Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Developmental Biology 426 (2017) 1–7

Contents lists available at ScienceDirect

Developmental Biology
journal homepage: www.elsevier.com/locate/developmentalbiology

Development or disease: duality of the mitochondrial permeability MARK


transition pore

María José Péreza, Rodrigo A. Quintanillaa,b,
a
Laboratory of Neurodegenerative Diseases, Universidad Autónoma de Chile, Santiago, Chile
b
Centro de Investigación y Estudio del Consumo de Alcohol en Adolescentes (CIAA), Santiago, Chile

A R T I C L E I N F O A BS T RAC T

Keywords: Mitochondria is not only a dynamic organelle that produces ATP, but is also an important contributor to cell
Mitochondria functions in both development and cell death processes. These paradoxical functions of mitochondria are
Mitochondrial permeability transition pore partially regulated by the mitochondrial permeability transition pore (mPTP), a high-conductance channel that
Cardiac development can induce loss of mitochondrial membrane potential, impairment of cellular calcium homeostasis, oxidative
Neuronal development
stress, and a decrease in ATP production upon pathological activation. Interestingly, despite their different
Oxidative stress
etiologies, several neurodegenerative diseases and heart ischemic injuries share mitochondrial dysfunction as a
Calcium
common element. Generally, mitochondrial impairment is triggered by calcium deregulation that could lead to
mPTP opening and cell death. Several studies have shown that opening of the mPTP not only induces
mitochondrial damage and cell death, but is also a physiological mechanism involved in different cellular
functions. The mPTP participates in regular calcium-release mechanisms that are required for proper metabolic
regulation; it is hypothesized that the transient opening of this structure could be the principal mediator of
cardiac and brain development. The mPTP also plays a role in protecting against different brain and cardiac
disorders in the elderly population. Therefore, the aim of this work was to discuss different studies that show
this controversial characteristic of the mPTP; although mPTP is normally associated with several pathological
events, new critical findings suggest its importance in mitochondrial function and cell development.

1. Introduction 1999; Hou et al., 2013; Kasahara and Scorrano, 2014; Kwong and
Molkentin, 2015). The mPTP is a high-conductance channel that, upon
The mitochondria is a highly dynamic organelle, commonly nick- opening, generates a sudden increase in inner mitochondrial mem-
named “the powerhouse of the cell” for its ability to produce cellular brane (IMM) permeability to ions and small solutes (Haworth and
energy in a highly efficient manner (Mattson et al., 2008). The main Hunter, 1979; Hunter and Haworth, 1979a, 1979b). These changes
function of the mitochondria is to convert the energy derived from also affect regulation of the mitochondrial membrane potential
nutrients into heat and ATP, but it is also a major contributor to cell (Galluzzi et al., 2009), calcium homeostasis (Elrod et al., 2010),
fate determination, as this organelle is able to control both apoptotic reactive oxygen species (ROS) production (Hou et al., 2014), ATP
and necrotic cell death (Folmes et al., 2012; Nunnari and Suomalainen, production (Budd and Nicholls, 1996), and cell death (Galluzzi et al.,
2012). These diametrically opposed functions of mitochondria are 2009). Recent advances have determined that the mPTP could also be
particularly relevant in the heart and brain, where mitochondria supply the principal mediator of the duality of cell development and cell death.
over 90% of the energy demands (Wallace, 2005). In fact, several The mPTP plays a role in cardiac and neurodegenerative pathologies
studies have demonstrated that, in these specific organs, metabolic (Hou et al., 2013; Kwong et al., 2015), but is also involved in heart and
control through mitochondria is not only related to cell fate (Folmes brain development (Hom et al., 2011; Porter et al., 2011; Mattson
et al., 2012), but also plays an important role in cell differentiation et al., 2008), highlighting the importance of the mPTP in health and
(Kasahara and Scorrano, 2014). disease.
Interestingly, it has been shown that all these mitochondrial
functions are at some point related to the opening or activation of
the mitochondrial permeability transition pore (mPTP) (Bernardi et al.,


Corresponding author at: Laboratory of Neurodegenerative Diseases, Universidad Autónoma de Chile, El Llano Subercaseaux 2801, 5to Piso, San Miguel 8910000, Santiago, Chile.
E-mail address: rodrigo.quintanilla@uautonoma.cl (R.A. Quintanilla).

http://dx.doi.org/10.1016/j.ydbio.2017.04.018
Received 7 February 2017; Received in revised form 26 April 2017; Accepted 26 April 2017
Available online 28 April 2017
0012-1606/ © 2017 Elsevier Inc. All rights reserved.
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

2. The mitochondrial permeability transition pore different working proposals about the structure of the pore itself.
However, both models suggest that ATP synthase forms part of the
The mPTP is a non-specific mega-channel that is sensitive to functional pore of the mPTP (Alavian et al., 2014; Giorgio et al., 2013).
perturbations in intracellular calcium homeostasis, permitting the One hypothesis indicates that the mPTP channel forms at the interface
passage of molecules smaller than 1.5 kDa in size (Elrod and of two monomers (associated into dimers) of ATP synthase (Giorgio
Molkentin, 2013). The opening of the pore results in an increase in et al., 2013). Purified dimers of the enzyme were reconstituted into
the permeability of the IMM and permits entry of metabolites into the lipid bilayers and stimulated with regulators of mPTP; it was shown
mitochondrial matrix space, which leads to a reduction in the efficiency that dimers of ATP synthase were capable of forming a current that was
of ATP production by uncoupling the electron transport system from electrophysiologically equivalent to that in the mPTP (Giorgio et al.,
ATP synthase activity (Elrod and Molkentin, 2013; Mnatsakanyan 2013). Since a current is only seen in dimers, and the genetic ablation
et al., 2016). Recent studies suggest that the mPTP is one of the of the subunits that form the dimers did not result in opening of the
principal mechanisms that leads to cell death, participating as a mPTP (Carraro et al., 2014), it was suggested that the pore structure of
metabolic regulator of cell energy homeostasis, a mitochondrial the mPTP forms at the membrane interface between two adjacent Fo
calcium transporter, and a superoxide (SO) efflux channel (Izzo et al., sectors (Giorgio et al., 2013).
2016). Another proposed model involves the c-subunit ring of the F1
The original mPTP model proposed that the channel was formed by subcomplex (Alavian et al., 2014). Both a- and c-subunits form the
three principal proteins: cyclophilin D (CyPD), located in the mito- proton channel of ATP synthase, permitting proton flux between the
chondrial matrix; the adenine nucleotide translocator (ANT), found in intermembrane space and the mitochondrial matrix (Pogoryelov et al.,
the inner membrane; the voltage-dependent anion channel (VDAC) in 2009). Interestingly, gene-silencing of isomers of the c-subunit in-
the outer membrane (Rao et al., 2014); and other interacting mito- hibited mPTP opening, and the overexpression of these subunits
chondrial molecules such as the phosphate carrier, BH3 proteins, and accelerated the kinetics of the opening response, increasing permeabi-
p53 (Elrod and Molkentin, 2013). However, genetic knock-out studies lization of the outer mitochondrial membrane (OMM) and resulting in
showed this model to be invalid, as they demonstrated that only the cell death (Azarashvili et al., 2014; Bonora et al., 2013). Furthermore,
deletion of the CyPD gene has a regulatory role in the activation of in purified ATP synthase extracts, it was found that the ring structure
mPTP (Baines et al., 2005; Nakagawa et al., 2006). In fact, studies in formed by c-subunits generated a current similar to that of the mPTP.
mouse models lacking ANT expression showed a functional state of These currents were inhibited by calcium, CyPD and CsA (Alavian,
mPTP that held two-fold more calcium than wild-type mitochondria; 2014). Therefore, the properties of the pore formed in the c-subunit
this suggests that ANT has an important regulatory role in the function ring suggests that this structure corresponds to the mPTP itself,
of the mPTP, but not in its structure (Kokoszka et al., 2004). On the indicating a strong relationship between the opening of the mPTP
other hand, the absence of one or more VDAC isoforms still allowed and ATP synthase activity (Alavian, 2014).
opening of the mPTP (Baines et al., 2007), indicating that CyPD is the It is still a matter of discussion whether the extraction methods and
major factor responsible for the activation of the mPTP (Elrod and protocols used were adequate to reach a clear conclusion about the
Molkentin, 2013; Izzo et al., 2016). components of mPTP (Bernardi et al., 2015a, 2015b; Chinopoulos
Interestingly, in recent years, it has been proposed that ATP et al., 2014; Konig et al., 2016; Maltecca et al., 2015); complementary
synthase is a major component of the mPTP (Jonas et al., 2015). evidence has indicated new roles of ATP synthase in mPTP regulation.
This enzyme is a highly conserved molecular machine that catalyzes the Although the mechanism for mPTP formation is an open question,
production of ATP from ADP and Pi, powered by an electrochemical studies suggest that ATP synthase could not only participate in mPTP
transmembrane gradient in mitochondria (Pogoryelov et al., 2009). opening by increasing IMM permeability, but could also directly affect
Furthermore, ATP synthase is located in the inner membrane cristae, respiratory chain functionality and cell survival (Alavian et al., 2014).
made up of a membrane component—known as Fo—composed of nine The permeability of the IMM is a critical and decisive point between life
polypeptides (a, b, c, d, e, f, g, F6, and F8) and a soluble catalytic core and death in the cell; the mPTP has been suggested as a master
complex—known as F1—consisting of five different subunits (α, β, γ, δ/ regulator of both metabolism and cell death. With that in mind, a
OSCP, and ε) (Pogoryelov et al., 2009). The F1 subcomplex forms a relative closure or opening of the mPTP could greatly affect cellular
catalytic head group and a rotating central stalk, and the Fo sub- functions in organs and tissues with a high energy demand, such as the
complex comprises the membrane region and the peripheral stalk; the heart and brain (Mnatsakanyan et al., 2016).
oligomycin sensitivity-conferrin protein (δ/OSCP) subunit connects the
peripheral stalk to the catalytic domain (Pogoryelov et al., 2009). 3. Pathological implications of mPTP opening
Regarding ATP synthase and its role in mPTP regulation, important
studies have described that CyPD associates with ATP synthase As we previously mentioned, the formation and consequent open-
specifically through the δ/OSCP subunit; this interaction is inhibited ing of the mPTP is a key factor in mitochondrial dysfunction and
by cyclosporine A (CsA), a fungus-derived drug that reduces the mitochondria-driven cell death (Bernardi et al., 2015a, 2015b; Bonora
probability of mPTP opening through the inhibition of CyPD (Giorgio et al., 2013; Jonas et al., 2015; Kroemer et al., 2007). In fact, the
et al., 2009, 2013). In fact, this inhibitory drug increased the activity of general consensus is that uncontrolled opening of the mPTP leads
both ATP synthesis and hydrolysis through displacement of CyPD from irreversibly to necrosis or apoptosis (Bernardi et al., 2015a, 2015b).
ATP synthase, suggesting that the interaction of CyPD with ATP Under physiological conditions, when the mitochondria is exposed to
synthase may switch the enzyme to a lower catalytic activity state high concentrations of calcium, it can undergo a massive and perma-
(Giorgio et al., 2009). On the other hand, it was suggested that the δ/ nent swelling that leads to an abrupt increase in permeability to small
OSCP subunit also influences the accessibility of mPTP calcium- solutes of the IMM, resulting in the collapse of the chemiosmotic
binding sites. It seems that δ/OSCP affects the affinity of the metal gradient across the IMM—also known as the mitochondrial perme-
binding sites of ATP synthase, and thus, the ability with which calcium ability transition (Elrod and Molkentin, 2013). The resultant uncou-
can interact with the enzyme. Interestingly, when calcium is bound to pling of oxidative phosphorylation results in a decrease in ATP
the enzyme, ATP synthase activity is not coupled to proton transloca- generation and a subsequent increase in reactive oxygen species
tion, suggesting that calcium induces conformational changes in ATP (ROS) production. In addition, mitochondrial swelling may rupture
synthase, which could open the mPTP and stop ATP synthesis (Giorgio the OMM, releasing cytochrome C and initiating apoptosis (Izzo et al.,
et al., 2013). 2016; Kroemer et al., 2007; Rao et al., 2014).
With respect to the mechanism for channel formation, there are two Interestingly, despite their different etiologies, several neurodegen-

2
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

erative diseases and ischemic heart lesions share the common patho- formulation of CsA used in the latter study might not have been
logic element of cytosolic calcium deregulation, which causes either effective for preventing myocardial reperfusion injury (Hausenloy and
overload or uptake defects leading to mPTP opening (Abeti and Yellon, 2015). Thus, although the administration of mPTP inhibitors to
Abramov, 2015). Several studies have suggested that in the heart, the reduce MI damage is still controversial, it remains a promising
opening of the mPTP during early reperfusion after ischemia is a alternative treatment.
harmful event that induces further damage to the myocardium
(Griffiths and Halestrap, 1993). In contrast, in the brain, mPTP is 3.2. mPTP and neurodegenerative disorders
persistently activated in neurons under pathological conditions, de-
creasing the ability of the cells to regulate calcium homeostasis, which In most common neurological disorders, mitochondria have been
contributes to neurodegeneration (Friberg and Wieloch, 2002). suggested as an important factor in the dysfunction and degeneration
Importantly, the genetic and pharmacological blockage of CyPD of neurons (Mattson et al., 2008). In these diseases, deregulated
prevents mitochondrial dysfunction in both cases (Rao et al., 2014). calcium homeostasis and increased ROS levels are apparently mediated
by mPTP opening. These disorders include excitotoxicity (Pivovarova
3.1. Role of the mPTP in heart injury and Andrews, 2010), Huntington's disease (HD) (Quintanilla et al.,
2013, 2016), Alzheimer's disease (AD) (Du et al., 2011), and
Cardiomyopathy is a general term used to describe heart muscle Parkinson's disease (PD) (Martin et al., 2014). Under normal condi-
diseases; these disorders are related to an increase in the size of the tions, mitochondria can buffer substantial amounts of calcium ions
heart or the thickening and stiffening of heart muscle, compromising during neurotransmission; the mitochondrial membrane potential and
the function of the whole organ (Kwong and Molkentin, 2015). One of several antioxidant mechanisms prevent the opening of the mPTP
the major causes of acute cardiomyopathies is myocardial ischemia (Friberg and Wieloch, 2002). Accumulating evidence indicates that
(MI), which is inherited—in a minority of cases—or caused by chronic these abnormalities are rescued by the action of CsA, or by the genetic
diseases such as hypertension and diabetes (Walters et al., 2012). ablation of CyPD (Abeti and Abramov, 2015). Indeed, CyPD-deficient
Regardless of the cause, alterations in mitochondrial bioenergetics mice displayed resistance to apoptotic and non-apoptotic neuronal cell
appear to play an important role in the inception and progression of death, as well as to mitochondrial dysfunction (Martin et al., 2014).
this disease (Walters et al., 2012), manifesting as an inhibition of Different groups have reported increased CyPD expression in AD;
respiratory complex activity, an increase in proton leakage from the this may facilitate the opening of the mPTP (Du et al., 2010). On the
IMM (Ingwall, 2009), increased ROS production (Keith et al., 1998), other hand, several studies have indicated that ROS and calcium
mitochondrial calcium overload (Luo and Anderson, 2013), and finally, overload are potent inducers of mPTP opening (Du and Yan, 2010).
opening of the mPTP (Kwong and Molkentin, 2015). During MI For example, in AD, the amyloid-beta peptide (Aβ), which accumulates
progression, oxygen deprivation alters mitochondrial function and in high amounts during the development of this disease, induced ROS
ATP synthesis, causing an important reduction in cardiac ATP produc- formation and calcium stress, which are optimal conditions for mPTP
tion (Jennings et al., 1991). Moreover, because calcium regulation formation; it is therefore logical to hypothesize that massive mPTP
requires ATP, and the mPTP remains open as ATP pools dissipate, formation occurs when Aβ interacts with mitochondria (Du and Yan,
calcium levels in the ischemic heart are further elevated (Kwong and 2010). Interestingly, the genetic absence of CyPD protects against this
Molkentin, 2015). In addition to calcium elevation, ROS production is phenomenon and slows cognitive loss in APP/PS1 mice, an animal
also increased in the ischemic heart (Murphy and Steenbergen, 2008). model of AD that recapitulates several features of this disease,
Thus, the ischemic heart is primed for a prolonged mPTP opening cycle including aggregation of Aβ in the brain (Du et al., 2010, 2011; Du
that ultimately leads to cardiomyocyte death (Murphy and and Yan, 2010). In addition, it was established that neurons that do not
Steenbergen, 2008). express CyPD show a decrease in Aβ-dependent ROS generation,
In support of these findings, CyPD-deficient mice subjected to increase in the buffering capacity of calcium, improved mitochondrial
prolonged MI displayed reduced mortality with a 50% reduction in the respiratory function, and attenuation of abnormalities in synaptic
final infarct size (Lim et al., 2011; Yellon and Hausenloy, 2007), plasticity present in AD (Du et al., 2011; Guo et al., 2013). Similarly,
suggesting that mPTP inhibition could be beneficial for treating this in HD, it was shown that mutant huntingtin affects mitochondrial
disease. In fact, additional studies using the CyPD KO mouse model function by modifying mitochondrial calcium homeostasis, mitochon-
showed protection not only in ischemia/reperfusion (I/R) models, but drial morphology, and ROS handling through the opening of mPTP
also in mice where the infarction occurred in the brain and kidney (Brustovetsky et al., 2003; Choo et al., 2004; Milakovic et al., 2006;
(Elrod et al., 2010). This effect should be considered important as a Quintanilla et al., 2013, 2016). These mitochondrial deficiencies in
potential pharmacological target for cardioprotection (Halestrap, mutant huntingtin-expressing cells only occurred after calcium over-
2009). Drugs that inhibit mPTP activation—such as sanglifehrin load, simulating the pathophysiological conditions that may occur in
(SfA), CsA, and other non-immunosuppressive analogs of CyPD, such HD (Milakovic et al., 2006; Quintanilla et al., 2013). As in AD, CsA
as [MeAla6]-cyclosporine (Griffiths and Halestrap, 1995), Debio-025 treatment ameliorated these toxic events; treatment also resulted in the
(Gomez et al., 2007), and NIM811 (Argaud et al., 2005)—have the reduction of neuronal loss induced by calcium stress in mutant
potential to be of great value in protecting the heart during cardiac huntingtin cells, supporting the role of mPTP in the mitochondrial
surgery for treatment of coronary thrombosis. More importantly, a dysfunction observed in HD (Quintanilla et al., 2013, 2016).
pilot clinical study in patients with MI showed that the administration More interestingly, it seems that opening of the mPTP has effects on
of CsA immediately when blood flow was restored reduced infarct size normal ageing (Gauba et al., 2017). For example, analysis of brain
by 40% compared with the placebo treatment (Piot et al., 2008). samples of ageing mice showed an increase in CyPD expression and an
Interestingly, in a subsequent follow-up study with the same group of increase in its interaction with ATP synthase (Gauba et al., 2017).
patients, reduction of the infarct size upon administration of one dose These alterations coincided with the presence of mitochondrial un-
of CsA was maintained over 6 months post-treatment, and these coupling, calcium-induced mitochondrial swelling, decreased ATP
patients displayed improved post-infarction cardiac function production by oxidative phosphorylation, and enhanced SO produc-
(Mewton et al., 2010). However, these findings were questioned in tion, indicating that the mPTP could be an important factor that
another study, where administration of CsA did not show a reduction in controls mitochondrial function during ageing. Interestingly, these
the infarct size (measured by the release of cardiac enzymes) or any mitochondrial alterations were significantly ameliorated by the genetic
improvement in clinical endpoints and mortality of the patients ablation of CyPD (Gauba et al., 2017). Those results are in agreement
(Ghaffari et al., 2013). However, subsequent analysis showed that the with studies where the knockout of CyPD presented an improvement in

3
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

the functional properties of brain mitochondria, and therefore pro- mitochondrial biology occur in cardiac stem cells during differentiation
tected mitochondrial function from calcium-induced stress (Chung et al., 2007, 2008; Porter et al., 2011). These changes appear to
(Gainutdinov et al., 2015). mirror what has been observed in the embryonic heart, where the
mPTP is open at early stages of myocyte differentiation (Beutner et al.,
4. Physiological roles of the mPTP 2014; Hom et al., 2011).
Prior to differentiation, cardiac embryonic stem cells have frag-
Pioneer studies showed that opening of the mPTP not only leads to mented mitochondrial networks with a specific perinuclear distribu-
mitochondrial damage and cell death, but is also a physiological tion; this event is interestingly related to the fact that the primary
mechanism that mitochondria use to promote cell health (Elrod and source of energy to the cells is from anaerobic glycolysis, while
Molkentin, 2013). In fact, evidence of the physiological mechanism of mitochondria show minimal participation (Beutner et al., 2014;
mPTP opening suggests two possible gating mechanisms: irreversible Porter et al., 2011). Upon induction of differentiation, mitochondrial
full-conductance opening for permanent permeability leading to cell fission decreases and the mitochondrial network spreads throughout
death, or a second strategy that consists of a transient and flickering the cell (Porter et al., 2011). There is also an increase in the complexity
short-term opening of mPTP with smaller and more variable conduc- of the IMM structure and the utilization of mitochondria and aerobic
tance (Hou et al., 2014; Wang et al., 2008). respiration as the primary source of energy (Porter et al., 2011).
The mPTP appears to act as a normal calcium-release mechanism Interestingly, It was found that mPTP opening in the early heart
that is required for proper metabolic regulation (Elrod and Molkentin, primordium is a physiological event in the development of the organ in
2013), and it is hypothesized that its transient opening may regulate which myocytes exhibit low mitochondrial membrane potential, high
cytosolic calcium when a calcium overload occurs in neurons and levels of ROS, and opening of the mPTP (Hom et al., 2011).
myocytes (Bernardi and von Stockum, 2012; Gainutdinov et al., 2015; Interestingly, pharmacological closure of the mPTP with CsA led to
Hom et al., 2011). Another possible role of the mPTP is its capacity to dramatic maturation of mitochondrial structure and function, decreas-
regulate energy production in the cell, because a physical interaction ing intracellular ROS levels and increasing mitochondrial membrane
between CyPD and ATP synthase has been shown; this also suggests potential, which accelerated myocyte differentiation (Hom et al., 2011).
that mPTP has roles as a metabolic regulator in the cell and in the Closure of the mPTP using CsA and antioxidant treatments also led to
generation of SO flashes (Beutner et al., 1998; Elrod and Molkentin, further myocyte differentiation; the latter treatment was effective
2013; Mnatsakanyan et al., 2016). regardless of the state of the mPTP (Hom et al., 2011). These studies
In this context, one of the primary functions of mitochondria is the indicate that mPTP function regulates myocyte differentiation via
regulation of redox homeostasis, which could be interrupted by a redox-signaling pathways and that changes in mPTP function, mito-
constant flow of ROS produced by the respiratory chain (Jensen et al., chondrial maturation, and ROS levels contribute to a novel mechanism
1996). Therefore, mitochondria have the ability to maintain oxidative that protects the embryonic heart during a period of increasing energy
equilibrium, mediated by the action of several antioxidant enzymes demand and low oxygen supply (Hom et al., 2011). These reports
such as catalase, thioredoxin, and glutathione peroxidase (Droge, highlight the mitochondria, and more specifically the mPTP, as a gating
2002). Even though ROS production by mitochondria could lead to mechanism underlying differentiation in the developing heart, impli-
oxidative damage, recent studies have suggested that ROS could also cating a cross-talk between genetic and metabolic signaling (Folmes
contribute to redox signaling in a variety of pathways in differentiation et al., 2012; Hom et al., 2011).
and organogenesis (Owusu-Ansah and Banerjee, 2009), cell fate In the same context, other studies have related ANT, a regulator of
regulation (Maryanovich and Gross, 2013), and the stress response mPTP, with the development and repair of cardiac tissue (Kokoszka
(Adler et al., 1999). In contrast to basal ROS production, it has been et al., 2016). Fetal and adult heart tissues express two ANT isoforms:
determined that the generation of SO flashes or “mitoflashes,” which ANT1, which is highly expressed in the heart and brain; and ANT2,
consist of an intermittent and dynamic liberation of SO in a frequency- which is expressed in the whole body (Bauer et al., 1999; Teixeira et al.,
modulated manner, could function in physiological signaling (Wang 2015). Mouse fetal and adult hearts express both ANT isoforms
et al., 2012, 2014). Further evidence strongly suggests that SO flashes (Haraguchi et al., 1993; Neckelmann et al., 1987). Interestingly,
and oxidative signaling mechanisms are involved in mPTP activation genetic knockdown of ANT1 does not impair fetal development, but
(Mnatsakanyan et al., 2016). Together with an increase in SO, cells does generate hypertrophic cardiomyopathy and induce apoptosis in
experience an abrupt dissipation of the mitochondrial membrane postnatal mice (Chevrollier et al., 2011). This effect is mediated
potential, irreversible loss of small solutes from the matrix, and through increased association with the IκBα–NFκB complex, which is
transient mitochondrial swelling; these events are all characteristic of later sequestered within the mitochondrial intermembrane space,
mPTP opening (Hou et al., 2014; Wang et al., 2008, 2012). However, impeding migration to the nucleus (Bauer et al., 1999). In addition,
there is still controversy surrounding the exact role of these SO flashes overexpression of ANT1 inhibited expression of the anti-apoptotic
in development (De Marchi et al., 2014). Recent advances have shown proteins Bcl-XL, c-IAP2, and Sod2, favoring the opening of the mPTP
that mitoflashes not only represent a fundamental step during mito- and eventually activating the intrinsic apoptosis pathway (Bauer et al.,
chondrial function, but also play an important role in metabolism, 1999). Further studies generated a heterozygous ANT2 null mouse
ageing, diabetes, wound healing, and, most importantly, a possible role model, in which fetuses presented with embryonic death at E14.5
in cell differentiation and stemness (Ding et al., 2015; Hom et al., (Kokoszka et al., 2016). ANT2 null mice showed significant cardiac
2011; Shen et al., 2014; Vega-Naredo et al., 2014). All these functions developmental failure with immature cardiomyocytes and a high
are mediated by the opening of the mPTP (Hou et al., 2014). degree of hyperproliferation (Kokoszka et al., 2016). Interestingly,
these mice suffered from cardiac failure due to hypertrabeculation and,
4.1. The mPTP in cardiac development more importantly, showed a high number of swollen mitochondria,
indicating that this mutation has an important effect on mitochondrial
Recently, important studies have demonstrated that mitochondria function (Kokoszka et al., 2016). ANTs have two main functions; they
are essential organelles for the heart and cardiac functions (Hom et al., function as mitochondrial–cytosol ATP/ADP exchangers, and in mod-
2011; Ingraham et al., 2009). In fact, during cardiac development and ulating the opening of the mPTP (Kokoszka et al., 2004). Previous
myocyte differentiation, the structures of individual mitochondria and studies suggest that ANT2 stimulates the closing of the mtPTP
the cellular mitochondrial network are in constant flux, suggesting (Chevrollier et al., 2011), while ANT1 enhances its opening (Bauer
direct participation in developmental mechanisms (Porter et al., 2011). et al., 1999). Therefore, the developmental toxicity shown in ANT2-null
Recent works have documented that several changes in energetics and mice may be the result of the mPTP being constantly opened in the

4
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

Fig. 1. Dual roles of the mitochondrial permeability transition pore (mPTP) in health and disease. (A). During myocyte stemness, which is defined as the ability to self-renew and
differentiate into different cell types, the mPTP is open; this results in a decrease in levels of ATP and mitochondrial membrane potential, as well as an increase in levels of cytosolic
calcium and reactive oxygen species (ROS). On the contrary, in neuronal stemness it seems that the mPTP remains closed, maintaining decreased levels of ROS and appropriate
production of ATP. (B). In cardiomyocytes, closure of the mPTP led to dramatic maturation of mitochondrial structure and function, decreasing intracellular ROS levels and increasing
mitochondrial membrane potential, which ultimately accelerated cell differentiation. However, neuronal development seems to be mediated by the opening of the mPTP, specifically by
the controlled increase in ROS levels in the form of superoxide mitoflashes leading to differentiation. (C). Despite their different etiologies, several neurodegenerative diseases and
cardiomyopathies share mitochondrial dysfunction as a common element, specifically the impairment of cytosolic calcium regulation; this may cause either calcium overload or uptake
defects in calcium homeostasis, leading to mPTP opening. In the heart, the opening of the mPTP during early reperfusion after ischemia is a harmful event that induces further damage
to the myocardium. In contrast, in the brain, mPTP is persistently activated in neurons exposed to pathological conditions, impairing their ability to regulate calcium. Only at this point is
the opening of the mPTP considered a pathological event.

cardiomyocytes, thus impairing cardiac cell maturation and resulting in of mPTP (Hou et al., 2012). Therefore, it has been shown that both
cardiomyocyte hyperproliferation (Kokoszka et al., 2016). mitochondrial ROS scavengers and mPTP inhibitors—such as CsA—
reduced the frequency of mitoflashes and enhanced NPC proliferation,
whereas prolonged mPTP opening and SO generation increased the
4.2. Role of the mPTP in nervous system development incidence of mitoflashes and promoted neuronal differentiation of
NPCs (Hou et al., 2012). This minimal and transient ROS production
During development of the nervous system, neural stem cells negatively regulates the self-renewal of NPCs and suggests that
proliferate and differentiate into neurons in the process of neurogen- dynamic mitochondrial ROS generation in the form of mPTP-depen-
esis (Mattson et al., 2008). The newly formed neurons then grow an dent SO flashes serves as a signaling mechanism that controls the fate
axon and dendrites and eventually form synapses; during this process, of NPCs (Hou et al., 2012). Interestingly, similar to these findings, CsA
many newly generated neurons undergo constant changes in metabo- inhibited the differentiation of hematopoietic progenitor cells and
lism and energy sources (Mattson et al., 2008). It was reported that the vascular progenitor cells (Arnold et al., 2000; Davies et al., 2005),
proliferation and differentiation of cultured embryonic mouse cortical but enhanced the differentiation of natural killer cells from hemato-
neural progenitor cells (NPCs) was affected by the concentration of poietic progenitors and cardiac cells from induced pluripotent stem
oxygen in the culture; lower oxygen levels promoted proliferation, cells (Fujiwara et al., 2011; Kosugi and Shearer, 1991). These findings
while higher oxygen levels suppressed this process (Morrison et al., reveal that there is great complexity and specificity in SO generation
2000; Smith et al., 2000; Tsatmali et al., 2005). and opening of the mPTP. These events are dependent on the specific
Interestingly, Hou et al. described that in NPCs, neuronal differ- cell type, the subcellular localization of mitochondria, and the temporal
entiation may be regulated by mitochondrial SO flashes; specific, focal, occurrence within the developmental stage (Hou et al., 2013, 2012).
and temporally-constrained bursts of mitochondrial SO negatively
regulate NPC proliferation independently of global changes in cytosolic
ROS levels (Hou et al., 2012). Indeed, when the authors limited the 5. Conclusions
mitoflash frequency using ROS scavengers, NPC proliferation in-
creased; conversely, increasing mitochondrial SO flashes promoted The opening of the mPTP was normally associated with pathologi-
neuronal differentiation (Hou et al., 2012). On the other hand, it has cal events such as neurodegenerative diseases or ischemic injury in
been shown that during neurogenesis of NPCs there is a switch in brain and heart, which all share a mitochondrial dysfunction condition
cellular energy production from anaerobic glycolysis to aerobic mito- directly related to the activation of mPTP (Fig. 1). However, several
chondrial oxidative phosphorylation in differentiated neurons novel physiological functions of mPTP in development have also been
(Candelario et al., 2013). Interestingly, increasing bursts of mitochon- discovered, as mPTP participates in the physiological calcium-release
drial SO production are associated with this switch mechanism in mechanisms that are required for proper metabolic regulation of the
cellular energy metabolism (Hou et al., 2012). cell; this structure could be an important player in the regulation of
With respect to the participation of the mPTP in neural develop- heart and brain development (Fig. 1). Furthermore, it seems that the
ment, it was reported that NPCs exhibit intermittent spontaneous conversion of a physiological opening event into a pathological channel
bursts of mitochondrial SO generation that require a transient opening opening may be related to the energetic context and the metabolic

5
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

status of the cell; nevertheless, the factors that regulate the transition 15980–15985.
Chevrollier, A., Loiseau, D., Reynier, P., Stepien, G., 2011. Adenine nucleotide
from physiological to pathological mPTP opening in the mitochondria translocase 2 is a key mitochondrial protein in cancer metabolism. Biochim. Biophys.
are not completely understood. Acta 1807, 562–567.
It is relevant to emphasize that these findings could help us to Chinopoulos, C., Kiss, G., Kawamata, H., Starkov, A.A., 2014. Measurement of ADP-ATP
exchange in relation to mitochondrial transmembrane potential and oxygen
understand the mechanisms of cell differentiation and the importance consumption. Methods Enzymol. 542, 333–348.
of mitochondria in health and disease, but they also highlight the Choo, Y.S., Johnson, G.V., MacDonald, M., Detloff, P.J., Lesort, M., 2004. Mutant
significance of research on drugs that may modulate the activation of huntingtin directly increases susceptibility of mitochondria to the calcium-induced
permeability transition and cytochrome c release. Human. Mol. Genet. 13,
mPTP in the search for possible therapies in degenerative diseases 1407–1420.
associated with either ageing or development. Chung, S., Dzeja, P.P., Faustino, R.S., Perez-Terzic, C., Behfar, A., Terzic, A., 2007.
Mitochondrial oxidative metabolism is required for the cardiac differentiation of
stem cells. Nat. Clin. Pract. Cardiovasc. Med. 4 (Suppl. 1), S60–S67.
Acknowledgments
Chung, S., Dzeja, P.P., Faustino, R.S., Terzic, A., 2008. Developmental restructuring of
the creatine kinase system integrates mitochondrial energetics with stem cell
This work has been supported by FONDECYT # 1140968 and cardiogenesis. Ann. N. Y. Acad. Sci. 1147, 254–263.
1170441 (RAQ), as well as CONICYT PIA Anillo ACT1411 (RAQ). Davies, W.R., Wang, S., Oi, K., Bailey, K.R., Tazelaar, H.D., Caplice, N.M., McGregor,
C.G., 2005. Cyclosporine decreases vascular progenitor cell numbers after cardiac
transplantation and attenuates progenitor cell growth in vitro. J. Heart Lung
References Transplant. 24, 1868–1877.
De Marchi, E., Bonora, M., Giorgi, C., Pinton, P., 2014. The mitochondrial permeability
transition pore is a dispensable element for mitochondrial calcium efflux. Cell
Abeti, R., Abramov, A.Y., 2015. Mitochondrial Ca(2+) in neurodegenerative disorders. Calcium 56, 1–13.
Pharmacol. Res. 99, 377–381. Ding, Y., Fang, H., Shang, W., Xiao, Y., Sun, T., Hou, N., Pan, L., Sun, X., Ma, Q., Zhou,
Adler, V., Yin, Z., Tew, K.D., Ronai, Z., 1999. Role of redox potential and reactive oxygen J., Wang, X., Zhang, X., Cheng, H., 2015. Mitoflash altered by metabolic stress in
species in stress signaling. Oncogene 18, 6104–6111. insulin-resistant skeletal muscle. J. Mol. Med. 93, 1119–1130.
Alavian, K.N., Beutner, G., Lazrove, E., Sacchetti, S., Park, H.A., Licznerski, P., Li, H., Droge, W., 2002. Free radicals in the physiological control of cell function. Physiol. Rev.
Nabili, P., Hockensmith, K., Graham, M., Porter, G.A., Jr., Jonas, E.A., 2014. An 82, 47–95.
uncoupling channel within the c-subunit ring of the F1FO ATP synthase is the Du, H., Guo, L., Yan, S., Sosunov, A.A., McKhann, G.M., Yan, S.S., 2010. Early deficits in
mitochondrial permeability transition pore. Proc. Natl. Acad. Sci. USA 111, synaptic mitochondria in an Alzheimer's disease mouse model. Proc. Natl. Acad. Sci.
10580–10585. USA 107, 18670–18675.
Argaud, L., Gomez, L., Gateau-Roesch, O., Couture-Lepetit, E., Loufouat, J., Robert, D., Du, H., Guo, L., Zhang, W., Rydzewska, M., Yan, S., 2011. Cyclophilin D deficiency
Ovize, M., 2005. Trimetazidine inhibits mitochondrial permeability transition pore improves mitochondrial function and learning/memory in aging Alzheimer disease
opening and prevents lethal ischemia-reperfusion injury. J. Mol. Cell. Cardiol. 39, mouse model. Neurobiol. Aging 32, 398–406.
893–899. Du, H., Yan, S.S., 2010. Mitochondrial permeability transition pore in Alzheimer's
Arnold, L.W., McCray, S.K., Tatu, C., Clarke, S.H., 2000. Identification of a precursor to disease: cyclophilin D and amyloid beta. Biochim. Biophys. Acta 1802, 198–204.
phosphatidyl choline-specific B-1 cells suggesting that B-1 cells differentiate from Elrod, J.W., Molkentin, J.D., 2013. Physiologic functions of cyclophilin D and the
splenic conventional B cells in vivo: cyclosporin A blocks differentiation to B-1. J. mitochondrial permeability transition pore. Circ. J. 77, 1111–1122.
Immunol. 164, 2924–2930. Elrod, J.W., Wong, R., Mishra, S., Vagnozzi, R.J., Sakthievel, B., Goonasekera, S.A.,
Azarashvili, T., Odinokova, I., Bakunts, A., Ternovsky, V., Krestinina, O., Tyynela, J., Karch, J., Gabel, S., Farber, J., Force, T., Brown, J.H., Murphy, E., Molkentin, J.D.,
Saris, N.E., 2014. Potential role of subunit c of F0F1-ATPase and subunit c of storage 2010. Cyclophilin D controls mitochondrial pore-dependent Ca(2+) exchange,
body in the mitochondrial permeability transition. Effect of the phosphorylation metabolic flexibility, and propensity for heart failure in mice. J. Clin. Investig. 120,
status of subunit c on pore opening. Cell Calcium 55, 69–77. 3680–3687.
Baines, C.P., Kaiser, R.A., Purcell, N.H., Blair, N.S., Osinska, H., Hambleton, M.A., Folmes, C.D., Dzeja, P.P., Nelson, T.J., Terzic, A., 2012. Mitochondria in control of cell
Brunskill, E.W., Sayen, M.R., Gottlieb, R.A., Dorn, G.W., Robbins, J., Molkentin, fate. Circ. Res. 110, 526–529.
J.D., 2005. Loss of cyclophilin D reveals a critical role for mitochondrial permeability Friberg, H., Wieloch, T., 2002. Mitochondrial permeability transition in acute
transition in cell death. Nature 434, 658–662. neurodegeneration. Biochimie 84, 241–250.
Baines, C.P., Kaiser, R.A., Sheiko, T., Craigen, W.J., Molkentin, J.D., 2007. Voltage- Fujiwara, M., Yan, P., Otsuji, T.G., Narazaki, G., Uosaki, H., Fukushima, H., Kuwahara,
dependent anion channels are dispensable for mitochondrial-dependent cell death. K., Harada, M., Matsuda, H., Matsuoka, S., Okita, K., Takahashi, K., Nakagawa, M.,
Nat. Cell Biol. 9, 550–555. Ikeda, T., Sakata, R., Mummery, C.L., Nakatsuji, N., Yamanaka, S., Nakao, K.,
Bauer, M.K., Schubert, A., Rocks, O., Grimm, S., 1999. Adenine nucleotide translocase-1, Yamashita, J.K., 2011. Induction and enhancement of cardiac cell differentiation
a component of the permeability transition pore, can dominantly induce apoptosis. J. from mouse and human induced pluripotent stem cells with cyclosporin-A. PLoS
Cell Biol. 147, 1493–1502. One 6, e16734.
Bernardi, P., Di Lisa, F., Fogolari, F., Lippe, G., 2015a. From ATP to PTP and back: a dual Gainutdinov, T., Molkentin, J.D., Siemen, D., Ziemer, M., Debska-Vielhaber, G.,
function for the mitochondrial ATP synthase. Circ. Res. 116, 1850–1862. Vielhaber, S., Gizatullina, Z., Orynbayeva, Z., Gellerich, F.N., 2015. Knockout of
Bernardi, P., Scorrano, L., Colonna, R., Petronilli, V., Di Lisa, F., 1999. Mitochondria and cyclophilin D in Ppif(-)/(-) mice increases stability of brain mitochondria against
cell death. Mechanistic aspects and methodological issues. Eur. J. Biochem. 264, Ca(2)(+) stress. Arch. Biochem. Biophys. 579, 40–46.
687–701. Galluzzi, L., Blomgren, K., Kroemer, G., 2009. Mitochondrial membrane
Bernardi, P., von Stockum, S., 2012. The permeability transition pore as a Ca(2+) release permeabilization in neuronal injury. Nat. Rev. Neurosci. 10, 481–494.
channel: new answers to an old question. Cell Calcium 52, 22–27. Gauba, E., Guo, L., Du, H., 2017. Cyclophilin D promotes brain mitochondrial F1FO ATP
Bernardi, P., Rasola, A., Forte, M., Lippe, G., 2015b. The mitochondrial permeability synthase dysfunction in aging mice. J. Alzheimer'S. Dis. 55, 1351–1362.
transition pore: channel formation by F-ATP synthase, integration in signal Ghaffari, S., Kazemi, B., Toluey, M., Sepehrvand, N., 2013. The effect of prethrombolytic
transduction, and role in pathophysiology. Physiol. Rev. 95 (4), 1111–1155. cyclosporine-A injection on clinical outcome of acute anterior ST-elevation
Beutner, G., Eliseev, R.A., Porter, G.A., Jr., 2014. Initiation of electron transport chain myocardial infarction. Cardiovasc. Ther. 31, e34–e39.
activity in the embryonic heart coincides with the activation of mitochondrial Giorgio, V., Bisetto, E., Soriano, M.E., Dabbeni-Sala, F., Basso, E., Petronilli, V., Forte,
complex 1 and the formation of supercomplexes. PLoS One 9, e113330. M.A., Bernardi, P., Lippe, G., 2009. Cyclophilin D modulates mitochondrial F0F1-
Beutner, G., Ruck, A., Riede, B., Brdiczka, D., 1998. Complexes between porin, ATP synthase by interacting with the lateral stalk of the complex. J. Biol. Chem. 284,
hexokinase, mitochondrial creatine kinase and adenylate translocator display 33982–33988.
properties of the permeability transition pore. Implication for regulation of Giorgio, V., von Stockum, S., Antoniel, M., Fabbro, A., Fogolari, F., Forte, M., Glick, G.D.,
permeability transition by the kinases. Biochim. Biophys. Acta 1368, 7–18. Petronilli, V., Zoratti, M., Szabo, I., Lippe, G., Bernardi, P., 2013. Dimers of
Bonora, M., Bononi, A., De Marchi, E., Giorgi, C., Lebiedzinska, M., Marchi, S., mitochondrial ATP synthase form the permeability transition pore. Proc. Natl. Acad.
Patergnani, S., Rimessi, A., Suski, J.M., Wojtala, A., Wieckowski, M.R., Kroemer, G., Sci. USA 110, 5887–5892.
Galluzzi, L., Pinton, P., 2013. Role of the c subunit of the FO ATP synthase in Gomez, L., Thibault, H., Gharib, A., Dumont, J.M., Vuagniaux, G., Scalfaro, P.,
mitochondrial permeability transition. Cell Cycle 12, 674–683. Derumeaux, G., Ovize, M., 2007. Inhibition of mitochondrial permeability transition
Brustovetsky, N., Brustovetsky, T., Purl, K.J., Capano, M., Crompton, M., Dubinsky, J.M., improves functional recovery and reduces mortality following acute myocardial
2003. Increased susceptibility of striatal mitochondria to calcium-induced infarction in mice. Am. J. Physiol. Heart Circ. Physiol. 293, H1654–H1661.
permeability transition. J. Neurosci. 23, 4858–4867. Griffiths, E.J., Halestrap, A.P., 1993. Protection by Cyclosporin A of ischemia/
Budd, S.L., Nicholls, D.G., 1996. Mitochondria, calcium regulation, and acute glutamate reperfusion-induced damage in isolated rat hearts. J. Mol. Cell. Cardiol. 25,
excitotoxicity in cultured cerebellar granule cells. J. Neurochem. 67, 2282–2291. 1461–1469.
Candelario, K.M., Shuttleworth, C.W., Cunningham, L.A., 2013. Neural stem/progenitor Griffiths, E.J., Halestrap, A.P., 1995. Mitochondrial non-specific pores remain closed
cells display a low requirement for oxidative metabolism independent of hypoxia during cardiac ischaemia, but open upon reperfusion. Biochem. J. 307 (Pt 1), 93–98.
inducible factor-1alpha expression. J. Neurochem. 125, 420–429. Guo, L., Du, H., Yan, S., Wu, X., McKhann, G.M., Chen, J.X., Yan, S.S., 2013. Cyclophilin
Carraro, M., Giorgio, V., Šileikyte, J., Sartori, G., Forte, M., Lippe, G., Zoratti, M., Szabó, D deficiency rescues axonal mitochondrial transport in Alzheimer's neurons. PLoS
I., Bernardi, P., 2014. Channel formation by yeast F-ATP synthase and the role of One 8, e54914.
dimerization in the mitochondrial permeability transition. J. Biol. Chem. 289, Halestrap, A.P., 2009. Mitochondria and reperfusion injury of the heart–a holey death

6
M.J. Pérez, R.A. Quintanilla Developmental Biology 426 (2017) 1–7

but not beyond salvation. J. Bioenerg. Biomembr. 41, 113–121. neurological disorders. Neuron 60, 748–766.
Haraguchi, Y., Chung, A.B., Torroni, A., Stepien, G., Shoffner, J.M., Wasmuth, J.J., Mewton, N., Croisille, P., Gahide, G., Rioufol, G., Bonnefoy, E., Sanchez, I., Cung, T.T.,
Costigan, D.A., Polak, M., Altherr, M.R., Winokur, S.T., et al., 1993. Genetic mapping Sportouch, C., Angoulvant, D., Finet, G., Andre-Fouet, X., Derumeaux, G., Piot, C.,
of human heart-skeletal muscle adenine nucleotide translocator and its relationship Vernhet, H., Revel, D., Ovize, M., 2010. Effect of cyclosporine on left ventricular
to the facioscapulohumeral muscular dystrophy locus. Genomics 16, 479–485. remodeling after reperfused myocardial infarction. J. Am. Coll. Cardiol. 55,
Hausenloy, D.J., Yellon, D.M., 2015. Targeting myocardial reperfusion injury–the search 1200–1205.
continues. N. Engl. J. Med. 373, 1073–1075. Milakovic, T., Quintanilla, R.A., Johnson, G.V., 2006. Mutant huntingtin expression
Haworth, R.A., Hunter, D.R., 1979. The Ca2+-induced membrane transition in induces mitochondrial calcium handling defects in clonal striatal cells: functional
mitochondria. II. Nature of the Ca2+ trigger site. Arch. Biochem. Biophys. 195, consequences. J. Biol. Chem. 281, 34785–34795.
460–467. Mnatsakanyan, N., Beutner, G., Porter, G.A., Alavian, K.N., Jonas, E.A., 2016.
Hom, J.R., Quintanilla, R.A., Hoffman, D.L., de Mesy Bentley, K.L., Molkentin, J.D., Physiological roles of the mitochondrial permeability transition pore. J. Bioenerg.
Sheu, S.S., Porter, G.A., Jr., 2011. The permeability transition pore controls cardiac Biomembr..
mitochondrial maturation and myocyte differentiation. Dev. Cell 21, 469–478. Morrison, S.J., Csete, M., Groves, A.K., Melega, W., Wold, B., Anderson, D.J., 2000.
Hou, Y., Ghosh, P., Wan, R., Ouyang, X., Cheng, H., Mattson, M.P., Cheng, A., 2014. Culture in reduced levels of oxygen promotes clonogenic sympathoadrenal
Permeability transition pore-mediated mitochondrial superoxide flashes mediate an differentiation by isolated neural crest stem cells. J. Neurosci. 20, 7370–7376.
early inhibitory effect of amyloid beta1-42 on neural progenitor cell proliferation. Murphy, E., Steenbergen, C., 2008. Mechanisms underlying acute protection from
Neurobiol. Aging 35, 975–989. cardiac ischemia-reperfusion injury. Physiol. Rev. 88, 581–609.
Hou, Y., Mattson, M.P., Cheng, A., 2013. Permeability transition pore-mediated Nakagawa, Y., Suzuki, T., Kamimura, H., Nagai, F., 2006. Role of mitochondrial
mitochondrial superoxide flashes regulate cortical neural progenitor differentiation. membrane permeability transition in N-nitrosofenfluramine-induced cell injury in
PLoS One 8, e76721. rat hepatocytes. Eur. J. Pharmacol. 529, 33–39.
Hou, Y., Ouyang, X., Wan, R., Cheng, H., Mattson, M.P., Cheng, A., 2012. Mitochondrial Neckelmann, N., Li, K., Wade, R.P., Shuster, R., Wallace, D.C., 1987. cDNA sequence of a
superoxide production negatively regulates neural progenitor proliferation and human skeletal muscle ADP/ATP translocator: lack of a leader peptide, divergence
cerebral cortical development. Stem Cells 30, 2535–2547. from a fibroblast translocator cDNA, and coevolution with mitochondrial DNA genes.
Hunter, D.R., Haworth, R.A., 1979a. The Ca2+-induced membrane transition in Proc. Natl. Acad. Sci. USA 84, 7580–7584.
mitochondria. I. The protective mechanisms. Arch. Biochem. Biophys. 195, Nunnari, J., Suomalainen, A., 2012. Mitochondria: in sickness and in health. Cell 148,
453–459. 1145–1159.
Hunter, D.R., Haworth, R.A., 1979b. The Ca2+-induced membrane transition in Owusu-Ansah, E., Banerjee, U., 2009. Reactive oxygen species prime Drosophila
mitochondria. III. Transitional Ca2+ release. Arch. Biochem. Biophys. 195, haematopoietic progenitors for differentiation. Nature 461, 537–541.
468–477. Piot, C., Croisille, P., Staat, P., Thibault, H., Rioufol, G., Mewton, N., Elbelghiti, R., Cung,
Ingraham, C.A., Burwell, L.S., Skalska, J., Brookes, P.S., Howell, R.L., Sheu, S.S., Pinkert, T.T., Bonnefoy, E., Angoulvant, D., Macia, C., Raczka, F., Sportouch, C., Gahide, G.,
C.A., 2009. NDUFS4: creation of a mouse model mimicking a Complex I disorder. Finet, G., Andre-Fouet, X., Revel, D., Kirkorian, G., Monassier, J.P., Derumeaux, G.,
Mitochondrion 9, 204–210. Ovize, M., 2008. Effect of cyclosporine on reperfusion injury in acute myocardial
Ingwall, J.S., 2009. On the control of metabolic remodeling in mitochondria of the failing infarction. N. Engl. J. Med. 359, 473–481.
heart. Circ. Heart Fail. 2, 275–277. Pivovarova, N.B., Andrews, S.B., 2010. Calcium-dependent mitochondrial function and
Izzo, V., Bravo-San Pedro, J.M., Sica, V., Kroemer, G., Galluzzi, L., 2016. Mitochondrial dysfunction in neurons. FEBS J. 277, 3622–3636.
permeability transition: new findings and persisting uncertainties. Trends Cell Biol.. Pogoryelov, D., Yildiz, O., Faraldo-Gomez, J.D., Meier, T., 2009. High-resolution
Jennings, R.B., Reimer, K.A., Steenbergen, C., 1991. Effect of inhibition of the structure of the rotor ring of a proton-dependent ATP synthase. Nat. Struct. Mol.
mitochondrial ATPase on net myocardial ATP in total ischemia. J. Mol. Cell. Cardiol. Biol. 16, 1068–1073.
23, 1383–1395. Porter, G.A., Jr., Hom, J., Hoffman, D., Quintanilla, R., de Mesy Bentley, K., Sheu, S.S.,
Jensen, W.A., Armstrong, J.M., De Giorgio, J., Hearn, M.T., 1996. Stability studies on pig 2011. Bioenergetics, mitochondria, and cardiac myocyte differentiation. Progress.
heart mitochondrial malate dehydrogenase: the effect of salts and amino acids. Pediatr. Cardiol. 31, 75–81.
Biochim. Biophys. Acta 1296, 23–34. Quintanilla, R.A., Godoy, J.A., Alfaro, I., Cabezas, D., von Bernhardi, R., Bronfman, M.,
Jonas, E.A., Porter, G.A., Jr., Beutner, G., Mnatsakanyan, N., Alavian, K.N., 2015. Cell Inestrosa, N.C., 2013. Thiazolidinediones promote axonal growth through the
death disguised: the mitochondrial permeability transition pore as the c-subunit of activation of the JNK pathway. PLoS One 8, e65140.
the F(1)F(O) ATP synthase. Pharmacol. Res. 99, 382–392. Quintanilla, R.A., Tapia, C., Perez, M.J., 2016. Possible role of mitochondrial
Kasahara, A., Scorrano, L., 2014. Mitochondria: from cell death executioners to permeability transition pore in the pathogenesis of Huntington disease. Biochem.
regulators of cell differentiation. Trends Cell Biol. 24, 761–770. Biophys. Res. Commun..
Keith, M., Geranmayegan, A., Sole, M.J., Kurian, R., Robinson, A., Omran, A.S., Rao, V.K., Carlson, E.A., Yan, S.S., 2014. Mitochondrial permeability transition pore is a
Jeejeebhoy, K.N., 1998. Increased oxidative stress in patients with congestive heart potential drug target for neurodegeneration. Biochim. Biophys. Acta 1842,
failure. J. Am. Coll. Cardiol. 31, 1352–1356. 1267–1272.
Kokoszka, J.E., Waymire, K.G., Flierl, A., Sweeney, K.M., Angelin, A., MacGregor, G.R., Shen, E.Z., Song, C.Q., Lin, Y., Zhang, W.H., Su, P.F., Liu, W.Y., Zhang, P., Xu, J., Lin, N.,
Wallace, D.C., 2016. Deficiency in the mouse mitochondrial adenine nucleotide Zhan, C., Wang, X., Shyr, Y., Cheng, H., Dong, M.Q., 2014. Mitoflash frequency in
translocator isoform 2 gene is associated with cardiac noncompaction. Biochim. early adulthood predicts lifespan in Caenorhabditis elegans. Nature 508, 128–132.
Biophys. Acta 1857, 1203–1212. Smith, J., Ladi, E., Mayer-Proschel, M., Noble, M., 2000. Redox state is a central
Kokoszka, J.E., Waymire, K.G., Levy, S.E., Sligh, J.E., Cai, J., Jones, D.P., MacGregor, modulator of the balance between self-renewal and differentiation in a dividing glial
G.R., Wallace, D.C., 2004. The ADP/ATP translocator is not essential for the precursor cell. Proc. Natl. Acad. Sci. USA 97, 10032–10037.
mitochondrial permeability transition pore. Nature 427, 461–465. Teixeira, F.K., Sanchez, C.G., Hurd, T.R., Seifert, J.R., Czech, B., Preall, J.B., Hannon,
Konig, T., Troder, S.E., Bakka, K., Korwitz, A., Richter-Dennerlein, R., Lampe, P.A., G.J., Lehmann, R., 2015. ATP synthase promotes germ cell differentiation
Patron, M., Muhlmeister, M., Guerrero-Castillo, S., Brandt, U., Decker, T., Lauria, I., independent of oxidative phosphorylation. Nat. Cell Biol. 17, 689–696.
Paggio, A., Rizzuto, R., Rugarli, E.I., De Stefani, D., Langer, T., 2016. The m-AAA Tsatmali, M., Walcott, E.C., Crossin, K.L., 2005. Newborn neurons acquire high levels of
protease associated with neurodegeneration limits MCU activity in mitochondria. reactive oxygen species and increased mitochondrial proteins upon differentiation
Mol. Cell 64, 148–162. from progenitors. Brain Res. 1040, 137–150.
Kosugi, A., Shearer, G.M., 1991. Effect of cyclosporin A on lymphopoiesis. III. Vega-Naredo, I., Loureiro, R., Mesquita, K.A., Barbosa, I.A., Tavares, L.C., Branco, A.F.,
Augmentation of the generation of natural killer cells in bone marrow transplanted Erickson, J.R., Holy, J., Perkins, E.L., Carvalho, R.A., Oliveira, P.J., 2014.
mice treated with cyclosporin A. J. Immunol. 146, 1416–1421. Mitochondrial metabolism directs stemness and differentiation in P19 embryonal
Kroemer, G., Galluzzi, L., Brenner, C., 2007. Mitochondrial membrane permeabilization carcinoma stem cells. Cell Death Differ. 21, 1560–1574.
in cell death. Physiol. Rev. 87, 99–163. Walters, A.M., Porter, G.A., Jr., Brookes, P.S., 2012. Mitochondria as a drug target in
Kwong, J.Q., Molkentin, J.D., 2015. Physiological and pathological roles of the ischemic heart disease and cardiomyopathy. Circ. Res. 111, 1222–1236.
mitochondrial permeability transition pore in the heart. Cell Metab. 21, 206–214. Wallace, D.C., 2005. A mitochondrial paradigm of metabolic and degenerative diseases,
Lim, S.Y., Hausenloy, D.J., Arjun, S., Price, A.N., Davidson, S.M., Lythgoe, M.F., Yellon, aging, and cancer: a dawn for evolutionary medicine. Annu. Rev. Genet. 39,
D.M., 2011. Mitochondrial cyclophilin-D as a potential therapeutic target for post- 359–407.
myocardial infarction heart failure. J. Cell. Mol. Med. 15, 2443–2451. Wang, W., Fang, H., Groom, L., Cheng, A., Zhang, W., Liu, J., Wang, X., Li, K., Han, P.,
Luo, M., Anderson, M.E., 2013. Mechanisms of altered Ca(2)(+) handling in heart Zheng, M., Yin, J., Wang, W., Mattson, M.P., Kao, J.P., Lakatta, E.G., Sheu, S.S.,
failure. Circ. Res. 113, 690–708. Ouyang, K., Chen, J., Dirksen, R.T., Cheng, H., 2008. Superoxide flashes in single
Maltecca, F., Baseggio, E., Consolato, F., Mazza, D., Podini, P., Young, S.M., Jr., Drago, I., mitochondria. Cell 134, 279–290.
Bahr, B.A., Puliti, A., Codazzi, F., Quattrini, A., Casari, G., 2015. Purkinje neuron Wang, X., Jian, C., Zhang, X., Huang, Z., Xu, J., Hou, T., Shang, W., Ding, Y., Zhang, W.,
Ca2+ influx reduction rescues ataxia in SCA28 model. J. Clin. Investig. 125, Ouyang, M., Wang, Y., Yang, Z., Zheng, M., Cheng, H., 2012. Superoxide flashes:
263–274. elemental events of mitochondrial ROS signaling in the heart. J. Mol. Cell. Cardiol.
Martin, L.J., Semenkow, S., Hanaford, A., Wong, M., 2014. Mitochondrial permeability 52, 940–948.
transition pore regulates Parkinson's disease development in mutant alpha-synuclein Wang, Z., Cai, F., Hu, L., Lu, Y., 2014. The role of mitochondrial permeability transition
transgenic mice. Neurobiol. Aging 35, 1132–1152. pore in regulating the shedding of the platelet GPIbalpha ectodomain. Platelets 25,
Maryanovich, M., Gross, A., 2013. A ROS rheostat for cell fate regulation. Trends Cell 373–381.
Biol. 23, 129–134. Yellon, D.M., Hausenloy, D.J., 2007. Myocardial reperfusion injury. N. Engl. J. Med. 357,
Mattson, M.P., Gleichmann, M., Cheng, A., 2008. Mitochondria in neuroplasticity and 1121–1135.

You might also like