Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Oct 1.
Published in final edited form as: Neurochem Res. 2019 Mar 12;44(10):2336–2345. doi: 10.1007/s11064-019-02769-6

Sex differences in ischemia/reperfusion injury: the role of mitochondrial permeability transition

Jasmine A Fels 1,2, Giovanni Manfredi 1,*
PMCID: PMC6742589  NIHMSID: NIHMS1523788  PMID: 30863968

Abstract

Brain and heart ischemia are among the leading causes of death and disability in both men and women, but there are significant sex differences in the incidence and severity of these diseases. Ca2+ dysregulation in response to ischemia/reperfusion injury (I/RI) is a well-recognized pathogenic mechanism leading to the death of affected cells. Excess intracellular Ca2+ causes mitochondrial matrix Ca2+ overload that can result in mitochondrial permeability transition (MPT), which can have severe consequences for mitochondrial function and trigger cell death. Recent findings indicate that estrogens and their related receptors are involved in the regulation of MPT, suggesting that sex differences in I/RI could be linked to estrogen-dependent modulation of mitochondrial Ca2+. Here, we review the evidence supporting sex differences in I/RI and the role of estrogen and estrogen receptors (ERs) in producing these differences, the involvement of mitochondrial Ca2+ overload in disease pathogenesis, and the estrogen-dependent modulation of MPT that may contribute to sex differences.

Keywords: Ischemia, sex, mitochondrial permeability transition, calcium, estrogen, estrogen receptor

1. Sex Differences in Ischemia/Reperfusion Injury

Ischemia, the blockage of blood supply to a part of the body, is a critical and prevalent public health problem. Common forms of ischemia are cerebral ischemic stroke (IS), caused by a blockage in blood flow to the brain, and myocardial ischemia (MI), caused by a blockage in the heart. Sex differences in both cerebral and myocardial ischemia have attracted considerable attention in recent years and are now well documented. Ischemic heart disease is the leading cause of death in both women and men [1]. Stroke, which includes both hemorrhagic and ischemic causes (although IS is far more common [2]) has dropped to the fifth leading cause of death in men, but remains the third in women [3]. There are clear age-related sex differences in both IS and MI incidence, as after age-matching it becomes apparent that women are protected from IS/MI until approximately age 75, when risk reverses and women become significantly more vulnerable [2, 46]. Strikingly, the overall lifetime incidence of stroke is higher in women [7, 8], and 60% of all stroke deaths occur in women [9]. Sex differences exist not only in the incidence of ischemia, but also in its severity. Regardless of age, women have more severe strokes and worse post-stroke outcomes [8, 10]. They are more likely to be disabled by stroke, have higher mortality rates [11], and are 3.5 times more likely to be placed in a nursing home post-stroke than men [3]. Similarly, after acute myocardial infarction, women are more likely to experience independence loss and poorer quality of life [12, 13]. Furthermore, some therapeutic strategies for IS, such as neuronal nitric oxide synthase (nNOS) and poly (ADP-ribose) polymerase (PARP) inhibitors, have been demonstrated pre-clinically to have better efficacy in males than females [3, 6], underscoring the importance of understanding the mechanisms underlying sex differences in IS/MI, so effective therapies can be developed for all patients.

Sex differences in IS/MI incidence and severity could partially derive from environmental factors, such as psychosocial risk factors [13] and a tendency for women to have atypical clinical presentation, resulting in delayed diagnosis and treatment [9]. However, there is a clear biological component, because sex differences are recapitulated in multiple animal models of ischemia/reperfusion injury (I/RI). Adult female mice have smaller brain lesions than males in a model of moderate hypoxia/ischemia [14]. Females are also preferentially protected in other widely used rodent models of cerebral ischemia [15], including the most commonly used middle cerebral artery occlusion (MCAO) model, which produces a lesion in the caudate/putamen and cortex in the hemisphere ipsilateral to occlusion. Interestingly, MCAO produces a significantly larger infarct and poorer functional motor recovery in aged female rats when compared to young females [16], experimentally recapitulating the increased vulnerability of older women to IS. Similar trends can be seen in the heart, as female rat hearts are more resistant than male hearts to damage after oxygen deprivation [1] and tend to have smaller infarcts after myocardial I/RI in the Langendorff preparation [17].

The effects of sex and age on the incidence and outcomes of IS/MI point to a role for sex hormones. Estrogens are a demonstrably important part of the sex differences, as risk for ischemia in women correlates well with lifetime fluctuations in estrogen levels. Women are protected from both cerebral and myocardial ischemia before menopause, but risk increases steeply after menopause or oophorectomy, when estrogen declines [1, 18, 19]. Additionally, women with natural menopause before age 42 have double the risk for IS than women with later menopause [20]. Estrogen replacement therapy (ERT) would seem an effective way to offset this increase in risk. Beneficial effects of ERT have been reported [18], but ERT is still controversial, because the Women’s Health Initiative (WHI) ERT clinical trial was discontinued after an increase in adverse effects, including cardiovascular disease. There is evidence that ERT may be effective if administered before age 50 [19], and post hoc analysis of the WHI data demonstrates that the therapy is advantageous when given to women recently post-menopause, but not to women who had been menopausal for many years [21]. Therefore, it would be premature to completely discount ERT as an option for reducing the risk of ischemia until more research is done to outline the precise conditions affecting its efficacy.

Pre-clinical studies using animal models have directly demonstrated that estrogens are important in the early protection and age-related increase in vulnerability to I/RI in females. The estrogen estradiol (17-beta estradiol, 17βE) is neuroprotective in ovariectomized female rats undergoing transient MCAO [22]. Additionally, hypertensive female rats, prone to spontaneous strokes, are more susceptible to neuronal damage after permanent MCAO in metestrus (low circulating estradiol) than in proestrus (high estradiol), although estradiol levels did not affect ischemic brain damage in another, non-hypertensive strain of rats [23]. Consistent with the re-analysis of the WHI study, estrogens are only neuroprotective in I/RI when given soon after ovariectomy in rats, as they lose all beneficial effects when given after a 10-week period of hypoestrogenism [24]. These data fit well with the estrogen-correlated resistance to IS seen clinically in premenopausal women, but also suggests that estrogen neuroprotection is not straightforward, and more work needs to be done to elucidate its mechanisms. Estrogen plays a role also in MI, as isolated cardiac myocytes from female rats are more resistant to simulated I/RI than male myocytes, but myocytes from older or ovariectomized rats lose this protection [25]. Moreover, 17βE is protective in myocardial I/RI in both female and male rabbit hearts [18, 26], indicating that exogenous estrogen could also be protective in males. These data indicate that animal models of myocardial I/RI can mirror the age- and hormone-related effects observed in patients with MI.

In addition to estradiol, estrogen receptors (ERs) are also involved in producing the sexually dimorphic phenotypes observed in I/RI. There are two canonical ERs, ERα and ERβ, that have both genomic (transcriptional) and non-genomic actions, and a more recently discovered G-protein coupled receptor with a high binding affinity for estrogens termed GPER. Both ERα and ERβ are involved estrogen-mediated neuroprotection. The synthetic ERα - selective agonist PPT and the ERβ -selective agonist WAY 200070-3 both attenuate neuronal death in the CA1 region of the hippocampus in ovariectomized rats after transient global ischemia [27]. Furthermore, 17βE, but not the hormonally inactive 17αE, exerts protective effects in myocardial I/RI of ovariectomized rabbits, demonstrating an ER-dependent mechanism [28]. GPER activation with the agonist G1 is also protective in myocardial I/RI in males [29], and estradiol mediates protection of ischemic male hearts through GPER [30], providing further evidence that estrogens/ERs exert protection in both males and females. Taken together, data from humans and animal models provide strong evidence that estrogen is protective against cerebral and myocardial ischemia, and that the decline of estrogen after menopause puts females at a particularly high risk.

2. Calcium-Dependent Mechanisms of Cell Death in I/RI

Differences in the mechanisms leading to cell death could contribute to the sex differences in the clinical outcomes of I/RI. Cell death in I/RI is a complex process involving multiple pathways. A well-recognized trigger of cell death in I/RI is a pathological rise of intracellular calcium (Ca2+), and evidence shows that estrogens and ERs can modulate Ca2+-dependent ischemic cell death. In neurons, this type of cell death is mainly mediated by glutamate excitotoxicity, which is defined as a glutamate-mediated accumulation of intracellular Ca2+ that produces cell damage [31]. 17βE prevents glutamate excitotoxicity in hippocampal and cortical neurons [32, 33] and reduces lesions following stereotactic perfusion of glutamate into the cerebral cortex of male rats [34]. DPN, an ERβ-selective agonist, is neuroprotective against excitotoxic glutamate in hippocampal neurons [35]. Interestingly, neurons from ERβ knockout animals are protected from glutamate excitotoxicity and oxidative stress [33, 36]. 17βE also prevents intracellular Ca2+ loading after hypoxia/reoxygenation in female cardiomyocytes [37].

During glutamate excitotoxicity Ca2+ enters the cytosol most notably through NMDA-type ionotropic glutamate receptors [38, 39], but also through additional routes, including other ionotropic glutamate receptors and voltage-gated Ca2+ channels (VGCCs), as well as release from endoplasmic reticulum stores induced by metabotropic receptors [40]. In the ischemic heart, a massive accumulation of intracellular Ca2+ occurs through an increase in the reverse activity of the plasma membrane Na+/Ca2+ exchanger, which allows Ca2+ into the cell and extrudes Na+. The reason for this reversal is a steep decline in intracellular pH due to increased glycolysis, which forces the Na+/H+ exchanger to pump H+ ions out of the cell, while allowing large amounts of Na+ in [41]. Non-Na+ dependent pathways of Ca2+ influx, including VGCCs, also appear to play a role in the heart [42]. Because efflux mechanisms depend on the plasma membrane Ca2+-ATPase pump and Na+/Ca2+ exchanger operating in forward mode and sequestration in the ER by the sarco/endoplasmic reticulum Ca2+-ATPase [40], ATP depletion during ischemia precipitates intracellular Ca2+ accumulation [43]. Figure 1 schematically illustrates the main mechanisms of cytosolic Ca2+ regulation in neurons and cardiomyocytes during I/RI.

Fig. 1. Modes of cellular Ca2+ entry and release in I/RI.

Fig. 1.

During ischemia, Ca2+ enters cardiac cells through the plasma membrane Na+/Ca2+ exchanger (NCX) operating in reverse mode, and through voltage-gated Ca2+ channels (VGCCs). In neurons, cytosolic Ca2+ influx occurs mainly through ionotropic NMDA-type glutamate receptors (NMDARs), other ionotropic glutamate receptors (AMPARs, KRs), and VGCCs. Additionally, in both neurons and cardiomyocytes, the endoplasmic reticulum can take up Ca2+ through the sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) and release stored Ca2+ through inositol triphosphate receptors (IP3Rs) and ryanodine receptors (RyRs). In neurons, ER Ca2+ release can be induced by metabotropic glutamate receptors (mGluRs). Intracellular Ca2+ overload occurs because efflux requires the NCX operating in forward mode or energy-dependent plasma membrane Ca2+-ATPases (PMCA), both of which fail in ischemia. Mitochondria are depolarized during ischemia, so Ca2+ import is minimal, but Ca2+ is imported into the matrix upon repolarization during reperfusion. Import occurs through the mitochondrial Ca2+ uniporter (MCU), and efflux occurs through the mitochondrial Na+/Ca2+ exchanger (NCLX) and Ca2+/H+ exchanger (LETM1).

The cytosolic accumulation of Ca2+ during I/RI leads to Ca2+ uptake by mitochondria. Mitochondria import Ca2+ into the matrix through the mitochondrial Ca2+ uniporter (MCU), which is regulated by several associated proteins [44]. Efflux of Ca2+ from the matrix involves both Na+-dependent (through the mitochondrial Na+/Ca2+ exchanger) and Na+-independent (through the Ca2+/H+ exchanger) pathways. Figure 1 illustrates the key players of mitochondrial Ca2+ regulation. The capacity of mitochondria for Ca2+ import exceeds their efflux capacity, leading to mitochondrial accumulation of Ca2+ when cytosolic Ca2+ is high [45]. This Ca2+ accumulation is functionally important in ischemia because it plays a key role in the excitotoxic neuronal death occurring in IS [46, 47] and in the mitochondrial dysfunction that precedes cell death in myocardial I/RI [42]. Blocking mitochondrial Ca2+ uptake largely decreases neuronal death after excitotoxic glutamate exposure [48, 49]. Furthermore, functional, polarized mitochondria are necessary for restoring intracellular Ca2+ homeostasis after excitotoxic stress [50]. Taken together, these data suggest that mitochondrial accumulation of Ca2+ is highly relevant for the Ca2+-mediated death of neuronal and cardiac cells in I/RI and could be an important contributor to sex differences in ischemia.

3. Mitochondrial Permeability Transition and its Role in I/RI

High amounts of matrix Ca2+ cause severe damage to mitochondria, and by extension to the cell as a whole, through a well-studied but still incompletely understood mechanism termed the mitochondrial permeability transition (MPT). Mitochondria rely on a highly impermeable inner mitochondrial membrane (IMM) to generate the proton motive force essential for ATP production. Matrix Ca2+ overload triggers MPT, a process mediated by the opening of a non-specific pore, which causes a sudden disruption in the impermeability of the IMM (Figure 2) [51, 52].

Fig. 2. 17βE and ERβ regulate MPT in I/RI.

Fig. 2

Mitochondria take up cytosolic Ca2+ into the matrix through the MCU and release Ca2+ through NCLX and LETM1. 17βE can affect matrix Ca2+ levels by modulating both MCU and NCLX. In healthy mitochondria, matrix Ca2+ does not reach the threshold for triggering MPT and the MPTP remains closed. Therefore, mitochondria maintain membrane integrity, energy production, and cell viability is preserved. During I/RI, matrix Ca2+ overload causes opening of the MPTP, resulting in MPT and collapse of mitochondrial energy-generating capacity. MPTP opening causes matrix swelling, rupture of mitochondrial membranes, and release of pro-apoptotic factors, such as cytochrome C, resulting in cell death. 17βE and ERβ can modulate MPT under conditions of I/RI by regulating both matrix Ca2+ levels and the MPTP.

The MPT pore (MPTP) has a conductance of approximately 1.5-2 nS, an estimated diameter between 2-3 nm, and can allow the passage of solutes smaller than 1.5 kDa through the IMM [53, 54]. A major consequence of MPT is the collapse of the mitochondrial membrane potential, which results in a loss of ion homeostasis and proton motive force and therefore depletion of ATP generation capacity and subsequent energy failure [54]. Additionally, when the MPTP opens the colloidal osmotic pressure of the proteins in the matrix results in a large influx of water, causing swelling which can burst mitochondrial membranes, facilitate the release of cytochrome c, and initiate apoptosis (Figure 2) [54, 55]. The main inducer for MPTP opening is high matrix Ca2+ [56], but other factors including matrix depolarization, neutral matrix pH, oxidative stress, inorganic phosphate, and de-energization can lower the Ca2+ threshold and increase MPTP opening probability [57, 58]. Factors decreasing MPTP opening probability are acidic matrix pH, adenine nucleotides, and magnesium [59]. The regulation of MPT is complex, because it is cell-type and context-specific and multiple interacting factors mediate MPTP opening. For example, matrix pH, inorganic phosphate levels, and mitochondrial energization state are highly interconnected factors that regulate MPT [54, 56, 58]. These are particularly relevant in the context of I/RI, because there are extensive and rapid changes in mitochondrial energization state and matrix pH in the different stages of ischemia and reperfusion.

The structural makeup of the MPTP remains controversial. The adenine nucleotide translocator ANT was originally believed to be a structural component of the pore [60], but genetic loss of function experiments ruled this out [61]. However, there is evidence that ANT may play a regulatory role in the Ca2+ threshold for pore opening, particularly in response to matrix depolarization and oxidative stress [54, 62]. The c subunit of the F1 region of the F0/F1 mammalian ATP synthase has been suggested as a pore forming component [63, 64]; however, simulations showed that the c-ring is thermodynamically unfavorable to occupation by water and therefore would be unlikely to form a pore [65]. The interface between F0/F1 ATP synthase dimers has also been suggested to form the MPTP [66, 67], but MPT can occur in cells with all three genes encoding subunit c of the ATP synthase knocked out [68] and in cells lacking either subunit b or the oligomycin-sensitivity conferring protein (OSCP) subunit of the ATP synthase [69]. Nevertheless, it is possible that multiple subunits would need to be knocked out simultaneously to fully abolish MPT.

While the proteins forming the pore itself remain unknown, CypD is a widely accepted pore regulator. CypD is a peptidyl-prolyl isomerase localized to the mitochondrial matrix [70]. CypD-deficient mitochondria are resistant to Ca2+-induced MPT, identifying it as a Ca2+ sensor for the pore [7173]. Cyclosporin A (CSA), a fungal-derived compound, is the most widely used inhibitor of MPT, and modulates pore opening by inhibiting CypD [74, 75]. However, CSA also inhibits all other cyclophilins, leading to powerful immunosuppressive and potentially neuromodulatory effects through its inhibition of calcineurin; so, results obtained with CSA must be interpreted with caution [70, 76]. CypD has been shown to interact with the OSCP subunit of the ATP synthase in an inorganic phosphate-stimulated and CSA-sensitive manner [77, 78]. This interaction lowers ATP generation capacity by uncoupling the ATP synthase from the electron transport chain and lowering its catalytic activity [79]. Loss of OSCP causes a CypD-independent increase in MPTP opening [80], suggesting that if the ATP synthase is a component of the pore, MPTP modulation by CypD could be dependent on its interaction with OSCP.

While MPT was traditionally thought of as an all-or-nothing event occurring exclusively in pathological situations, a large body of evidence suggests that the MPTP can also have reversible low-conductance states, which may play a physiological role. Low-conductance states (approximately 0.5 nS) with permeability to solutes under 300 Da have been measured in isolated mitochondria and in cells [75, 8185] and in vivo in astrocytes [86]. Additionally, low conductance reversible opening of the MPTP in cardiac mitochondria is protective against Ca2+ overload. [87]. Mitochondria have limited Ca2+ efflux mechanisms, which can easily become overloaded during periods of high Ca2+ entry into the cell. Transient MPTP opening could provide a fast and thermodynamically favorable route for mitochondrial Ca2+ efflux, although its role as a Ca2+ release channel is controversial [45, 88, 89]. MPT is also a variable and highly heterogeneous phenomenon across the mitochondrial pool, as not all mitochondria in a cell will undergo the transition simultaneously or at all [90]. Furthermore, some populations of mitochondria appear to be more vulnerable to MPT. For example, neuronal synaptic mitochondria have a lower Ca2+ threshold than non-synaptic mitochondria [91, 92]. This heterogeneity could be explained in part by variations in the expression of CypD, whose levels have been shown to differ based on subcellular localization, tissue, and age [77, 92, 93].

MPT has been demonstrated to play a crucial role in cell death in a variety of pathological conditions, but most notably in I/RI [94, 95]. Although reperfusion is necessary to limit cell death due to hypoxia during ischemia, it can also promote MPT. While mitochondria are depolarized during ischemia, Ca2+ accumulates in the cytosol. After reoxygenation mitochondria reestablish their membrane potential, allowing for massive Ca2+ entry into the matrix. Ca2+ influx depolarizes the matrix, which, together with the depletion of adenine nucleotides, a burst of reactive oxygen species, and reestablishment of matrix pH, primes mitochondria for MPT [96].

There is convincing evidence that blocking CypD-sensitive MPT is protective in models of ischemia. CSA treatment significantly decreases infarct size after transient MCAO [97] or global brain ischemia [98], and CypD knockout dramatically reduces infarct size in MCAO [71]. Inhibiting MPT at the time of reperfusion with the non-immunosuppressive CypD inhibitor NIM811 protects the heart against I/RI [99], and CypD knockout reduces infarct size by up to 40% [100, 101]. Furthermore, mitochondria isolated from CypD knockout animals are resistant to swelling induced by high Ca2+, and CypD overexpression in the heart increases the probability of mitochondrial swelling, cardiac cell death, and cardiac hypertrophy [101]. Interestingly, CypD deficient hepatocytes and embryonic fibroblasts are protected from necrosis but not against several apoptotic stimuli [71, 100, 102]. Taken together, this evidence suggests that CypD-regulated MPT is an important contributor to necrotic, but not apoptotic, cell death in I/RI. However, a role for MPT not mediated by or insensitive to CypD in apoptotic cell death cannot be ruled out.

4. Sex Differences in Mitochondrial Ca2+ Handling and Permeability Transition

The regulation of protection by estrogens and ERs that is observed in I/RI must be mediated at least in part through their maintenance of mitochondrial function, as the neuroprotective ability of estrogenic compounds correlates highly with their ability to maintain mitochondrial membrane potential after insult [103]. Some studies have addressed this issue directly. For example, it was shown that mitochondria from ERβ knockout animals repolarize and maintain ATP production more effectively than wild type controls after oxidative insult [36]. In addition, estradiol and ER ligands can directly modulate mitochondrial Ca2+ uptake and release dynamics. 17βE, DPN, and ERα-selective agonist PPT increase Ca2+ uptake into the matrix through the MCU in a receptor-independent fashion, while tamoxifen, an ER antagonist, inhibits it [104]. In rat synaptosomal mitochondria, 17βE at physiological concentrations decreases Na+-dependent Ca2+ efflux, but at higher concentrations increases it [105]. Estradiol can modulate the ability of mitochondria to tolerate high Ca2+ loads, as it was shown to increase the levels of Bcl-2 protein [106]. 17βE also prevents age-related Ca2+ dysregulation in neurons from male rats by preserving the ability of mitochondria to accumulate Ca2+ [32]. Altogether, this suggests that males and pre- or post-menopausal females may possess distinct dynamics of mitochondrial calcium homeostasis, which could contribute to sex differences in I/RI.

Importantly, sex differences have been demonstrated in MPT. Female mitochondria from both brain and spinal cord have lower Ca2+ capacity, a measure of susceptibility to MPT, and greater Ca2+ induced depolarization. Genetic ablation of CypD abolishes these differences, suggesting that they are dependent on MPT [33, 107]. Additionally, treating both female and male brain mitochondria with a high concentration of 17βE decreases their Ca2+ capacity, and this was prevented by CypD knockout, pointing to a direct role for 17βE in modulating MPT [107]. These results demonstrate that female brain mitochondria are more prone to MPT than male brain mitochondria. Further, mitochondria isolated from female rat heart have higher Ca2+ capacity than male mitochondria [108], but also take up less Ca2+ initially, have reduced swelling in response to high matrix Ca2+, and recover membrane potential more quickly after Ca2+-induced depolarization [109, 110]. These results suggest that female cardiac mitochondria are less susceptible to MPT than male mitochondria, while female brain mitochondria are more susceptible, suggesting that sex-dependent regulation of MPT is tissue-specific.

In addition to estradiol, ERβ is an important mediator of the susceptibility of female brain mitochondria to MPT and could potentially modulate MPT sensitivity in other tissues as well. We have shown that knockout of ERβ decreases the sex difference in brain mitochondrial Ca2+ capacity, and that this effect depends on CypD. ERβ localizes in various cell compartments, including the mitochondrial matrix [111], suggesting that the mitochondrial pool of ERβ could be responsible for modulating MPT. We showed that ERβ knockout decreases the interaction between CypD and OSCP, a proposed MPTP component, while 17βE increases it, likely by acting as a ligand for ERβ and modulating its protein-protein interactions [33]. In addition, 17βE binds to OSCP directly [112]. At pharmacological concentrations 17βE promotes the “slip” rate of the ATP synthase by decreasing the efficiency of the coupling between proton flow and ADP phosphorylation [113]. Therefore, modulation of the CypD-OSCP interaction by 17βE and/or ERβ could provide the mechanism for the estrogen-dependent sex differences in MPT.

5. Conclusions and Future Perspectives

IS/MI are common causes of disability and death that pose a severe burden for public health worldwide, and this will only increase as the population continues to age. It is clear that sex differences exist in ischemic injury, due to the modulatory effects of estrogen and ERs. Understanding the basis for these differences would provide insight into fundamental mechanisms of cell death in I/RI that can be targeted to develop therapeutics effective in both sexes. Ca2+-dependent cell death is a crucial part of the injury, in both brain and heart. One of the main mechanisms mediating this type of cell death is mitochondrial Ca2+ overload and MPT. Estrogen/ER-dependent sex differences in mitochondrial Ca2+ handling and MPT are likely linked to the sexual dimorphism seen in IS/MI. Based on studies by our group and others, we propose that sex differences in MPT are related to the modulation of the interactions between CypD and MPTP components by 17βE and ERβ. Figure 2 schematically summarizes the processes leading to mitochondrial Ca2+ overload and MPT in I/RI, the downstream consequences of MPT, and the proposed modulatory role of estrogens and ERβ on mitochondrial Ca2+ dynamics.

In the future, it will be important to better understand the extent to which estrogens and ERs modulate the sexual dimorphism in ischemia. Estrogens and ERs are undoubtedly involved, but sex differences could also result from other hormones including androgens, as well as genetic and epigenetic dissimilarities. More work needs to be done to directly compare in both sexes the influence of estrogens and ERs on mitochondrial function, Ca2+ handling, and MPT, in physiological and pathological conditions. More research is also necessary to examine the mechanisms by which estrogens and ERs regulate MPTP opening. Identifying the targets of 17βE/ERβ in modulating MPT will provide more clarity about the true components of the pore, as well as new approaches to regulate MPTP opening and promote cell survival in I/RI. Additionally, it is still unclear whether the sex differences in MPT reflect variations in irreversible high conductance MPTP opening, reversible low conductance opening, or both. Reversible MPTP opening could potentially be a protective mechanism against mitochondrial Ca2+ overload by allowing limited ionic efflux. Whether reversible MPTP opening occurs differentially in I/RI in males and females, and if it is modulated by estrogens and ERs, remain to be elucidated. This will be important when considering targeting the MPTP as a therapeutic strategy in a sex-dependent manner. Finally, it will be important to investigate whether sex differences in MPT play a role in other pathologies besides ischemia. Many neurodegenerative diseases have significant differences in occurrence and progression in men and women, including Alzheimer’s disease, amyotrophic lateral sclerosis, and multiple sclerosis, and early evidence indicates that MPT could be an important contributor to pathogenesis in these disorders [107, 114, 115]. However, whether differences in MPT and its modulation by estrogen and ERs contribute to the sex differences in these diseases is yet to be determined.

Acknowledgments

This work was supported by NIH/NINDS grant 1R01NS095692.

References

  • 1.Ostadal B, Ostadal P (2014) Sex-based differences in cardiac ischaemic injury and protection: therapeutic implications. Br J Pharmacol 171:541–554 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Gibson CL, Attwood L (2016) The impact of gender on stroke pathology and treatment. Neurosci Biobehav Rev 67:119–124 [DOI] [PubMed] [Google Scholar]
  • 3.Spychala MS, Honarpisheh P, McCullough LD (2017) Sex differences in neuroinflammation and neuroprotection in ischemic stroke. J Neurosci Res 95:462–471 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Herson PS, Palmateer J, Hurn PD (2013) Biological sex and mechanisms of ischemic brain injury. Transl Stroke Res 4:413–419 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Haast RA, Gustafson DR, Kiliaan AJ (2012) Sex differences in stroke. J Cereb Blood Flow Metab 32:2100–2107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Ahnstedt H, McCullough LD, Cipolla MJ (2016) The Importance of Considering Sex Differences in Translational Stroke Research. Transl Stroke Res 7:261–273 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Gibson CL (2013) Cerebral ischemic stroke: is gender important? J Cereb Blood Flow Metab 33:1355–1361 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Chauhan A, Moser H, McCullough LD (2017) Sex differences in ischaemic stroke: potential cellular mechanisms. Clin Sci (Lond) 131:533–552 [DOI] [PubMed] [Google Scholar]
  • 9.Girijala RL, Sohrabji F, Bush RL (2017) Sex differences in stroke: Review of current knowledge and evidence. Vasc Med 22:135–145 [DOI] [PubMed] [Google Scholar]
  • 10.Appelros P, Stegmayr B, Terent A (2009) Sex differences in stroke epidemiology: a systematic review. Stroke 40:1082–1090 [DOI] [PubMed] [Google Scholar]
  • 11.Di Carlo A, Lamassa M, Baldereschi M, Pracucci G, Basile AM, Wolfe CD, Giroud M, Rudd A, Ghetti A, Inzitari D, European BSoSCG (2003) Sex differences in the clinical presentation, resource use, and 3-month outcome of acute stroke in Europe: data from a multicenter multinational hospital-based registry. Stroke 34:1114–1119 [DOI] [PubMed] [Google Scholar]
  • 12.Dodson JA, Arnold SV, Reid KJ, Gill TM, Rich MW, Masoudi FA, Spertus JA, Krumholz HM, Alexander KP (2012) Physical function and independence 1 year after myocardial infarction: observations from the Translational Research Investigating Underlying disparities in recovery from acute Myocardial infarction: Patients’ Health status registry. Am Heart J 163:790–796 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Humphries KH, Izadnegahdar M, Sedlak T, Saw J, Johnston N, Schenck-Gustafsson K, Shah RU, Regitz-Zagrosek V, Grewal J, Vaccarino V, Wei J, Bairey Merz CN (2017) Sex differences in cardiovascular disease - Impact on care and outcomes. Front Neuroendocrinol 46:46–70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Zhu C, Xu F, Wang X, Shibata M, Uchiyama Y, Blomgren K, Hagberg H (2006) Different apoptotic mechanisms are activated in male and female brains after neonatal hypoxia-ischaemia. J Neurochem 96:1016–1027 [DOI] [PubMed] [Google Scholar]
  • 15.Hurn PD, Vannucci SJ, Hagberg H (2005) Adult or perinatal brain injury: does sex matter? Stroke 36:193–195 [DOI] [PubMed] [Google Scholar]
  • 16.DiNapoli VA, Huber JD, Houser K, Li X, Rosen CL (2008) Early disruptions of the blood-brain barrier may contribute to exacerbated neuronal damage and prolonged functional recovery following stroke in aged rats. Neurobiol Aging 29:753–764 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Johnson MS, Moore RL, Brown DA (2006) Sex differences in myocardial infarct size are abolished by sarcolemmal KATP channel blockade in rat. Am J Physiol Heart Circ Physiol 290:H2644–2647 [DOI] [PubMed] [Google Scholar]
  • 18.Booth EA, Lucchesi BR (2008) Estrogen-mediated protection in myocardial ischemia-reperfusion injury. Cardiovasc Toxicol 8:101–113 [DOI] [PubMed] [Google Scholar]
  • 19.Rocca WA, Grossardt BR, Miller VM, Shuster LT, Brown RD Jr. (2012) Premature menopause or early menopause and risk of ischemic stroke. Menopause 19:272–277 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Lisabeth LD, Beiser AS, Brown DL, Murabito JM, Kelly-Hayes M, Wolf PA (2009) Age at natural menopause and risk of ischemic stroke: the Framingham heart study. Stroke 40:1044–1049 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Harman SM, Naftolin F, Brinton EA, Judelson DR (2005) Is the estrogen controversy over? Deconstructing the Women’s Health Initiative study: a critical evaluation of the evidence. Ann N Y Acad Sci 1052:43–56 [DOI] [PubMed] [Google Scholar]
  • 22.Simpkins JW, Rajakumar G, Zhang YQ, Simpkins CE, Greenwald D, Yu CJ, Bodor N, Day AL (1997) Estrogens may reduce mortality and ischemic damage caused by middle cerebral artery occlusion in the female rat. J Neurosurg 87:724–730 [DOI] [PubMed] [Google Scholar]
  • 23.Carswell HV, Dominiczak AF, Macrae IM (2000) Estrogen status affects sensitivity to focal cerebral ischemia in stroke-prone spontaneously hypertensive rats. Am J Physiol Heart Circ Physiol 278:H290–294 [DOI] [PubMed] [Google Scholar]
  • 24.Suzuki S, Brown CM, Wise PM (2009) Neuroprotective effects of estrogens following ischemic stroke. Front Neuroendocrinol 30:201–211 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ross JL, Howlett SE (2012) Age and ovariectomy abolish beneficial effects of female sex on rat ventricular myocytes exposed to simulated ischemia and reperfusion. PLoS One 7:e38425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Hale SL, Birnbaum Y, Kloner RA (1997) Estradiol, Administered Acutely, Protects Ischemic Myocardium in Both Female and Male Rabbits. J Cardiovasc Pharmacol Ther 2:47–52 [DOI] [PubMed] [Google Scholar]
  • 27.Miller NR, Jover T, Cohen HW, Zukin RS, Etgen AM (2005) Estrogen can act via estrogen receptor alpha and beta to protect hippocampal neurons against global ischemia-induced cell death. Endocrinology 146:3070–3079 [DOI] [PubMed] [Google Scholar]
  • 28.Booth EA, Marchesi M, Kilbourne EJ, Lucchesi BR (2003) 17Beta-estradiol as a receptor-mediated cardioprotective agent. J Pharmacol Exp Ther 307:395–401 [DOI] [PubMed] [Google Scholar]
  • 29.Bopassa JC, Eghbali M, Toro L, Stefani E (2010) A novel estrogen receptor GPER inhibits mitochondria permeability transition pore opening and protects the heart against ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol 298:H16–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Kabir ME, Singh H, Lu R, Olde B, Leeb-Lundberg LM, Bopassa JC (2015) G Protein-Coupled Estrogen Receptor 1 Mediates Acute Estrogen-Induced Cardioprotection via MEK/ERK/GSK-3beta Pathway after Ischemia/Reperfusion. PLoS One 10:e0135988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Choi DW (1985) Glutamate neurotoxicity in cortical cell culture is calcium dependent. Neurosci Lett 58:293–297 [DOI] [PubMed] [Google Scholar]
  • 32.Brewer GJ, Reichensperger JD, Brinton RD (2006) Prevention of age-related dysregulation of calcium dynamics by estrogen in neurons. Neurobiol Aging 27:306–317 [DOI] [PubMed] [Google Scholar]
  • 33.Burstein SR, Kim HJ, Fels JA, Qian L, Zhang S, Zhou P, Starkov AA, Iadecola C, Manfredi G (2018) Estrogen receptor beta modulates permeability transition in brain mitochondria. Biochim Biophys Acta Bioenerg 1859:423–433 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Mendelowitsch A, Ritz MF, Ros J, Langemann H, Gratzl O (2001) 17beta-Estradiol reduces cortical lesion size in the glutamate excitotoxicity model by enhancing extracellular lactate: a new neuroprotective pathway. Brain Res 901:230–236 [DOI] [PubMed] [Google Scholar]
  • 35.Zhao L, Brinton RD (2007) Estrogen receptor alpha and beta differentially regulate intracellular Ca(2+) dynamics leading to ERK phosphorylation and estrogen neuroprotection in hippocampal neurons. Brain Res 1172:48–59 [DOI] [PubMed] [Google Scholar]
  • 36.Yang SH, Sarkar SN, Liu R, Perez EJ, Wang X, Wen Y, Yan LJ, Simpkins JW (2009) Estrogen receptor beta as a mitochondrial vulnerability factor. J Biol Chem 284:9540–9548 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Jovanovic S, Jovanovic A, Shen WK, Terzic A (2000) Low concentrations of 17beta-estradiol protect single cardiac cells against metabolic stress-induced Ca2+ loading. J Am Coll Cardiol 36:948–952 [DOI] [PubMed] [Google Scholar]
  • 38.Tymianski M, Charlton MP, Carlen PL, Tator CH (1993) Source specificity of early calcium neurotoxicity in cultured embryonic spinal neurons. J Neurosci 13:2085–2104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Peng TI, Greenamyre JT (1998) Privileged access to mitochondria of calcium influx through N-methyl-D-aspartate receptors. Mol Pharmacol 53:974–980 [PubMed] [Google Scholar]
  • 40.Cross JL, Meloni BP, Bakker AJ, Lee S, Knuckey NW (2010) Modes of Neuronal Calcium Entry and Homeostasis following Cerebral Ischemia. Stroke Res Treat 2010:316862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Garcia-Dorado D, Ruiz-Meana M, Inserte J, Rodriguez-Sinovas A, Piper HM (2012) Calcium-mediated cell death during myocardial reperfusion. Cardiovasc Res 94:168–180 [DOI] [PubMed] [Google Scholar]
  • 42.Nayler WG (1981) The role of calcium in the ischemic myocardium. Am J Pathol 102:262–270 [PMC free article] [PubMed] [Google Scholar]
  • 43.Chiong M, Wang ZV, Pedrozo Z, Cao DJ, Troncoso R, Ibacache M, Criollo A, Nemchenko A, Hill JA, Lavandero S (2011) Cardiomyocyte death: mechanisms and translational implications. Cell Death Dis 2:e244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Pallafacchina G, Zanin S, Rizzuto R (2018) Recent advances in the molecular mechanism of mitochondrial calcium uptake. F1000Res 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Bernardi P, von Stockum S (2012) The permeability transition pore as a Ca(2+) release channel: new answers to an old question. Cell Calcium 52:22–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Duchen MR (2012) Mitochondria, calcium-dependent neuronal death and neurodegenerative disease. Pflugers Arch 464:111–121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Gouriou Y, Demaurex N, Bijlenga P, De Marchi U (2011) Mitochondrial calcium handling during ischemia-induced cell death in neurons. Biochimie 93:2060–2067 [DOI] [PubMed] [Google Scholar]
  • 48.Stout AK, Raphael HM, Kanterewicz BI, Klann E, Reynolds IJ (1998) Glutamate-induced neuron death requires mitochondrial calcium uptake. Nat Neurosci 1:366–373 [DOI] [PubMed] [Google Scholar]
  • 49.Castilho RF, Hansson O, Ward MW, Budd SL, Nicholls DG (1998) Mitochondrial control of acute glutamate excitotoxicity in cultured cerebellar granule cells. J Neurosci 18:10277–10286 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Vergun O, Keelan J, Khodorov BI, Duchen MR (1999) Glutamate-induced mitochondrial depolarisation and perturbation of calcium homeostasis in cultured rat hippocampal neurones. J Physiol 519 Pt 2:451–466 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hunter DR, Haworth RA (1979) The Ca2+-induced membrane transition in mitochondria. III. Transitional Ca2+ release. Arch Biochem Biophys 195:468–477 [DOI] [PubMed] [Google Scholar]
  • 52.Szabo I, Zoratti M (1992) The mitochondrial megachannel is the permeability transition pore. J Bioenerg Biomembr 24:111–117 [DOI] [PubMed] [Google Scholar]
  • 53.Crompton M (1999) The mitochondrial permeability transition pore and its role in cell death. Biochem J 341 (Pt 2):233–249 [PMC free article] [PubMed] [Google Scholar]
  • 54.Halestrap AP, Richardson AP (2015) The mitochondrial permeability transition: a current perspective on its identity and role in ischaemia/reperfusion injury. J Mol Cell Cardiol 78:129–141 [DOI] [PubMed] [Google Scholar]
  • 55.Petronilli V, Penzo D, Scorrano L, Bernardi P, Di Lisa F (2001) The mitochondrial permeability transition, release of cytochrome c and cell death. Correlation with the duration of pore openings in situ. J Biol Chem 276:12030–12034 [DOI] [PubMed] [Google Scholar]
  • 56.Giorgio V, Guo L, Bassot C, Petronilli V, Bernardi P (2018) Calcium and regulation of the mitochondrial permeability transition. Cell Calcium 70:56–63 [DOI] [PubMed] [Google Scholar]
  • 57.Petronilli V, Cola C, Bernardi P (1993) Modulation of the mitochondrial cyclosporin A-sensitive permeability transition pore. II. The minimal requirements for pore induction underscore a key role for transmembrane electrical potential, matrix pH, and matrix Ca2+. J Biol Chem 268:1011–1016 [PubMed] [Google Scholar]
  • 58.Doczi J, Turiak L, Vajda S, Mandi M, Torocsik B, Gerencser AA, Kiss G, Konrad C, Adam-Vizi V, Chinopoulos C (2011) Complex contribution of cyclophilin D to Ca2+-induced permeability transition in brain mitochondria, with relation to the bioenergetics state. J Biol Chem 286:6345–6353 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Zoratti M, Szabo I (1995) The mitochondrial permeability transition. Biochim Biophys Acta 1241:139–176 [DOI] [PubMed] [Google Scholar]
  • 60.Crompton M, Virji S, Ward JM (1998) Cyclophilin-D binds strongly to complexes of the voltage-dependent anion channel and the adenine nucleotide translocase to form the permeability transition pore. Eur J Biochem 258:729–735 [DOI] [PubMed] [Google Scholar]
  • 61.Kokoszka JE, Waymire KG, Levy SE, Sligh JE, Cai J, Jones DP, MacGregor GR, Wallace DC (2004) The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature 427:461–465 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Doczi J, Torocsik B, Echaniz-Laguna A, Mousson de Camaret B, Starkov A, Starkova N, Gal A, Molnar MJ, Kawamata H, Manfredi G, Adam-Vizi V, Chinopoulos C (2016) Alterations in voltage-sensing of the mitochondrial permeability transition pore in ANT1-deficient cells. Sci Rep 6:26700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Alavian KN, Beutner G, Lazrove E, Sacchetti S, Park HA, Licznerski P, Li H, Nabili P, Hockensmith K, Graham M, Porter GA Jr., Jonas EA (2014) An uncoupling channel within the c-subunit ring of the F1FO ATP synthase is the mitochondrial permeability transition pore. Proc Natl Acad Sci U S A 111:10580–10585 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Bonora M, Morganti C, Morciano G, Pedriali G, Lebiedzinska-Arciszewska M, Aquila G, Giorgi C, Rizzo P, Campo G, Ferrari R, Kroemer G, Wieckowski MR, Galluzzi L, Pinton P (2017) Mitochondrial permeability transition involves dissociation of F1FO ATP synthase dimers and C-ring conformation. EMBO Rep 18:1077–1089 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Zhou W, Marinelli F, Nief C, Faraldo-Gomez JD (2017) Atomistic simulations indicate the c-subunit ring of the F1Fo ATP synthase is not the mitochondrial permeability transition pore. Elife 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M, Glick GD, Petronilli V, Zoratti M, Szabo I, Lippe G, Bernardi P (2013) Dimers of mitochondrial ATP synthase form the permeability transition pore. Proc Natl Acad Sci U S A 110:5887–5892 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Bernardi P, Di Lisa F, Fogolari F, Lippe G (2015) From ATP to PTP and Back: A Dual Function for the Mitochondrial ATP Synthase. Circ Res 116:1850–1862 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.He J, Ford HC, Carroll J, Ding S, Fearnley IM, Walker JE (2017) Persistence of the mitochondrial permeability transition in the absence of subunit c of human ATP synthase. Proc Natl Acad Sci U S A 114:3409–3414 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.He J, Carroll J, Ding S, Fearnley IM, Walker JE (2017) Permeability transition in human mitochondria persists in the absence of peripheral stalk subunits of ATP synthase. Proc Natl Acad Sci U S A 114:9086–9091 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Giorgio V, Soriano ME, Basso E, Bisetto E, Lippe G, Forte MA, Bernardi P (2010) Cyclophilin D in mitochondrial pathophysiology. Biochim Biophys Acta 1797:1113–1118 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Schinzel AC, Takeuchi O, Huang Z, Fisher JK, Zhou Z, Rubens J, Hetz C, Danial NN, Moskowitz MA, Korsmeyer SJ (2005) Cyclophilin D is a component of mitochondrial permeability transition and mediates neuronal cell death after focal cerebral ischemia. Proc Natl Acad Sci U S A 102:12005–12010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Basso E, Fante L, Fowlkes J, Petronilli V, Forte MA, Bernardi P (2005) Properties of the permeability transition pore in mitochondria devoid of Cyclophilin D. J Biol Chem 280:18558–18561 [DOI] [PubMed] [Google Scholar]
  • 73.Izzo V, Bravo-San Pedro JM, Sica V, Kroemer G, Galluzzi L (2016) Mitochondrial Permeability Transition: New Findings and Persisting Uncertainties. Trends Cell Biol 26:655–667 [DOI] [PubMed] [Google Scholar]
  • 74.Crompton M, Ellinger H, Costi A (1988) Inhibition by cyclosporin A of a Ca2+-dependent pore in heart mitochondria activated by inorganic phosphate and oxidative stress. Biochem J 255:357–360 [PMC free article] [PubMed] [Google Scholar]
  • 75.Hansson MJ, Morota S, Chen L, Matsuyama N, Suzuki Y, Nakajima S, Tanoue T, Omi A, Shibasaki F, Shimazu M, Ikeda Y, Uchino H, Elmer E (2011) Cyclophilin D-sensitive mitochondrial permeability transition in adult human brain and liver mitochondria. J Neurotrauma 28:143–153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Diaz-Ruiz A, Vergara P, Perez-Severiano F, Segovia J, Guizar-Sahagun G, Ibarra A, Rios C (2005) Cyclosporin-A inhibits constitutive nitric oxide synthase activity and neuronal and endothelial nitric oxide synthase expressions after spinal cord injury in rats. Neurochem Res 30:245–251 [DOI] [PubMed] [Google Scholar]
  • 77.Gauba E, Guo L, Du H (2017) Cyclophilin D Promotes Brain Mitochondrial F1FO ATP Synthase Dysfunction in Aging Mice. J Alzheimers Dis 55:1351–1362 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Giorgio V, Bisetto E, Soriano ME, Dabbeni-Sala F, Basso E, Petronilli V, Forte MA, Bernardi P, Lippe G (2009) Cyclophilin D modulates mitochondrial F0F1-ATP synthase by interacting with the lateral stalk of the complex. J Biol Chem 284:33982–33988 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Chinopoulos C, Konrad C, Kiss G, Metelkin E, Torocsik B, Zhang SF, Starkov AA (2011) Modulation of F0F1-ATP synthase activity by cyclophilin D regulates matrix adenine nucleotide levels. FEBS J 278:1112–1125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Beck SJ, Guo L, Phensy A, Tian J, Wang L, Tandon N, Gauba E, Lu L, Pascual JM, Kroener S, Du H (2016) Deregulation of mitochondrial F1FO-ATP synthase via OSCP in Alzheimer’s disease. Nat Commun 7:11483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Bernardi P, Petronilli V (1996) The permeability transition pore as a mitochondrial calcium release channel: a critical appraisal. J Bioenerg Biomembr 28:131–138 [DOI] [PubMed] [Google Scholar]
  • 82.Ichas F, Jouaville LS, Mazat JP (1997) Mitochondria are excitable organelles capable of generating and conveying electrical and calcium signals. Cell 89:1145–1153 [DOI] [PubMed] [Google Scholar]
  • 83.Ichas F, Mazat JP (1998) From calcium signaling to cell death: two conformations for he mitochondrial permeability transition pore. Switching from low- to high-conductance state. Biochim Biophys Acta 1366:33–50 [DOI] [PubMed] [Google Scholar]
  • 84.Huser J, Blatter LA (1999) Fluctuations in mitochondrial membrane potential caused by repetitive gating of the permeability transition pore. Biochem J 343 Pt 2:311–317 [PMC free article] [PubMed] [Google Scholar]
  • 85.Wang W, Fang H, Groom L, Cheng A, Zhang W, Liu J, Wang X, Li K, Han P, Zheng M, Yin J, Wang W, Mattson MP, Kao JP, Lakatta EG, Sheu SS, Ouyang K, Chen J, Dirksen RT, Cheng H (2008) Superoxide flashes in single mitochondria. Cell 134:279–290 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Agarwal A, Wu PH, Hughes EG, Fukaya M, Tischfield MA, Langseth AJ, Wirtz D, Bergles DE (2017) Transient Opening of the Mitochondrial Permeability Transition Pore Induces Microdomain Calcium Transients in Astrocyte Processes. Neuron 93:587–605 e587 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Korge P, Yang L, Yang JH, Wang Y, Qu Z, Weiss JN (2011) Protective role of transient pore openings in calcium handling by cardiac mitochondria. J Biol Chem 286:34851–34857 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Altschuld RA, Hohl CM, Castillo LC, Garleb AA, Starling RC, Brierley GP (1992) Cyclosporin inhibits mitochondrial calcium efflux in isolated adult rat ventricular cardiomyocytes. Am J Physiol 262:H1699–1704 [DOI] [PubMed] [Google Scholar]
  • 89.De Marchi E, Bonora M, Giorgi C, Pinton P (2014) The mitochondrial permeability transition pore is a dispensable element for mitochondrial calcium efflux. Cell Calcium 56:1–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Lemasters JJ, Nieminen AL, Qian T, Trost LC, Elmore SP, Nishimura Y, Crowe RA, Cascio WE, Bradham CA, Brenner DA, Herman B (1998) The mitochondrial permeability transition in cell death: a common mechanism in necrosis, apoptosis and autophagy. Biochim Biophys Acta 1366:177–196 [DOI] [PubMed] [Google Scholar]
  • 91.Brown MR, Sullivan PG, Geddes JW (2006) Synaptic mitochondria are more susceptible to Ca2+ overload than nonsynaptic mitochondria. J Biol Chem 281:11658–11668 [DOI] [PubMed] [Google Scholar]
  • 92.Naga KK, Sullivan PG, Geddes JW (2007) High cyclophilin D content of synaptic mitochondria results in increased vulnerability to permeability transition. J Neurosci 27:7469–7475 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Porter GA Jr., Beutner G (2018) Cyclophilin D, Somehow a Master Regulator of Mitochondrial Function. Biomolecules 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Lemasters JJ, Theruvath TP, Zhong Z, Nieminen AL (2009) Mitochondrial calcium and the permeability transition in cell death. Biochim Biophys Acta 1787:1395–1401 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Rasola A, Bernardi P (2007) The mitochondrial permeability transition pore and its involvement in cell death and in disease pathogenesis. Apoptosis 12:815–833 [DOI] [PubMed] [Google Scholar]
  • 96.Starkov AA, Chinopoulos C, Fiskum G (2004) Mitochondrial calcium and oxidative stress as mediators of ischemic brain injury. Cell Calcium 36:257–264 [DOI] [PubMed] [Google Scholar]
  • 97.Shiga Y, Onodera H, Matsuo Y, Kogure K (1992) Cyclosporin A protects against ischemia-reperfusion injury in the brain. Brain Res 595:145–148 [DOI] [PubMed] [Google Scholar]
  • 98.Uchino H, Elmer E, Uchino K, Lindvall O, Siesjo BK (1995) Cyclosporin A dramatically ameliorates CA1 hippocampal damage following transient forebrain ischaemia in the rat. Acta Physiol Scand 155:469–471 [DOI] [PubMed] [Google Scholar]
  • 99.Argaud L, Gateau-Roesch O, Muntean D, Chalabreysse L, Loufouat J, Robert D, Ovize M (2005) Specific inhibition of the mitochondrial permeability transition prevents lethal reperfusion injury. J Mol Cell Cardiol 38:367–374 [DOI] [PubMed] [Google Scholar]
  • 100.Nakagawa T, Shimizu S, Watanabe T, Yamaguchi O, Otsu K, Yamagata H, Inohara H, Kubo T, Tsujimoto Y (2005) Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature 434:652–658 [DOI] [PubMed] [Google Scholar]
  • 101.Baines CP, Kaiser RA, Purcell NH, Blair NS, Osinska H, Hambleton MA, Brunskill EW, Sayen MR, Gottlieb RA, Dorn GW, Robbins J, Molkentin JD (2005) Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature 434:658–662 [DOI] [PubMed] [Google Scholar]
  • 102.Tsujimoto Y, Shimizu S (2007) Role of the mitochondrial membrane permeability transition in cell death. Apoptosis 12:835–840 [DOI] [PubMed] [Google Scholar]
  • 103.Simpkins JW, Dykens JA (2008) Mitochondrial mechanisms of estrogen neuroprotection. Brain Res Rev 57:421–430 [DOI] [PubMed] [Google Scholar]
  • 104.Lobaton CD, Vay L, Hernandez-Sanmiguel E, Santodomingo J, Moreno A, Montero M, Alvarez J (2005) Modulation of mitochondrial Ca(2+) uptake by estrogen receptor agonists and antagonists. Br J Pharmacol 145:862–871 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Horvat A, Petrovic S, Nedeljkovic N, Martinovic JV, Nikezic G (2000) Estradiol affect Na-dependent Ca2+ efflux from synaptosomal mitochondria. Gen Physiol Biophys 19:59–71 [PubMed] [Google Scholar]
  • 106.Nilsen J, Diaz Brinton R (2003) Mechanism of estrogen-mediated neuroprotection: regulation of mitochondrial calcium and Bcl-2 expression. Proc Natl Acad Sci U S A 100:2842–2847 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Kim HJ, Magrane J, Starkov AA, Manfredi G (2012) The mitochondrial calcium regulator cyclophilin D is an essential component of oestrogen-mediated neuroprotection in amyotrophic lateral sclerosis. Brain 135:2865–2874 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Ribeiro RF Jr., Ronconi KS, Morra EA, Do Val Lima PR, Porto ML, Vassallo DV, Figueiredo SG, Stefanon I(2016) Sex differences in the regulation of spatially distinct cardiac mitochondrial subpopulations. Mol Cell Biochem 419:41–51 [DOI] [PubMed] [Google Scholar]
  • 109.Milerova M, Drahota Z, Chytilova A, Tauchmannova K, Houstek J, Ostadal B (2016) Sex difference in the sensitivity of cardiac mitochondrial permeability transition pore to calcium load. Mol Cell Biochem 412:147–154 [DOI] [PubMed] [Google Scholar]
  • 110.Arieli Y, Gursahani H, Eaton MM, Hernandez LA, Schaefer S (2004) Gender modulation of Ca(2+) uptake in cardiac mitochondria. J Mol Cell Cardiol 37:507–513 [DOI] [PubMed] [Google Scholar]
  • 111.Yang SH, Liu R, Perez EJ, Wen Y, Stevens SM Jr., Valencia T, Brun-Zinkernagel AM, Prokai L, Will Y, Dykens J, Koulen P, Simpkins JW (2004) Mitochondrial localization of estrogen receptor beta. Proc Natl Acad Sci U S A 101:4130–4135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Zheng J, Ramirez VD (1999) Purification and identification of an estrogen binding protein from rat brain: oligomycin sensitivity-conferring protein (OSCP), a subunit of mitochondrial F0F1-ATP synthase/ATPase. J Steroid Biochem Mol Biol 68:65–75 [DOI] [PubMed] [Google Scholar]
  • 113.Moreno AJ, Moreira PI, Custodio JB, Santos MS (2013) Mechanism of inhibition of mitochondrial ATP synthase by 17beta-estradiol. J Bioenerg Biomembr 45:261–270 [DOI] [PubMed] [Google Scholar]
  • 114.Du H, Guo L, Fang F, Chen D, Sosunov AA, McKhann GM, Yan Y, Wang C, Zhang H, Molkentin JD, Gunn-Moore FJ, Vonsattel JP, Arancio O, Chen JX, Yan SD (2008) Cyclophilin D deficiency attenuates mitochondrial and neuronal perturbation and ameliorates learning and memory in Alzheimer’s disease. Nat Med 14:1097–1105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Warne J, Pryce G, Hill JM, Shi X, Lenneras F, Puentes F, Kip M, Hilditch L, Walker P, Simone MI, Chan AW, Towers GJ, Coker AR, Duchen MR, Szabadkai G, Baker D, Selwood DL (2016) Selective Inhibition of the Mitochondrial Permeability Transition Pore Protects against Neurodegeneration in Experimental Multiple Sclerosis. J Biol Chem 291:4356–4373 [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES