By: Meharvan Singh and Chang Su


Abstract


Both progesterone and estradiol have well-described neuroprotective effects against numerous insults
in a variety of cell culture models, animal models and in humans. However, the efficacy of these
hormones may depend on a variety of factors, including the type of hormone used (ex. progesterone
versus medroxyprogesterone acetate), the duration of the postmenopausal period prior to initiating the
hormone intervention, and potentially, the age of the subject. The latter two factors relate to the
proposed existence of a “window of therapeutic opportunity” for steroid hormones in the brain. While
such a window of opportunity has been described for estrogen, there is a paucity of information to
address whether such a window of opportunity exists for progesterone and its related progestins. Here,
we review known cellular mechanisms likely to underlie the protective effects of progesterone and
furthermore, describe key differences in the neurobiology of progesterone and the synthetic progestin,
medroxyprogesterone acetate (MPA). Based on the latter, we offer a model that defines some of the key
cellular and molecular players that predict the neuroprotective efficacy of progesterone. Accordingly,
we suggest how changes in the expression or function of these cellular and molecular targets of
progesterone with age or prolonged duration of hormone withdrawal (such as following surgical or
natural menopause) may impact the efficacy of progesterone.


Progesterone, like estrogen, is a gonadal steroid hormone that has classically been associated with
reproductive function, and accordingly, most of the literature related to progesterone and the brain 

focuses on the hypothalamus as the relevant target. However, extra-hypothalamic functions of these
steroids are now also apparent. Certainly, the widespread expression of receptors for progesterone and
estrogen would support the broader scope of gonadal hormone action, and such diversity of function is
underscored by the ability of gonadal hormones to regulate neurite outgrowth and differentiation
(Toran-Allerand, 1976; Toran-Allerand, 1980), survival, plasticity and regeneration (Behl et al., 1995;
Matsumoto and Arai, 1981; Simpkins et al., 1997), and also the ability to modulate cognitive function
(Luine et al., 1998; Singh et al., 1994).


With age, circulating gonadal hormone levels decline in both males and females. In women, however,
such age-associated decreases are much more dramatic as a result of the menopause. The menopause
occurs at an average age of 51 (Source: National Institute on Aging), and is characterized by a
precipitous decline in circulating gonadal hormones. Since the average lifespan of women has
increased to approximately 80 yrs of age (U.S. National Center for Health Statistics, National Vital
Statistics Reports (NVSR), Deaths: Final Data for 2007. Vol. 58, No. 19, May 2010), a substantial
portion of a woman’s life is spent in a hormone-deprived state. Given that many neuronal and glial cell
populations are normal targets of hormones such as progesterone, it is critical to address the impact of
this age-related hormonal decline on brain function. Further, the rather glaring discrepancy between
numerous basic (bench) science, epidemiological and clinical studies that documented the protective
effects of gonadal hormones on brain structure and function, and the results from the Women’s Health
Initiative (WHI), warrant clarification. In fact, these contrasting results underscore our limited
understanding of the neurobiology of these hormones. One hypothesis that has emerged to reconcile
some of the basic (bench) research with the results of the WHI is that there exists a “critical window”
of therapeutic opportunity for gonadal steroid hormones (Bohacek and Daniel, 2010; Daniel and
Bohacek, 2010; Dunkin et al., 2005; Henderson, 2006; Pinna et al., 2008; Sherwin, 2005; Smith et al.,
2010; Suzuki et al., 2007; Wu et al., 2011; Zandi et al., 2002). While such a window of opportunity has
been described for estrogen, it is still relatively unclear if such a window of opportunity exists for
progesterone and its related progestins. In this review, we summarize data from our laboratory and that
of others that describe the various mechanisms underlying progesterone-induced brain protection, and
further, through our current understanding of key differences in the neurobiology of progesterone and
the synthetic progestin, medroxyprogesterone acetate (MPA), offer a model that defines some of the
key cellular and molecular players that predict the neuroprotective efficacy of progesterone.
Accordingly, we suggest how changes in the expression or function of these cellular and molecular
targets of progesterone with age or prolonged duration of hormone withdrawal (such as following
surgical or natural menopause) may impact the efficacy of progesterone. We suggest that a better
understanding of the regulation of these cellular mediators of progesterone-induced protection could
yield insight into the extension of the therapeutic window of opportunity.


Progesterone-induced neuroprotection

Progesterone has been reported to exert protective effects in a variety of experimental models that
mimic certain pathogenic aspects of brain dysfunction seen with advanced age- or age-related
neurodegenerative diseases such as Alzheimer’s disease. For example, physiologically relevant
concentrations of progesterone have been shown to significantly attenuate oxidative injury resulting
from glutamate (Kaur et al., 2007; Nilsen and Brinton, 2002a; Nilsen and Brinton, 2002b; Nilsen and
Brinton, 2003) and glucose deprivation–induced toxicity (Goodman et al., 1996), and also protects
against FeSO - and amyloid β-peptide – induced toxicity in primary hippocampal cultures (Goodman
et al., 1996).


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 (Jiang et al., 1996). Interestingly, post-ischemic administration of progesterone was also
found to be protective (Kumon et al., 2000; Morali et al., 2005), and resulted in improvements in
various functional measures, including the rotarod test, and adhesive-backed somatosensory and
neurological scores (Chen et al., 1999). The ability of progesterone to protect even when administered
after the insult (albeit within a relatively narrow window) may suggest that both rapid/immediate and
long-term mechanisms of progesterone action are involved in the protective effects of progesterone.
Progesterone has also been shown to reduce the amount of cell death following an acute episode of
global ischemia (Cervantes et al., 2002), and is thought to be related to the ability of progesterone to
reduce lipid peroxidation, the generation of isoprostanes (Roof et al., 1997) and the expression of proinflammatory
genes (Pettus et al., 2005). It is worth pointing out that in these studies, the dose of
progesterone used may also be relevant since supraphysiological serum/plasma levels of progesterone
were achieved. With such doses, the resulting levels of allopregnanolone, the major progesterone
metabolite, could underlie some of the neuroprotective levels (see below for discussion of
allopregnanolone and neuroprotection).


Another model in which progesterone has been shown to exert protective effects is in the traumatic
brain injury (TBI) model. The administration of progesterone reduces cerebral edema for up to 24
hours after injury (Roof et al., 1996). In a rodent model of medial frontal cortex impact injury,
progesterone reduced complement factor C3, glial fibrillary acidic protein (GFAP), and nuclear factor
kappa beta (NFκB) (Pettus et al., 2005), all of which can be interpreted as protective mechanisms.
Progesterone also decreased the levels of lipid peroxidation in male rats when administered after TBI
(Roof and Hall, 2000).


Interestingly, there appears to be a sex difference in terms of the severity of impairment following TBI.
Females appeared to have less spatial learning impairments when compared to their male counterparts.
And though the lesion size was similar, females exhibited less ventricular dilation indicating lower
edema and water retention (Attella et al., 1987). In fact, direct assessments of edema reveal that
progesterone treatment significantly attenuates the level of edema seen in injured animals in contrast to
non-progesterone treated animals that had undergone experimental TBI (Roof et al., 1996).


The protective effects of progesterone are also evident in other regions of the central nervous system in
addition to the hippocampus and cerebral cortex. For example, progesterone has also been shown to
have a beneficial effect on spinal cord contusion injuries as supported by the work of Thomas et al.
who found that there was a marked reduction in the size of the lesion and a prevention of secondary
neuronal loss with progesterone treatment (Thomas et al., 1999). Further support for progesterone’s
protective actions in the spinal cord comes from the observation that progesterone has been shown to
promote morphological and functional recovery in the Wobbler mouse, an animal model of spinal cord
degeneration (Gonzalez Deniselle et al., 2002a; Gonzalez Deniselle et al., 2002b). Progesterone can
also induce re-myelination as supported by the increased expression of myelin proteins in the damaged
sciatic nerves of both young adult rats and in 22–24-month-old males (Ibanez et al., 2003). Thus,
progesterone may be of potential therapeutic benefit in diseases where demyelination is an important
component of its pathogenesis.


While the studies described above were all derived from animal models and cell/tissue culture models,
it is worth mentioning that a relatively recently completed phase II, randomized, double-blind, placebocontrolled
clinical trial assessing the efficacy of progesterone treatment for acute traumatic brain injury
yielded promising results. The data suggested that progesterone treatment can improve functional
recovery, at least when administered to those who experienced moderate, but not severe, traumatic
brain injury (Junpeng et al., 2011; Vandromme et al., 2008; Wali et al., 2011; Wright et al., 2007; Xiao
et al., 2008).


Mechanisms underlying progesterone’s protective effects
Numerous mechanisms of action likely underlie the protective effects of progesterone. The classical
genomic mechanism of progesterone action, for example, may be involved in the regulation of
neurotrophin expression (Kaur et al., 2007), which in turn, can promote cell survival. Alternatively,
progesterone may act through novel receptor systems, such as a membrane associated progesterone
receptor or the sigma receptor (another putative receptor for progesterone), to activate certain signal
transduction pathways, which in turn, triggers cellular events that are relevant and important for
neuroprotection (Singh and Su, 2012; Su et al., 2012b). Additionally, major metabolites of
progesterone, such as allopregnanolone, have been reported to participate in the neuroprotective effects
of progesterone (Ciriza et al., 2004).


With regards to the relationship between progesterone and neurotrophins, we (Kaur et al., 2007; Singh
et al., 1995) and others (Gonzalez Deniselle et al., 2007; Gonzalez et al., 2004; Sohrabji et al., 1995)
have shown that steroid hormones, including progesterone, increase the expression of BDNF. Further,
we found that neurotrophin signaling was necessary for progesterone induced protection (Jodhka et al.,
2009). With respect to “non-genomic” or cell signaling mechanisms underlying progesterone’s
protective effects, progesterone has been shown to elicit rapid effects on specific signaling pathways
including the cAMP/PKA (Collado et al., 1985), MAPK (ERK1/2) (Migliaccio et al., 1998; Nilsen and
Brinton, 2002a; Singh, 2001) and the PI-3K/Akt pathway (Singh, 2001), all of which have been
implicated in mediating neuroprotective effects. Progesterone-induced neuroprotection has not only
been correlated with activation of the MAPK and Akt signaling pathways (Nilsen and Brinton, 2002a;
Nilsen and Brinton, 2003) but has also been shown to depend on the activation of these pathways (Kaur
et al., 2007). Activation of these signaling pathways, in turn, may also lead to increased expression of
anti-apoptotic proteins such as Bcl-2 (Nilsen and Brinton, 2002a).


Another mechanism by which progesterone can exert protective effects is through its metabolites,
which in turn, can interact with membrane-associated receptors coupled to ion-channels, such as the
GABA receptor system (see (Deutsch et al., 1992) for review). Such metabolites include
allopregnanolone (or 3α, 5α tetrahydroprogesterone), which bind to discrete sites within the
hydrophobic domain of the GABA receptor complex, and result in the potentiation of GABA-induced
chloride conductance. Indeed, allopregnanolone has been suggested to play a role in mediating the

protective effects of progesterone (Ardeshiri et al., 2006; Djebaili et al., 2004; He et al., 2004a; He et
al., 2004b; Sayeed et al., 2009; Vitarbo et al., 2004). In addition to the effects of allopregnanolone on
the GABAA receptor, as outlined above, allopregnanolone may also elicit its protective effects through
its actions on the mitochondria (Robertson et al., 2006). For example, allopregnanolone was reported to
inhibit currents associated with the opening of the mitochondrial permeability transition pore (mtPTP)
(Sayeed et al., 2009), and as such, may help reduce the potential apoptotic consequences of mtPTP
opening (such as cytochrome c release) during insult or injury.


In addition to the allosteric effects described above, progesterone itself may have non-allosteric
influences on the GABAA receptor. Progesterone may influence the GABAA receptor via the
activation of a signal transduction pathway, which in turn, influences GABA-gated currents through
phosphorylation of discrete sites within certain subunits of the GABAA receptor (Bell-Horner et al.,
2006; Vasan et al., 2003). Since the regulation of the GABAA receptor has been shown to modulate
cell survival, particularly in models of excitotoxicity, the regulation of the GABAA receptor by
progesterone may be relevant to the protective effect of progesterone seen against kainate-induced
seizure activity and subsequent cell death (Hoffman et al., 2003).


Receptor pharmacology of progesterone’s protective effects
It is clear that the classical, intracellular/nuclear PR certainly plays an important role in mediating the
effects of progesterone. For example, our laboratory has determined that the ability of progesterone to
increase the expression (mRNA and protein levels) of brain-derived neurotrophic factor (BDNF), a key
mediator of progesterone’s protective effects, requires the classical PR (Jodhka et al., 2009). Further,
Cai and colleagues (Cai et al., 2008) have implicated the classical/intracellular PR in the protective
effects of progesterone against an experimental model (middle cerebral artery occlusion) of stroke.
More recently, Liu et al., describe the key role of the classical PR in neuroprotection after experimental
stroke (Liu et al., 2012), using the PR knockout model. This experimental model (at least the
homozygous knockout) has clear reproductive behavior deficits (Conneely and Lydon, 2000), but does
not appear to, in and of itself, result in overt phenotypic changes in brain morphology.


However, evidence also exists for alternative mechanisms of action, including that which involves
integral membrane progesterone receptors. For example, the effect of progesterone has been reported in
the brain of PR knock-out (PRKO) mice (Krebs et al., 2000), suggesting PRs other than the classical
PR may mediate the effect of progesterone in the CNS. In fact, several lines of evidence recently
obtained suggest that the rapid effects of progesterone are mediated by cell membrane-associated PRs
expressed in the brain (Balasubramanian et al., 2008a; Balasubramanian et al., 2008b; Liu et al., 2009;
Tokmakov and Fukami, 2009). If nothing else, progesterone’s high degree of lipophilicity (having a
logP value, or octanol/water partition coefficient, of approximately 4), may be consistent with the idea
that progesterone interacts with a plasma membrane associated receptor.


Membrane receptors for progesterone, though proposed for many years based on the existence of
specific, displaceable binding sites observed in synaptosomal membrane preparations (Ke and
Ramirez, 1990; Towle and Sze, 1983), have only recently been cloned. For example, Zhu and
colleagues discovered a novel membrane-associated progesterone receptor, termed mPR (Zhu et al.,
2003a), that has a predicted seven transmembrane-spanning domain, and is coupled to the Gi/o class of

G-proteins (Zhu et al., 2003b). Other membrane progesterone receptors include 25-Dx (also called
Pgrmc1) (Falkenstein et al., 1998; Krebs et al., 2000; Meyer et al., 1996), that is involved in numerous
aspects of cell function, ranging from neuronal development (Sakamoto et al., 2004), steroidogenesis
(Min et al., 2004), regulation of CSF production and osmoregulation (Meffre et al., 2005), and the
regulation of reproductive behavior (Krebs et al., 2000). Though our laboratory has determined that the
classical PR, mPRα, mPRβ and Pgrmc1 are expressed in our experimental models of the CNS wherein
we have shown progesterone-induced neuroprotection, we have determined that while progesterone’s
ability to increase BDNF expression is dependent on the classical PR (Jodhka et al., 2009), it is the
membrane associated receptor, Pgrmc1, that mediates the effect of progesterone on BDNF release (Su
et al., 2012a). Further, this effect on BDNF release appears to be mediated by ERK5 (Su et al., 2011;
Su et al., 2012a). As such, we believe that these effects on BDNF are critical to progesterone’s
neuroprotective capacity (Kaur et al., 2007). Moreover, a putative ligand of membrane associated
progesterone receptors, the BSA-conjugated progesterone (P4-BSA), that does not bind to the
intracellular localized classical PR, fails to increase BDNF levels but yet, is effective in increasing the
phosphorylation of ERK1/2 (Jodhka et al., 2009), an effect that appears to be mediate by a membrane
receptor other than Pgrmc1 or the classical PR (Su et al., 2012a), and is yet another proposed mediator
of progesterone’s neuroprotective effects (Kaur et al., 2007). As such, the ability of a progestin to have
maximal neuroprotective efficacy may depend on the complement of progesterone receptors that it is
capable of binding/activating.


And finally, progesterone has also been found to interact with sigma 1 (σ1) receptor (Selmin et al.,
1996; Seth et al., 1998). Given the reported role of the sigma 1 receptor in neuroprotection (for review,
see (Maurice et al., 2006)), this mechanism may also be relevant to progesterone’s protective actions.


The neurobiology of progesterone versus medroxyprogesterone acetate

Recent results from the Women’s Health Initiative-Memory Study (WHIMS) failed to reveal beneficial
effects in reducing the risk of Alzheimer’s disease or “all-cause” dementia. As a consequence, these
reports left the field unsettled as to the future of hormone therapy. Since the publication of these
studies, it became apparent that there were important caveats to the data that needed to be considered.
Among these included consideration of the type of hormone used. Indeed there are important
differences in the neurobiology of two major progestins, the “natural” progestin, progesterone, and the
synthetic medroxyprogesterone acetate (MPA), the most commonly used progestin in hormone therapy
regimens.


Medroxyprogesterone acetate (MPA), a synthetic progestin derived from 17α-hydroxyprogesterone, is
often used in conjunction with estrogens to reduce the risk of certain cancers (cervical cancer, for
example) resulting from unopposed estrogen therapy (Gambrell, 1986; Hirvonen, 1996). First, though
both progesterone and MPA can bind to the classical PR, it is important to recognize that there are
important pharmacological and pharmacokinetic differences between MPA and progesterone. For
example, orally administered MPA does not undergo any first pass effects (Schindler et al., 2003),
unlike progesterone. Furthermore, MPA has little binding affinity for sex hormone binding globulin
(Schindler et al., 2003). In addition to differences in bioavailability and half-life, MPA also displays
many non-progestagenic effects (Schindler et al., 2003), including the ability to bind to the androgen
receptor (AR) where it acts as a partial agonist (Winneker et al., 2003) with a binding affinity (Kd) of
approximately 2.1 nM (Hackenberg et al., 1990). Progesterone, in contrast, does not bind to the AR
(Schindler et al., 2003). MPA can also bind to, and activate, glucocorticoid receptors (Koubovec et al.,
2005; Schindler et al., 2003) with an effective concentration (EC50) that is nearly 300-fold lower than
that for progesterone (Koubovec et al., 2005).


While progesterone and MPA may be equally effective at reducing the uterotrophic effects of unopposed
estrogen treatment, their effects on the brain are far from identical. In fact, it has become
increasingly clear that while progesterone is neuroprotective, MPA is not. For example, our laboratory
described that in cerebral cortical explants, the difference in neuroprotective efficacy between
progesterone and MPA may have been attributed to their differential regulation of BDNF. Specifically,
while progesterone increased both the mRNA and protein levels of BDNF in the cerebral cortex, MPA
treatment resulted in a substantial inhibition (Jodhka et al., 2009). Combined with the observation that
progesterone’s protective effects may be dependent on neurotrophin signaling (Kaur et al., 2007), this
inhibition of BDNF expression by MPA may actually suggest that it have adverse consequences to
brain function. Similarly, the Brinton laboratory has shown in hippocampal cultures that while
progesterone is protective, MPA is not. In this model, the protective effects of progesterone appeared to
be mediated, in part, by attenuating the glutamate-induced increase in intracellular Ca levels. MPA,
in contrast, failed to alter the glutamate-induced influx of Ca . Of significance was that MPA not only
failed to elicit protective effects, but also blocked the beneficial effect of estradiol. In sharp contrast,
progesterone did not inhibit the effect of estradiol (Nilsen and Brinton, 2002b). Furthermore, while
some of the neuroprotective effects of progesterone are mediated by its neuroactive metabolite,
alloprognanolone (see discussion above), it is unclear if MPA is a substrate for the progesterone
metabolizing enzymes 5alpha-reductase and 3alpha-hydroxysteroid dehydrogenase. If anything, MPA
has been shown to inhibit the biosynthetic enzymes associated with the conversion of progesterone to
allopregnanolone. Thus, both the inability of MPA to be converted to neuroactive steroid metabolites in
conjunction with its effect in reducing potential conversion of progesterone to allopregnanolone may
contribute to its lack of neuroprotection.


As stated above, progesterone’s protective effects, in at least two neuronal models (cerebral cortical
neurons and hippocampal neurons), was dependent on activation of the ERK/MAPK pathway (Kaur et
al., 2007; Nilsen and Brinton, 2002b; Nilsen and Brinton, 2003). While both progesterone and MPA
can elicit ERK phosphorylation, only progesterone treatment resulted in nuclear translocation of ERK
(Nilsen and Brinton, 2003), the consequence of which is likely to regulate key genes, whose protein
products may enable more long term/sustainable protection. In fact, progesterone, but not MPA,
increased the expression of the anti-apoptotic Bcl-2 protein (Nilsen and Brinton, 2002a). And as
observed in the model of glutamate-induced Ca influx, MPA not only failed to increase expression of
Bcl-2, but actually inhibited that elicited by estradiol (Nilsen and Brinton, 2002b).


The disparity between the effects of progesterone and MPA has also been observed in vivo. For
example, a study using rhesus monkeys illustrated that combined treatment with estradiol and
progesterone protects against coronary vasospasm, whereas estradiol + MPA treatment did not
(Miyagawa et al., 1997). And once again, in contrast to the antagonistic effects of MPA on estrogen’s
effects, progesterone enhanced the protective effects of estrogen against exercise-induced myocardial

ischemia in post-menopausal women to, whereas MPA did not (Rosano et al., 2000). Moreover, 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 (Littleton-Kearney et al., 2005). The functional
antagonistic effects of MPA were also noted in the cholinergic system of monkeys, where MPA
administered in conjunction with CEE reduced choline acetyl transferase (ChAT) in such cognitionrelevant
areas of the brain as the medial septum (Gibbs et al., 2002). Similar consequences of MPA
were seen in the cardiovascular system of cynomolgus monkeys. Adams et al., demonstrated that
monkeys treated with CEE showed a 72% reduction in coronary artery atherosclerosis whereas there
were no benefits observed in CEE plus MPA group (Adams et al., 1997). Interestingly, with regards to
the traumatic brain injury model, MPA required a larger dose than progesterone 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 (Stein, 2005)).


Progesterone’s protective effects are influenced by age
Unfortunately, there is very limited information on how the cytoprotective effects of progesterone are
altered with age. A few studies have, however, suggested that progesterone protective effects are
diminished with age. For example, in one study progesterone significantly reduced the experimental
stroke-induced lesion volume in young adult (3 month old) ovariectomized (OVX) C57Bl/6 mice, but
had no effect on neurological outcome in older (12 month old) OVX mice (Gibson et al., 2011).
Further, recent data from our laboratory suggests that while progesterone increased BDNF expression
in the hippocampus of young adult rats (4 months of age), the response of the hippocampus of older
(10 month old) mice was significantly reduced (Singh, Su, Yang, Sumien and Forster, unpublished
data). Given the implicated importance of neurotrophin signaling in the protective effects of
progesterone, we suggest that deficits in capacity of progesterone to promote the synthesis and
availability of neurotrophins to other cells may underlie, at least in part, this diminished response with
age.


The biological basis for a window of opportunity for progesterone
Based on the mechanistic differences between progesterone and MPA that our lab and others have
described to explain why progesterone is protective but MPA is not, we propose that the biological
basis for a window of opportunity for progesterone may, in part, be influenced by: 1) the relative
expression, binding activity and distribution of the certain progesterone receptors in the brain and, 2)
potential changes in neurotrophin regulation and it’s associated signaling cascades coupling
progesterone receptors and their downstream targets.


Given the importance of BDNF as a cellular mediator of progesterone-induced brain protection (Kaur
et al., 2007), and the role of the PR and Pgrmc1 in mediating the effects of progesterone on BDNF
synthesis and release (Su et al., 2012a), respectively, we suggest that the complementary actions of
membrane and intracellular progesterone receptors are required to afford protection. Given that the
classical progesterone receptor is necessary for the induction of BDNF synthesis, but a membraneassociated
progesterone receptor (Pgrmc1) is required for progesterone-induced release of BDNF, we
propose that the two must act in tandem for protection to be afforded. As such, a decline in the

expression of either the PR or Pgrmc1 would contribute to a diminished response to progesterone,
within the context of cytoprotection, and thus, implicate these two receptors in defining a therapeutic
window of opportunity. Alternatively, since the protective effects of progesterone are noted to depend
on neurotrophin signaling, including activation of the ERK/MAPK or PI3K/Akt signaling pathways, a
change in the relative expression of the neurotrophins, their cognate receptors (Trk receptors) or
specific signaling proteins may similarly impact the capacity of progesterone to promote brain health.
Supporting the latter is the observation of an age-related decrease in ERK1/2 activity in the rat dentate
gyrus (McGahon et al., 1999) and cerebral cortex (Zhen et al., 1999).


While a finite window of therapeutic opportunity for estrogens in the aging brain have been defined to
some extent, along with some of the mechanisms to explain the altered response of the brain to
estrogens with age, we are unaware of any published data that addresses whether such a limited
window of opportunity for progesterone (or its related progestins) exists. Recent data from our
laboratory support the conclusion that there does, indeed, exist an age-associated loss of sensitivity of
the brain to progesterone (at least with respect to the regulation of BDNF expression). Further, we
believe that the data from mechanistic studies aimed at explaining the difference in neuroprotective
efficacy between progesterone and MPA offer unique insight into the mechanisms that may be required
for progesterone to elicit protective effects, and as such, provide a basis for explaining a potentially
limited window of opportunity of therapeutic efficacy for progesterone. Future studies, from our
laboratory and others, will undoubtedly provide more clarity to this important topic and consequently
has the potential to help refine the future of hormone therapy.


Acknowledgments
The work from our laboratory cited herein was supported, in part, from grants from the NIH
(AG022550 and AG027956), an IIRG from the Alzheimer’s Association and the Texas Garvey Fund to
MS, and an American Federation of Aging (AFAR) grant to CS.


References
1. Adams MR, Register TC, Golden DL, Wagner JD, Williams JK. Medroxyprogesterone acetate
antagonizes inhibitory effects of conjugated equine estrogens on coronary artery atherosclerosis.
Arterioscler Thromb Vasc Biol. 1997;17:217–221.
2. Ardeshiri A, Kelley MH, Korner IP, Hurn PD, Herson PS. Mechanism of progesterone
neuroprotection of rat cerebellar Purkinje cells following oxygen-glucose deprivation. Eur J
Neurosci. 2006;24:2567–2574.
3. Attella MJ, Nattinville A, Stein DG. Hormonal state affects recovery from frontal cortex lesions
in adult female rats. Behav Neural Biol. 1987;48:352–367. 

4. Balasubramanian B, Portillo W, Reyna A, Chen JZ, Moore AN, Dash PK, Mani SK.
Nonclassical mechanisms of progesterone action in the brain: I. Protein kinase C activation in the
hypothalamus of female rats. Endocrinology. 2008a;149:5509–5517.
5. Balasubramanian B, Portillo W, Reyna A, Chen JZ, Moore AN, Dash PK, Mani SK.
Nonclassical mechanisms of progesterone action in the brain: II. Role of calmodulin-dependent
protein kinase II in progesterone-mediated signaling in the hypothalamus of female rats.
Endocrinology. 2008b;149:5518–5526.
6. 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. [PubMed]
7. Bell-Horner CL, Dohi A, Nguyen Q, Dillon GH, Singh M. ERK/MAPK pathway regulates
GABAA receptors. J Neurobiol. 2006;66:1467–1474.
8. Bohacek J, Daniel JM. The beneficial effects of estradiol on attentional processes are dependent
on timing of treatment initiation following ovariectomy in middle-aged rats.
Psychoneuroendocrinology. 2010;35:694–705.
9. Cai W, Zhu Y, Furuya K, Li Z, Sokabe M, Chen L. Two different molecular mechanisms
underlying progesterone neuroprotection against ischemic brain damage. Neuropharmacology.
2008;55:127–138.
10. 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.
11. 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.
12. Ciriza I, Azcoitia I, Garcia-Segura LM. Reduced progesterone metabolites protect rat
hippocampal neurones from kainic acid excitotoxicity in vivo. J Neuroendocrinol. 2004;16:58–
63.
13. Collado ML, Rodriguez-Manzo G, Cruz ML. Effect of progesterone upon adenylate cyclase
activity and cAMP levels on brain areas. Pharmacol Biochem Behav. 1985;23:501–504.

14. Conneely OM, Lydon JP. Progesterone receptors in reproduction: functional impact of the A and
B isoforms. Steroids. 2000;65:571–577.
15. Daniel JM, Bohacek J. The critical period hypothesis of estrogen effects on cognition: Insights
from basic research. Biochim Biophys Acta. 2010;1800:1068–1076.
16. Deutsch SI, Mastropaolo J, Hitri A. GABA-active steroids: endogenous modulators of GABAgated
chloride ion conductance. Clin Neuropharmacol. 1992;15:352–364.
17. Djebaili M, Hoffman SW, Stein DG. Allopregnanolone and progesterone decrease cell death and
cognitive deficits after a contusion of the rat pre-frontal cortex. Neuroscience. 2004;123:349–
359.
18. Dunkin J, Rasgon N, Wagner-Steh K, David S, Altshuler L, Rapkin A. Reproductive events
modify the effects of estrogen replacement therapy on cognition in healthy postmenopausal
women. Psychoneuroendocrinology. 2005;30:284–296.
19. Falkenstein E, Schmieding K, Lange A, Meyer C, Gerdes D, Welsch U, Wehling M. Localization
of a putative progesterone membrane binding protein in porcine hepatocytes. Cell Mol Biol (Noisy-le-grand) 1998;44:571–578.

20. Gambrell RD., Jr The role of hormones in the etiology and prevention of endometrial cancer.
Clin Obstet Gynaecol. 1986;13:695–723.
21. Gibbs RB, Nelson D, Anthony MS, Clarkson TB. Effects of long-term hormone replacement and
of tibolone on choline acetyltransferase and acetylcholinesterase activities in the brains of
ovariectomized, cynomologus monkeys. Neuroscience. 2002;113:907–914.
22. Gibson CL, Coomber B, Murphy SP. Progesterone is neuroprotective following cerebral
ischaemia in reproductively ageing female mice. Brain. 2011;134:2125–2133.
23. 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. 2002a;83:199–209. [
24. 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. 2002b;11:457–468.
25. Gonzalez Deniselle MC, Garay L, Gonzalez S, Saravia F, Labombarda F, Guennoun R,
Schumacher M, De Nicola AF. Progesterone modulates brain-derived neurotrophic factor and
choline acetyltransferase in degenerating Wobbler motoneurons. Exp Neurol. 2007;203:406–414.
26. Gonzalez SL, Labombarda F, Gonzalez Deniselle MC, Guennoun R, Schumacher M, De Nicola
AF. Progesterone up-regulates neuronal brain-derived neurotrophic factor expression in the
injured spinal cord. Neuroscience. 2004;125:605–614.
27. Goodman Y, Bruce AJ, Cheng B, Mattson MP. Estrogens attenuate and corticosterone
exacerbates excitotoxicity, oxidative injury, and amyloid betapeptide toxicity in hippocampal
neurons. J Neurochem. 1996;66:1836–1844.
28. Hackenberg R, Hofmann J, Wolff G, Holzel F, Schulz KD. Down-regulation of androgen
receptor by progestins and interference with estrogenic or androgenic stimulation of mammary
carcinoma cell growth. J Cancer Res Clin Oncol. 1990;116:492–498.
29. He J, Evans CO, Hoffman SW, Oyesiku NM, Stein DG. Progesterone and allopregnanolone
reduce inflammatory cytokines after traumatic brain injury. Exp Neurol. 2004a;189:404–412.
30. He J, J, Hoffman SW, Stein DG. Allopregnanolone, a progesterone metabolite, enhances
behavioral recovery and decreases neuronal loss after traumatic brain injury. Restor Neurol
Neurosci. 2004b;22:19–31.
31. Henderson VW. Estrogen-containing hormone therapy and Alzheimer's disease risk:
understanding discrepant inferences from observational and experimental research.
Neuroscience. 2006;138:1031–1039.
32. Hirvonen E. Progestins. Maturitas. 1996;23(Suppl):S13–S18.
33. Hoffman GE, Moore N, Fiskum G, Murphy AZ. Ovarian steroid modulation of seizure severity
and hippocampal cell death after kainic acid treatment. Exp Neurol. 2003;182:124–134.
34. 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.

35. Jiang N, Chopp M, Stein D, Feit H. Progesterone is neuroprotective after transient middle
cerebral artery occlusion in male rats. Brain Res. 1996;735:101–107.
36. Jodhka PK, Kaur P, Underwood W, Lydon JP, Singh M. The differences in neuroprotective
efficacy of progesterone and medroxyprogesterone acetate correlate with their effects on brainderived
neurotrophic factor expression. Endocrinology. 2009;150:3162–3168.
37. Junpeng M, Huang S, Qin S. Progesterone for acute traumatic brain injury. Cochrane Database
Syst Rev. 2011:CD008409.
38. Kaur P, Jodhka PK, Underwood WA, Bowles CA, de Fiebre NC, de Fiebre CM, Singh M.
Progesterone increases brain-derived neuroptrophic factor expression and protects against
glutamate toxicity in a mitogen-activated protein kinase- and phosphoinositide-3 kinasedependent
manner in cerebral cortical explants. J Neurosci Res. 2007;85:2441–2449.
[PMC free article] [
39. Ke FC, Ramirez VD. Binding of progesterone to nerve cell membranes of rat brain using
progesterone conjugated to 125I-bovine serum albumin as a ligand. J Neurochem. 1990;54:467–
472.
40. Koubovec D, Ronacher K, Stubsrud E, Louw A, Hapgood JP. Synthetic progestins used in HRT
have different glucocorticoid agonist properties. Mol Cell Endocrinol. 2005;242:23–32.
41. Krebs CJ, Jarvis ED, Chan J, Lydon JP, Ogawa S, Pfaff DW. A membrane-associated
progesterone-binding protein 25-Dx, is regulated by progesterone in brain regions involved in
female reproductive behaviors. Proc Natl Acad Sci U S A. 2000;97:12816–12821.
42. 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;92:848–852.
43. 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;25:421–426.
44. Liu A, Margaill I, Zhang S, Labombarda F, Coqueran B, Delespierre B, Liere P, Marchand-
Leroux C, O'Malley BW, Lydon JP, De Nicola AF, Sitruk-Ware R, Mattern C, Plotkine M,
Schumacher M, Guennoun R. Progesterone Receptors: A Key for Neuroprotection in
Experimental Stroke. Endocrinology. 2012
45. Liu L, Wang J, Zhao L, Nilsen J, McClure K, Wong K, Brinton RD. Progesterone increases rat
neural progenitor cell cycle gene expression and proliferation via extracellularly regulated kinase
and progesterone receptor membrane components 1 and 2. Endocrinology. 2009;150:3186–3196.
46. Luine VN, Richards ST, Wu VY, Beck KD. Estradiol enhances learning and memory in a spatial
memory task and effects levels of monoaminergic neurotransmitters [In Process Citation] Horm
Behav. 1998;34:149–162.
47. Matsumoto A, Arai Y. Neuronal plasticity in the deafferented hypothalamic arcuate nucleus of
adult female rats and its enhancement by treatment with estrogen. J Comp Neurol. 1981;197:197–205.

48. Maurice T, Gregoire C, Espallergues J. Neuro(active)steroids actions at the neuromodulatory
sigma1 (sigma1) receptor: biochemical and physiological evidences, consequences in
neuroprotection. Pharmacol Biochem Behav. 2006;84:581–597.
49. McGahon B, Maguire C, Kelly A, Lynch MA. Activation of p42 mitogen-activated protein
kinase by arachidonic acid and trans-1-amino-cyclopentyl-1,3- dicarboxylate impacts on longterm
potentiation in the dentate gyrus in the rat: analysis of age-related changes. Neuroscience.
1999;90:1167–1175.
50. Meffre D, Delespierre B, Gouezou M, Leclerc P, Vinson GP, Schumacher M, Stein DG,
Guennoun R. The membrane-associated progesterone-binding protein 25-Dx is expressed in
brain regions involved in water homeostasis and is up-regulated after traumatic brain injury. J
Neurochem. 2005;93:1314–1326.
51. Meyer C, Schmid R, Scriba PC, Wehling M. Purification and partial sequencing of high-affinity
progesterone-binding site(s) from porcine liver membranes. Eur J Biochem. 1996;239:726–731.
52. Migliaccio A, Piccolo D, Castoria G, Di Domenico M, Bilancio A, Lombardi M, Gong W, Beato
M, Auricchio F. Activation of the Src/p21ras/Erk pathway by progesterone receptor via crosstalk
with estrogen receptor. Embo. J. 1998;17:2008–2018.
53. Min L, Takemori H, Nonaka Y, Katoh Y, Doi J, Horike N, Osamu H, Raza FS, Vinson GP,
Okamoto M. Characterization of the adrenal-specific antigen IZA (inner zone antigen) and its
role in the steroidogenesis. Mol Cell Endocrinol. 2004;215:143–148.
54. Miyagawa K, Vidgoff J, Hermsmeyer K. Ca2+ release mechanism of primate drug-induced
coronary vasospasm. Am J Physiol. 1997;272:H2645–H2654.
55. 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;382:286–290.
56. Nilsen J, Brinton RD. Impact of progestins on estrogen-induced neuroprotection: synergy by
progesterone and 19-norprogesterone and antagonism by medroxyprogesterone acetate.
Endocrinology. 2002a;143:205–212.
57. Nilsen J, Brinton RD. Impact of progestins on estradiol potentiation of the glutamate calcium
response. Neuroreport. 2002b;13:825–830.
58. 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.
59. Pettus EH, Wright DW, Stein DG, Hoffman SW. Progesterone treatment inhibits the
inflammatory agents that accompany traumatic brain injury. Brain Res. 2005;1049:112–119.
60. Pinna C, Cignarella A, Sanvito P, Pelosi V, Bolego C. Prolonged ovarian hormone deprivation
impairs the protective vascular actions of estrogen receptor alpha agonists. Hypertension.
2008;51:1210–1217.
61. Robertson CL, Puskar A, Hoffman GE, Murphy AZ, Saraswati M, Fiskum G. Physiologic
progesterone reduces mitochondrial dysfunction and hippocampal cell loss after traumatic brain
injury in female rats. Exp Neurol. 2006;197:235–243.
62. Roof RL, Duvdevani R, Heyburn JW, Stein DG. Progesterone rapidly decreases brain edema:
treatment delayed up to 24 hours is still effective. Exp Neurol. 1996;138:246–251.
63. Roof RL, Hoffman SW, Stein DG. Progesterone protects against lipid peroxidation following
traumatic brain injury in rats. Mol Chem Neuropathol. 1997;31:1–11.
64. Roof RL, Hall ED. Gender differences in acute CNS trauma and stroke: neuroprotective effects
of estrogen and progesterone. J Neurotrauma. 2000;17:367–388.
65. Rosano GM, Webb CM, Chierchia S, Morgani GL, Gabraele M, Sarrel PM, de Ziegler D, Collins
P. Natural progesterone, but not medroxyprogesterone acetate, enhances the beneficial effect of
estrogen on exercise-induced myocardial ischemia in postmenopausal women. J Am Coll
Cardiol. 2000;36:2154–2159.
66. Sakamoto H, Ukena K, Takemori H, Okamoto M, Kawata M, Tsutsui K. Expression and
localization of 25-Dx, a membrane-associated putative progesterone-binding protein, in the
developing Purkinje cell. Neuroscience. 2004;126:325–334.
67. Sayeed I, Parvez S, Wali B, Siemen D, Stein DG. Direct inhibition of the mitochondrial
permeability transition pore: a possible mechanism for better neuroprotective effects of
allopregnanolone over progesterone. Brain Res. 2009;1263:165–173.
68. Schindler AE, Campagnoli C, Druckmann R, Huber J, Pasqualini JR, Schweppe KW, Thijssen
JH. Classification and pharmacology of progestins. Maturitas. 2003;46(Suppl 1):S7–S16.
69. Selmin O, Lucier GW, Clark GC, Tritscher AM, Vanden Heuvel JP, Gastel JA, Walker NJ, Sutter
TR, Bell DA. Isolation and characterization of a novel gene induced by 2,3,7,8-
tetrachlorodibenzo-p-dioxin in rat liver. Carcinogenesis. 1996;17:2609–2615.
70. Seth P, Fei YJ, Li HW, Huang W, Leibach FH, Ganapathy V. Cloning and functional
characterization of a sigma receptor from rat brain. J Neurochem. 1998;70:922–931.
71. Sherwin BB. Surgical menopause, estrogen, and cognitive function in women: what do the
findings tell us? Ann N Y Acad Sci. 2005;1052:3–10.
72. Simpkins JW, Green PS, Gridley KE, Singh M, de Fiebre NC, Rajakumar G. Role of estrogen
replacement therapy in memory enhancement and the prevention of neuronal loss associated with
Alzheimer's disease. Am J Med. 1997;103:19S–25S.
73. 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;644:305–312.
74. 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–
2324.
75. Singh M. Ovarian hormones elicit phosphorylation of Akt and extracellular-signal regulated
kinase in explants of the cerebral cortex. Endocrine. 2001;14:407–415.
76. Singh M, Su C. Progesterone and neuroprotection. Horm Behav. 2012
Smith CC, Vedder LC, Nelson AR, Bredemann TM, McMahon LL. Duration of estrogen
deprivation, not chronological age prevents estrogen's ability to enhance hippocampal synaptic
physiology. Proc Natl Acad Sci U S A. 2010;107:19543–19548
78. Sohrabji F, Miranda RC, Toran-Allerand CD. Identification of a putative estrogen response
element in the gene encoding brain-derived neurotrophic factor. Proc Natl Acad Sci U S A.
1995;92:11110–11114.
79. Stein DG. The case for progesterone. Ann N Y Acad Sci. 2005;1052:152–169. [PubMed]
80. Su C, Underwood W, Rybalchenko N, Singh M. ERK1/2 and ERK5 have distinct roles in the
regulation of brain-derived neurotrophic factor expression. J Neurosci Res. 2011;89:1542–1550.
81. Su C, Cunningham RL, Rybalchenko N, Singh M. Progesterone Increases the Release of Brain-
Derived Neurotrophic Factor from Glia via Progesterone Receptor Membrane Component 1
(Pgrmc1)-Dependent ERK5 Signaling. Endocrinology. 2012a
82. Su C, Cunningham RL, Rybalchenko N, Singh M. Progesterone increases the release of brainderived
neurotrophic factor from glia via progesterone receptor membrane component 1
(Pgrmc1)-dependent ERK5 signaling. Endocrinology. 2012b;153:4389–4400.
83. Suzuki S, Gerhold LM, Bottner M, Rau SW, Dela Cruz C, Yang E, Zhu H, Yu J, Cashion AB,
Kindy MS, Merchenthaler I, Gage FH, Wise PM. Estradiol enhances neurogenesis following
ischemic stroke through estrogen receptors alpha and beta. J Comp Neurol. 2007;500:1064–
1075.
84. Thomas AJ, Nockels RP, Pan HQ, Shaffrey CI, Chopp M. Progesterone is neuroprotective after
acute experimental spinal cord trauma in rats. Spine. 1999;24:2134–2138.
85. Tokmakov AA, Fukami Y. [Nongenomic mechanisms of progesterone] Tsitologiia. 2009;51:403–
416.
86. Toran-Allerand C. Sex steroids and the development of the newborn mouse hypothalamus and
preoptic area in vitro: implications for sexual differentiation. Brain Res. 1976;106:407–412.
87. Toran-Allerand C. Sex steroids and the development of the newborn mouse hypothalamus and
preoptic area in vitro. II. Morphological correlates and hormonal specificity. Brain Res.
1980;189:413–427.
88. Towle AC, Sze PY. Steroid binding to synaptic plasma membrane: differential binding of
glucocorticoids and gonadal steroids. J Steroid Biochem. 1983;18:135–143.
89. Vandromme M, Melton SM, Kerby JD. Progesterone in traumatic brain injury: time to move on
to phase III trials. Crit Care. 2008;12:153.
90. Vasan R, Vali M, Bell-Horner C, Kaur P, Dillon GH, Singh M. Regulation of the GABA-A
receptor by the MAPK pathway and progesterone. 33rd Annual Society for Neuroscience
Meeting; New Orleans, LA. 2003. pp. 472–412. Vol., ed.^eds.,
91. Vitarbo EA, Chatzipanteli K, Kinoshita K, Truettner JS, Alonso OF, Dietrich WD. Tumor
necrosis factor alpha expression and protein levels after fluid percussion injury in rats: the effect
of injury severity and brain temperature. Neurosurgery. 2004;55:416–424. discussion 424–5.
Wali B, Sayeed I, Stein DG. Improved behavioral outcomes after progesterone administration in
aged male rats with traumatic brain injury. Restor Neurol Neurosci. 2011;29:61–71.
93. Winneker RC, Bitran D, Zhang Z. The preclinical biology of a new potent and selective
progestin: trimegestone. Steroids. 2003;68:915–920.
Wright DW, Kellermann AL, Hertzberg VS, Clark PL, Frankel M, Goldstein FC, Salomone JP,
Dent LL, Harris OA, Ander DS, Lowery DW, Patel MM, Denson DD, Gordon AB, Wald MM,
Gupta S, Hoffman SW, Stein DG. ProTECT: a randomized clinical trial of progesterone for acute
traumatic brain injury. Ann Emerg Med. 2007;49:391–402. e1–e2.
95. Wu WW, Adelman JP, Maylie J. Ovarian hormone deficiency reduces intrinsic excitability and
abolishes acute estrogen sensitivity in hippocampal CA1 pyramidal neurons. J Neurosci.
2011;31:2638–2648.
96. Xiao G, Wei J, Yan W, Wang W, Lu Z. Improved outcomes from the administration of
progesterone for patients with acute severe traumatic brain injury: a randomized controlled trial.
Crit Care. 2008;12:R61.
97. Zandi PP, Carlson MC, Plassman BL, Welsh-Bohmer KA, Mayer LS, Steffens DC, Breitner JC.
Hormone replacement therapy and incidence of Alzheimer disease in older women: the Cache
County Study. Jama. 2002;288:2123–2129.
98. Zhen X, Uryu K, Cai G, Johnson GP, Friedman E. Age-associated impairment in brain MAPK
signal pathways and the effect of caloric restriction in Fischer 344 rats. J Gerontol A Biol Sci
Med Sci. 1999;54:B539–B548.
99. Zhu Y, Bond J, Thomas P. Identification, classification, and partial characterization of genes in
humans and other vertebrates homologous to a fish membrane progestin receptor. Proc Natl Acad
Sci U S A. 2003a;100:2237–2242.
100. Zhu Y, Rice CD, Pang Y, Pace M, Thomas P. Cloning, expression, and characterization of a
membrane progestin receptor and evidence it is an intermediary in meiotic maturation of fish
oocytes. Proc Natl Acad Sci U S A. 2003b;100:2231–2236