversible learning impairment and compromised cholinergic function in female Sprague-Dawley rats. Brain Res. 1994;644:305–12. doi: 10.1016/0006-8993(94)91694-2. [DOI] [PubMed] [Google Scholar]
  • 3.Singh M, Meyer EM, Simpkins JW. The effect of ovariectomy and estradiol replacement on brain-derived neurotrophic factor messenger ribonucleic acid expression in cortical and hippocampal brain regions of female Sprague-Dawley rats. Endocrinology. 1995;136:2320–4. doi: 10.1210/endo.136.5.7720680. [DOI] [PubMed] [Google Scholar]
  • 4.Woolley CS, McEwen BS. Roles of estradiol and progesterone in regulation of hippocampal dendritic spine density during the estrous cycle in the rat. J Comp Neurol. 1993;336:293–306. doi: 10.1002/cne.903360210. [DOI] [PubMed] [Google Scholar]
  • 5.Luine VN. Estradiol increases choline acetyltransferase activity in specific basal forebrain nuclei and projection areas of female rats. Exp Neurol. 1985;89:484–90. doi: 10.1016/0014-4886(85)90108-6. [DOI] [PubMed] [Google Scholar]
  • 6.Sohrabji F, Miranda RC, Toran-Allerand CD. Estrogen differentially regulates estrogen and nerve growth factor receptor mRNAs in adult sensory neurons. J Neurosci. 1994;14:459–71. doi: 10.1523/JNEUROSCI.14-02-00459.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Jezierski MK, Sohrabji F. Region- and peptide-specific regulation of the neurotrophins by estrogen. Brain Res Mol Brain Res. 2000;85:77–84. doi: 10.1016/s0169-328x(00)00244-8. [DOI] [PubMed] [Google Scholar]
  • 8.Gibbs RB. Levels of trkA and BDNF mRNA, but not NGF mRNA, fluctuate across the estrous cycle and increase in response to acute hormone replacement. Brain Res. 1998;787:259–68. doi: 10.1016/s0006-8993(97)01511-4. [DOI] [PubMed] [Google Scholar]
  • 9.Alkayed NJ, Harukuni I, Kimes AS, London ED, Traystman RJ, Hurn PD. Gender-linked brain injury in experimental stroke. Stroke. 1998;29:159–65. 166 . doi: 10.1161/01.str.29.1.159. [DOI] [PubMed] [Google Scholar]
  • 10.Dubal DB, Kashon ML, Pettigrew LC, Ren JM, Finklestein SP, Rau SW, Wise PM. Estradiol protects against ischemic injury. J Cereb Blood Flow Metab. 1998;18:1253–8. doi: 10.1097/00004647-199811000-00012. [DOI] [PubMed] [Google Scholar]
  • 11.Simpkins JW, Rajakumar G, Zhang YQ, Simpkins CE, Greenwald D, Yu CJ, Bodor N, Day AL. Estrogens may reduce mortality and ischemic damage caused by middle cerebral artery occlusion in the female rat. J Neurosurg. 1997;87:724–30. doi: 10.3171/jns.1997.87.5.0724. [DOI] [PubMed] [Google Scholar]
  • 12.He Z, He YJ, Day AL, Simpkins JW. Proestrus levels of estradiol during transient global cerebral ischemia improves the histological outcome of the hippocampal CA1 region: perfusion-dependent and-independent mechanisms. J Neurol Sci. 2002;193:79–87. doi: 10.1016/s0022-510x(01)00648-7. [DOI] [PubMed] [Google Scholar]
  • 13.Sudo S, Wen TC, Desaki J, Matsuda S, Tanaka J, Arai T, Maeda N, Sakanaka M. Beta-estradiol protects hippocampal CA1 neurons against transient forebrain ischemia in gerbil. Neurosci Res. 1997;29:345–54. doi: 10.1016/s0168-0102(97)00106-5. [DOI] [PubMed] [Google Scholar]
  • 14.Fukuda K, Yao H, Ibayashi S, Nakahara T, Uchimura H, Fujishima M, Hall ED. Ovariectomy exacerbates and estrogen replacement attenuates photothrombotic focal ischemic brain injury in rats. Stroke. 2000;31:155–60. doi: 10.1161/01.str.31.1.155. [DOI] [PubMed] [Google Scholar]
  • 15.Mendelowitsch A, Ritz MF, Ros J, Langemann H, Gratzl O. 17beta-Estradiol reduces cortical lesion size in the glutamate excitotoxicity model by enhancing extracellular lactate: a new neuroprotective pathway. Brain Res. 2001;901:230–6. doi: 10.1016/s0006-8993(01)02359-9. [DOI] [PubMed] [Google Scholar]
  • 16.Yang SH, He Z, Wu SS, He YJ, Cutright J, Millard WJ, Day AL, Simpkins JW. 17-beta estradiol can reduce secondary ischemic damage and mortality of subarachnoid hemorrhage. J Cereb Blood Flow Metab. 2001;21:174–81. doi: 10.1097/00004647-200102000-00009. [DOI] [PubMed] [Google Scholar]
  • 17.Culmsee C, Vedder H, Ravati A, Junker V, Otto D, Ahlemeyer B, Krieg JC, Krieglstein J. Neuroprotection by estrogens in a mouse model of focal cerebral ischemia and in cultured neurons: evidence for a receptor-independent antioxidative mechanism. J Cereb Blood Flow Metab. 1999;19:1263–9. doi: 10.1097/00004647-199911000-00011. [DOI] [PubMed] [Google Scholar]
  • 18.Chen J, Xu W, Jiang H. 17 beta-estradiol protects neurons from ischemic damage and attenuates accumulation of extracellular excitatory amino acids. Anesth Analg. 2001;92:1520–3. doi: 10.1097/00000539-200106000-00033. [DOI] [PubMed] [Google Scholar]
  • 19.Jiang N, Chopp M, Stein D, Feit H. Progesterone is neuroprotective after transient middle cerebral artery occlusion in male rats. Brain Res. 1996;735:101–7. doi: 10.1016/0006-8993(96)00605-1. [DOI] [PubMed] [Google Scholar]
  • 20.Kumon Y, Kim SC, Tompkins P, Stevens A, Sakaki S, Loftus CM. Neuroprotective effect of postischemic administration of progesterone in spontaneously hypertensive rats with focal cerebral ischemia. J Neurosurg. 2000;92:848–52. doi: 10.3171/jns.2000.92.5.0848. [DOI] [PubMed] [Google Scholar]
  • 21.Morali G, Letechipia-Vallejo G, Lopez-Loeza E, Montes P, Hernandez-Morales L, Cervantes M. Post-ischemic administration of progesterone in rats exerts neuroprotective effects on the hippocampus. Neurosci Lett. 2005;382:286–90. doi: 10.1016/j.neulet.2005.03.066. [DOI] [PubMed] [Google Scholar]
  • 22.Chen J, Chopp M, Li Y. Neuroprotective effects of progesterone after transient middle cerebral artery occlusion in rat. J Neurol Sci. 1999;171:24–30. doi: 10.1016/s0022-510x(99)00247-6. [DOI] [PubMed] [Google Scholar]
  • 23.Cervantes M, Gonzalez-Vidal MD, Ruelas R, Escobar A, Morali G. Neuroprotective effects of progesterone on damage elicited by acute global cerebral ischemia in neurons of the caudate nucleus. Arch Med Res. 2002;33:6–14. doi: 10.1016/s0188-4409(01)00347-2. [DOI] [PubMed] [Google Scholar]
  • 24.Roof RL, Hoffman SW, Stein DG. Progesterone protects against lipid peroxidation following traumatic brain injury in rats. Mol Chem Neuropathol. 1997;31:1–11. doi: 10.1007/BF02815156. [DOI] [PubMed] [Google Scholar]
  • 25.Pettus EH, Wright DW, Stein DG, Hoffman SW. Progesterone treatment inhibits the inflammatory agents that accompany traumatic brain injury. Brain Res. 2005;1049:112–9. doi: 10.1016/j.brainres.2005.05.004. [DOI] [PubMed] [Google Scholar]
  • 26.Roof RL, Hall ED. Gender differences in acute CNS trauma and stroke: neuroprotective effects of estrogen and progesterone. J Neurotrauma. 2000;17:367–88. doi: 10.1089/neu.2000.17.367. [DOI] [PubMed] [Google Scholar]
  • 27.Ibanez C, Shields SA, El-Etr M, Leonelli E, Magnaghi V, Li WW, Sim FJ, Baulieu EE, Melcangi RC, Schumacher M, Franklin RJ. Steroids and the reversal of age-associated changes in myelination and remyelination. Prog Neurobiol. 2003;71:49–56. doi: 10.1016/j.pneurobio.2003.09.002. [DOI] [PubMed] [Google Scholar]
  • 28.Gonzalez Deniselle MC, Lopez Costa JJ, Gonzalez SL, Labombarda F, Garay L, Guennoun R, Schumacher M, De Nicola AF. Basis of progesterone protection in spinal cord neurodegeneration. J Steroid Biochem Mol Biol. 2002;83:199–209. doi: 10.1016/s0960-0760(02)00262-5. [DOI] [PubMed] [Google Scholar]
  • 29.Gonzalez Deniselle MC, Lopez-Costa JJ, Saavedra JP, Pietranera L, Gonzalez SL, Garay L, Guennoun R, Schumacher M, De Nicola AF. Progesterone neuroprotection in the Wobbler mouse, a genetic model of spinal cord motor neuron disease. Neurobiol Dis. 2002;11:457–68. doi: 10.1006/nbdi.2002.0564. [DOI] [PubMed] [Google Scholar]
  • 30.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;291:2947–58. doi: 10.1001/jama.291.24.2947. [DOI] [PubMed] [Google Scholar]
  • 31.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;291:2959–68. doi: 10.1001/jama.291.24.2959. [DOI] [PubMed] [Google Scholar]
  • 32.Wassertheil-Smoller S, Hendrix SL, Limacher M, Heiss G, Kooperberg C, Baird A, Kotchen T, Curb JD, Black H, Rossouw JE, Aragaki A, Safford M, Stein E, Laowattana S, Mysiw WJ. Effect of estrogen plus progestin on stroke in postmenopausal women: the Women’s Health Initiative: a randomized trial. Jama. 2003;289:2673–84. doi: 10.1001/jama.289.20.2673. [DOI] [PubMed] [Google Scholar]
  • 33.Shumaker SA, Legault C, Rapp SR, Thal L, Wallace RB, Ockene JK, Hendrix SL, Jones BN, 3rd, 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. 2003;289:2651–62. doi: 10.1001/jama.289.20.2651. [DOI] [PubMed] [Google Scholar]
  • 34.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. 2003;289:2663–72. doi: 10.1001/jama.289.20.2663. [DOI] [PubMed] [Google Scholar]
  • 35.Savonenko AV, Markowska AL. The cognitive effects of ovariectomy and estrogen replacement are modulated by aging. Neuroscience. 2003;119:821–30. doi: 10.1016/s0306-4522(03)00213-6. [DOI] [PubMed] [Google Scholar]
  • 36.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;24:733–43. doi: 10.1016/s0197-4580(02)00193-8. [DOI] [PubMed] [Google Scholar]
  • 37.Henderson VW, Espeland MA, Hogan PE, Rapp SR, Stefanick ML, Wactawski-Wende J, Johnson KC, Wassertheil-Smoller S, Freeman R, Curb D. Prior Use of Hormone Therapy and Incident Alzheimer’s Disease in the Women’s Health Initiative Memory Study. American Academy of Neurology. 2007 [Google Scholar]
  • 38.Nilsen J, Brinton RD. Impact of progestins on estrogen-induced neuroprotection: synergy by progesterone and 19-norprogesterone and antagonism by medroxyprogesterone acetate. Endocrinology. 2002;143:205–12. doi: 10.1210/endo.143.1.8582. [DOI] [PubMed] [Google Scholar]
  • 39.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;100:10506–11. doi: 10.1073/pnas.1334098100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Kaur P, Jodhka P, Underwood W, Bowles C, de Fiebre N, de Fiebre C, Singh M. Progesterone increases BDNF expression and protects against glutamate toxicity in a MAPK- and PI3K-dependent manner in cerebral cortical explants. J Neurosci Res. 2007 doi: 10.1002/jnr.21370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Littleton-Kearney MT, Klaus JA, Hurn PD. Effects of combined oral conjugated estrogens and medroxyprogesterone acetate on brain infarction size after experimental stroke in rat. J Cereb Blood Flow Metab. 2005;25:421–6. doi: 10.1038/sj.jcbfm.9600052. [DOI] [PubMed] [Google Scholar]
  • 42.Stein DG. The case for progesterone. Ann N Y Acad Sci. 2005;1052:152–69. doi: 10.1196/annals.1347.011. [DOI] [PubMed] [Google Scholar]
  • 43.van den Berg MP, Merkus P, Romeijn SG, Verhoef JC, Merkus FW. Uptake of melatonin into the cerebrospinal fluid after nasal and intravenous delivery: studies in rats and comparison with a human study. Pharm Res. 2004;21:799–802. doi: 10.1023/b:pham.0000026431.55383.69. [DOI] [PubMed] [Google Scholar]
  • 44.Lee BJ, Kang JK, Jung EY, Yun YW, Baek IJ, Yon JM, Lee YB, Sohn HS, Lee JY, Kim KS, Nam SY. Exposure to genistein does not adversely affect the reproductive system in adult male mice adapted to a soy-based commercial diet. J Vet Sci. 2004;5:227–34. [PubMed] [Google Scholar]
  • 45.Jung EY, Lee BJ, Yun YW, Kang JK, Baek IJ, Jurg MY, Lee YB, Sohn HS, Lee JY, Kim KS, Yu WJ, Do JC, Kim YC, Nam SY. Effects of exposure to genistein and estradiol on reproductive development in immature male mice weaned from dams adapted to a soy-based commercial diet. J Vet Med Sci. 2004;66:1347–54. doi: 10.1292/jvms.66.1347. [DOI] [PubMed] [Google Scholar]
  • 46.Gordon MN, Osterburg HH, May PC, Finch CE. Effective oral administration of 17 beta-estradiol to female C57BL/6J mice through the drinking water. Biol Reprod. 1986;35:1088–95. doi: 10.1095/biolreprod35.5.1088. [DOI] [PubMed] [Google Scholar]
  • 47.Jung ME, Yang SH, Brun-Zinkernagel AM, Simpkins JW. Estradiol protects against cerebellar damage and motor deficit in ethanol-withdrawn rats. Alcohol. 2002;26:83–93. doi: 10.1016/s0741-8329(01)00199-9. [DOI] [PubMed] [Google Scholar]
  • 48.Miller L, Hunt JS. Regulation of TNF-alpha production in activated mouse macrophages by progesterone. J Immunol. 1998;160:5098–104. [PubMed] [Google Scholar]
  • 49.Green PS, Bales K, Paul S, Bu G. Estrogen therapy fails to alter amyloid deposition in the PDAPP model of Alzheimer’s disease. Endocrinology. 2005;146:2774–81. doi: 10.1210/en.2004-1433. [DOI] [PubMed] [Google Scholar]
  • 50.Mobbs CV, Flurkey K, Gee DM, Yamamoto K, Sinha YN, Finch CE. Estradiol-induced adult anovulatory syndrome in female C57BL/6J mice: age-like neuroendocrine, but not ovarian, impairments. Biol Reprod. 1984;30:556–63. doi: 10.1095/biolreprod30.3.556. [DOI] [PubMed] [Google Scholar]
  • 51.Mobbs CV, Kannegieter LS, Finch CE. Delayed anovulatory syndrome induced by estradiol in female C57BL/6J mice: age-like neuroendocrine, but not ovarian, impairments. Biol Reprod. 1985;32:1010–7. doi: 10.1095/biolreprod32.5.1010. [DOI] [PubMed] [Google Scholar]
  • 52.Iguchi T, Takei T, Takase M, Takasugi N. Estrogen participation in induction of cervicovaginal adenosis-like lesions in immature mice exposed prenatally to diethylstilbestrol. Acta Anat (Basel) 1986;127:110–4. doi: 10.1159/000146264. [DOI] [PubMed] [Google Scholar]
  • 53.Rosenblum WI, el-Sabban F, Allen AD, Nelson GH, Bhatnagar AS, Choi S. Effects of estradiol on platelet aggregation in mouse mesenteric arterioles and ex vivo. Thromb Res. 1985;39:253–62. doi: 10.1016/0049-3848(85)90220-8. [DOI] [PubMed] [Google Scholar]
  • 54.Rosenblum WI, el-Sabban F, Allen AD, Nelson GH, Bhatnagar AS, Choi SC. Effects of estradiol on platelet aggregation in cerebral microvessels of mice. Stroke. 1985;16:980–4. doi: 10.1161/01.str.16.6.980. [DOI] [PubMed] [Google Scholar]
  • 55.Chin M, Isono M, Isshiki K, Araki S, Sugimoto T, Guo B, Sato H, Haneda M, Kashiwagi A, Koya D. Estrogen and raloxifene, a selective estrogen receptor modulator, ameliorate renal damage in db/db mice. Am J Pathol. 2005;166:1629–36. doi: 10.1016/s0002-9440(10)62473-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Latham KA, Zamora A, Drought H, Subramanian S, Matejuk A, Offner H, Rosloniec EF. Estradiol treatment redirects the isotype of the autoantibody response and prevents the development of autoimmune arthritis. J Immunol. 2003;171:5820–7. doi: 10.4049/jimmunol.171.11.5820. [DOI] [PubMed] [Google Scholar]
  • 57.Greenberg LH, Slayden OD. Human endometriotic xenografts in immunodeficient RAG-2/gamma (c)KO mice. Am J Obstet Gynecol. 2004;190:1788–95. 1795–6 . doi: 10.1016/j.ajog.2004.02.047. [DOI] [PubMed] [Google Scholar]
  • 58.Tse J, Martin-McNaulty B, Halks-Miller M, Kauser K, DelVecchio V, Vergona R, Sullivan ME, Rubanyi GM. Accelerated atherosclerosis and premature calcified cartilaginous metaplasia in the aorta of diabetic male Apo E knockout mice can be prevented by chronic treatment with 17 beta-estradiol. Atherosclerosis. 1999;144:303–13. doi: 10.1016/s0021-9150(98)00325-6. [DOI] [PubMed] [Google Scholar]
  • 59.Dabrosin C, Gyorffy S, Margetts P, Ross C, Gauldie J. Therapeutic effect of angiostatin gene transfer in a murine model of endometriosis. Am J Pathol. 2002;161:909–18. doi: 10.1016/S0002-9440(10)64251-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Verdu EF, Deng Y, Bercik P, Collins SM. Modulatory effects of estrogen in two murine models of experimental colitis. Am J Physiol Gastrointest Liver Physiol. 2002;283:G27–36. doi: 10.1152/ajpgi.00460.2001. [DOI] [PubMed] [Google Scholar]
  • 61.Theodorsson A, Theodorsson E. Estradiol increases brain lesions in the cortex and lateral striatum after transient occlusion of the middle cerebral artery in rats: No effect of ischemia on galanin in the stroke area but decreased levels in the hippocampus. Peptides. 2005 doi: 10.1016/j.peptides.2005.04.013. [DOI] [PubMed] [Google Scholar]
  • 62.Shi J, Bui JD, Yang SH, He Z, Lucas TH, Buckley DL, Blackband SJ, King MA, Day AL, Simpkins JW. Estrogens decrease reperfusion-associated cortical ischemic damage: an MRI analysis in a transient focal ischemia model. Stroke. 2001;32:987–92. doi: 10.1161/01.str.32.4.987. [DOI] [PubMed] [Google Scholar]
  • 63.Wen Y, Yang S, Liu R, Perez E, Yi KD, Koulen P, Simpkins JW. Estrogen attenuates nuclear factor-kappa B activation induced by transient cerebral ischemia. Brain Res. 2004;1008:147–54. doi: 10.1016/j.brainres.2004.02.019. [DOI] [PubMed] [Google Scholar]
  • 64.Gollapudi L, Oblinger MM. Stable transfection of PC12 cells with estrogen receptor (ERalpha): protective effects of estrogen on cell survival after serum deprivation. J Neurosci Res. 1999;56:99–108. doi: 10.1002/(SICI)1097-4547(19990401)56:1<99::AID-JNR13>3.0.CO;2-G. [DOI] [PubMed] [Google Scholar]
  • 65.Sawada H, Ibi M, Kihara T, Urushitani M, Honda K, Nakanishi M, Akaike A, Shimohama S. Mechanisms of antiapoptotic effects of estrogens in nigral dopaminergic neurons. Faseb J. 2000;14:1202–14. doi: 10.1096/fasebj.14.9.1202. [DOI] [PubMed] [Google Scholar]
  • 66.Couse JF, Curtis SW, Washburn TF, Lindzey J, Golding TS, Lubahn DB, Smithies O, Korach KS. Analysis of transcription and estrogen insensitivity in the female mouse after targeted disruption of the estrogen receptor gene. Mol Endocrinol. 1995;9:1441–54. doi: 10.1210/mend.9.11.8584021. [DOI] [PubMed] [Google Scholar]
  • 67.McCullough LD, Alkayed NJ, Traystman RJ, Williams MJ, Hurn PD. Postischemic estrogen reduces hypoperfusion and secondary ischemia after experimental stroke. Stroke. 2001;32:796–802. doi: 10.1161/01.str.32.3.796. [DOI] [PubMed] [Google Scholar]
  • 68.Sugo N, Alkayed NJ, Blizzard KK, McCullough LD, Traystman RJ, Crain BJ, Korach KS, Hurn PD. Estrogen Protects ischemic Brain of Estrogen Receptor-alpha or -beta knockout mice. Society for Neuroscience 31st Annual Meeting. 2001 Abstract #: 437.13. [Google Scholar]

RESOURCES