6
Cell Calcium 52 (2012) 22–27 Contents lists available at SciVerse ScienceDirect Cell Calcium jo u rn al hom epa ge: www.elsevier.com/locate/ceca The permeability transition pore as a Ca 2+ release channel: New answers to an old question Paolo Bernardi , Sophia von Stockum Department of Biomedical Sciences and CNR Institute of Neuroscience, University of Padova, Padova, Italy a r t i c l e i n f o Article history: Received 16 February 2012 Received in revised form 21 March 2012 Accepted 21 March 2012 Available online 17 April 2012 Keywords: Mitochondria Permeability transition Ca 2+ release a b s t r a c t Mitochondria possess a sophisticated array of Ca 2+ transport systems reflecting their key role in physi- ological Ca 2+ homeostasis. With the exception of most yeast strains, energized organelles are endowed with a very fast and efficient mechanism for Ca 2+ uptake, the ruthenium red (RR)-sensitive mitochondrial Ca 2+ uniporter (MCU); and one main mechanism for Ca 2+ release, the RR-insensitive 3Na + –Ca 2+ antiporter. An additional mechanism for Ca 2+ release is provided by a Na + and RR-insensitive release mechanism, the putative 3H + –Ca 2+ antiporter. A potential kinetic imbalance is present, however, because the V max of the MCU is of the order of 1400 nmol Ca 2+ mg 1 protein min 1 while the combined V max of the efflux path- ways is about 20 nmol Ca 2+ mg 1 protein min 1 . This arrangement exposes mitochondria to the hazards of Ca 2+ overload when the rate of Ca 2+ uptake exceeds that of the combined efflux pathways, e.g. for sharp increases of cytosolic [Ca 2+ ]. In this short review we discuss the hypothesis that transient opening of the Ca 2+ -dependent permeability transition pore may provide mitocondria with a fast Ca 2+ release channel preventing Ca 2+ overload. We also address the relevance of a mitochondrial Ca 2+ release channel recently discovered in Drosophila melanogaster, which possesses intermediate features between the permeability transition pore of yeast and mammals. © 2012 Elsevier Ltd. All rights reserved. 1. Mitochondria have a large Ca 2+ problem In energized mitochondria Ca 2+ uptake is an electrophoretic process driven by the Ca 2+ electrochemical gradient, ˜ Ca: ˜ Ca = zF + RT ln [Ca 2+ ] i [Ca 2+ ] o (1) In respiring, coupled mitochondria the (negative inside) drives uptake of Ca 2+ , which is transported with a net charge of 2 [1,2] via an inner membrane channel [3], the mitochondrial Ca 2+ uniporter, MCU [4,5]. Ca 2+ uptake is charge-compensated by increased H + pumping by the respiratory chain [1,2], resulting in increased matrix pH that prevents the recovery of , limiting the further ability to accumulate Ca 2+ [6–8]. Uptake of substan- tial amounts of Ca 2+ therefore requires both buffering of matrix pH (to allow regeneration of the ); and buffering of matrix Ca 2+ (to prevent the buildup of a Ca 2+ concentration gradient) [6–8]. Buffering of matrix pH is achieved by the simultaneous uptake of protons and anions via diffusion of the undissociated acid through Corresponding author at: Department of Biomedical Sciences, University of Padova, Viale Giuseppe Colombo 3, I-35121 Padova, Italy. Tel.: +39 049 827 6365; fax: +39 049 827 6361. E-mail address: [email protected] (P. Bernardi). the inner membrane (as in the case of acetate), of CO 2 (which then regenerates bicarbonate and H + in the matrix) or through transport proteins (like the H + –Pi symporter) [9]. Buffering of accumulated Ca 2+ (and therefore the final [Ca 2+ ] in the matrix) thus depends in part on the cotransported anion and in part on remarkably ill- characterized matrix constituents. If Pi is the prevailing anion, free matrix [Ca 2+ ] becomes invariant with the matrix Ca 2+ load [10] and the ˜ Ca favors the accumulation of large loads of both Ca 2+ and Pi [11], with a predicted Ca 2+ equilibrium accumulation of 10 6 if the is 180 mV [6]. This is never reached because at rest- ing cytosolic Ca 2+ levels the rate of Ca 2+ uptake is comparable to that of the efflux pathways, and Ca 2+ distribution is governed by a kinetic steady state rather than by thermodynamic equilibrium [6,7]. Thus, in energized mitochondria coupling of Ca 2+ uptake with Ca 2+ efflux on separate pathways allows regulation of both cytoso- lic and matrix [Ca 2+ ]. Energy is required both for Ca 2+ uptake and for Ca 2+ release, owing to the electrophoretic nature of transport on MCU and the 3Na + –1Ca 2+ stoichiometry of NCLX [12], which dissipates the ; the requirement for H + extrusion posed by oper- ation of the Na + –H + exchanger (NHE) coupled to NCLX; and the fact that Ca 2+ efflux on the putative H + –Ca 2+ exchanger is favored by the [13,14], suggesting a H + –Ca 2+ stoichiometry higher than 2H + per Ca 2+ . A kinetic imbalance is apparent, however, because the V max of the MCU is of the order of 1400 nmol Ca 2+ mg 1 pro- tein min 1 while the combined V max of the efflux pathways is about 0143-4160/$ see front matter © 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ceca.2012.03.004

The permeability transition pore as a Ca2+ release channel: New answers to an old question

Embed Size (px)

Citation preview

Page 1: The permeability transition pore as a Ca2+ release channel: New answers to an old question

To

PD

a

ARRAA

KMPC

1

p

doCiitt((Bp

Pf

0h

Cell Calcium 52 (2012) 22– 27

Contents lists available at SciVerse ScienceDirect

Cell Calcium

jo u rn al hom epa ge: www.elsev ier .com/ locate /ceca

he permeability transition pore as a Ca2+ release channel: New answers to anld question

aolo Bernardi ∗, Sophia von Stockumepartment of Biomedical Sciences and CNR Institute of Neuroscience, University of Padova, Padova, Italy

r t i c l e i n f o

rticle history:eceived 16 February 2012eceived in revised form 21 March 2012ccepted 21 March 2012vailable online 17 April 2012

eywords:itochondria

ermeability transition

a b s t r a c t

Mitochondria possess a sophisticated array of Ca2+ transport systems reflecting their key role in physi-ological Ca2+ homeostasis. With the exception of most yeast strains, energized organelles are endowedwith a very fast and efficient mechanism for Ca2+ uptake, the ruthenium red (RR)-sensitive mitochondrialCa2+ uniporter (MCU); and one main mechanism for Ca2+ release, the RR-insensitive 3Na+–Ca2+ antiporter.An additional mechanism for Ca2+ release is provided by a Na+ and RR-insensitive release mechanism, theputative 3H+–Ca2+ antiporter. A potential kinetic imbalance is present, however, because the Vmax of theMCU is of the order of 1400 nmol Ca2+ mg−1 protein min−1 while the combined Vmax of the efflux path-ways is about 20 nmol Ca2+ mg−1 protein min−1. This arrangement exposes mitochondria to the hazards

2+ 2+

a2+ release of Ca overload when the rate of Ca uptake exceeds that of the combined efflux pathways, e.g. for sharp

increases of cytosolic [Ca2+]. In this short review we discuss the hypothesis that transient opening of theCa2+-dependent permeability transition pore may provide mitocondria with a fast Ca2+ release channelpreventing Ca2+ overload. We also address the relevance of a mitochondrial Ca2+ release channel recentlydiscovered in Drosophila melanogaster, which possesses intermediate features between the permeabilitytransition pore of yeast and mammals.

© 2012 Elsevier Ltd. All rights reserved.

. Mitochondria have a large Ca2+ problem

In energized mitochondria Ca2+ uptake is an electrophoreticrocess driven by the Ca2+ electrochemical gradient, � �Ca:

�Ca = zF�� + RT ln[Ca2+]i[Ca2+]o

(1)

In respiring, coupled mitochondria the � (negative inside)rives uptake of Ca2+, which is transported with a net chargef 2 [1,2] via an inner membrane channel [3], the mitochondriala2+ uniporter, MCU [4,5]. Ca2+ uptake is charge-compensated by

ncreased H+ pumping by the respiratory chain [1,2], resulting inncreased matrix pH that prevents the recovery of � , limitinghe further ability to accumulate Ca2+ [6–8]. Uptake of substan-ial amounts of Ca2+ therefore requires both buffering of matrix pHto allow regeneration of the � ); and buffering of matrix Ca2+

to prevent the buildup of a Ca2+ concentration gradient) [6–8].uffering of matrix pH is achieved by the simultaneous uptake ofrotons and anions via diffusion of the undissociated acid through

∗ Corresponding author at: Department of Biomedical Sciences, University ofadova, Viale Giuseppe Colombo 3, I-35121 Padova, Italy. Tel.: +39 049 827 6365;ax: +39 049 827 6361.

E-mail address: [email protected] (P. Bernardi).

143-4160/$ – see front matter © 2012 Elsevier Ltd. All rights reserved.ttp://dx.doi.org/10.1016/j.ceca.2012.03.004

the inner membrane (as in the case of acetate), of CO2 (which thenregenerates bicarbonate and H+ in the matrix) or through transportproteins (like the H+–Pi symporter) [9]. Buffering of accumulatedCa2+ (and therefore the final [Ca2+] in the matrix) thus dependsin part on the cotransported anion and in part on remarkably ill-characterized matrix constituents. If Pi is the prevailing anion, freematrix [Ca2+] becomes invariant with the matrix Ca2+ load [10]and the � �Ca favors the accumulation of large loads of both Ca2+

and Pi [11], with a predicted Ca2+ equilibrium accumulation of 106

if the � is −180 mV [6]. This is never reached because at rest-ing cytosolic Ca2+ levels the rate of Ca2+ uptake is comparable tothat of the efflux pathways, and Ca2+ distribution is governed bya kinetic steady state rather than by thermodynamic equilibrium[6,7]. Thus, in energized mitochondria coupling of Ca2+ uptake withCa2+ efflux on separate pathways allows regulation of both cytoso-lic and matrix [Ca2+]. Energy is required both for Ca2+ uptake andfor Ca2+ release, owing to the electrophoretic nature of transporton MCU and the 3Na+–1Ca2+ stoichiometry of NCLX [12], whichdissipates the � ; the requirement for H+ extrusion posed by oper-ation of the Na+–H+ exchanger (NHE) coupled to NCLX; and the factthat Ca2+ efflux on the putative H+–Ca2+ exchanger is favored by

the � [13,14], suggesting a H+–Ca2+ stoichiometry higher than2H+ per Ca2+. A kinetic imbalance is apparent, however, becausethe Vmax of the MCU is of the order of 1400 nmol Ca2+ mg−1 pro-tein min−1 while the combined Vmax of the efflux pathways is about
Page 2: The permeability transition pore as a Ca2+ release channel: New answers to an old question

/ Cell

2cuis

ewMs2ar(tgbeteov(C

2m

n3ccomifs

sMannbibosp(aoliittw[waii

P. Bernardi, S. von Stockum

0 nmol Ca2+ mg−1 protein min−1. This arrangement exposes mito-hondria to the hazards of Ca2+ overload when the rate of Ca2+

ptake exceeds that of the combined efflux pathways, e.g. for sharpncreases of cytosolic [Ca2+]. Why is then the rate of Ca2+ efflux solow?

The rate of Ca2+ uptake via the MCU is a steep function ofxtramitochondrial [Ca2+] [15]. Increasing rates of Ca2+ effluxould increase extramitochondrial Ca2+, stimulate Ca2+ uptake viaCU and increase overall Ca2+ cycling, resulting in energy dis-

ipation [16]. This can be observed by adding the electroneutralH+–Ca2+ ionophore A23187 to respiring mitochondria that haveccumulated Ca2+, a condition where Ca2+ is released and all of theespiratory capacity can be diverted into Ca2+ cycling [17]. Thusand as long as the membrane potential is high) net Ca2+ effluxhrough stimulation of the efflux pathways would have a high ener-etic cost. The low Vmax and early saturation of the efflux pathwaysy matrix Ca2+ are probably designed to pose an upper limit to thenergy that can be spent in regulation of matrix and cytosolic [Ca2+]hrough mitochondrial “Ca2+ cycling”. As mentioned above, how-ver, this situation exposes mitochondria to the constant threatf Ca2+ overload. We have proposed that this event may be pre-ented by transient openings of the permeability transition porePTP), which does mediate mitochondrial depolarization and fasta2+ release in vitro and possibly in vivo [18].

. Properties of the permeability transition pore inammals

The PTP is a high-conductance (≈1.3 nS) inner membrane chan-el [19–21]. In the fully open state its apparent diameter is about

nm [22] and its solute exclusion size ≈ 1500 Da. The PTP open-losed transitions are regulated by physiological effectors, and theonsequences of pore opening vary dramatically depending on thepen time [16]. Here we shall cover only the basic features of PTPodulation, and we refer the Reader to previous reviews for more

nformation on the rise of the PTP from the status of in vitro arti-act to that of effector mechanism of cell death regulated by keyignaling cascades [23–25].

Matrix modulators of the PTP include Ca2+ through a “permis-ive” site for opening that can be competitively inhibited by othere2+ ions like Mg2+, Sr2+ and Mn2+; and Pi, which in most species

cts as a powerful PTP inducer through a still undefined mecha-ism. Pore opening is promoted by an oxidized state of pyridineucleotides and of critical dithiols at discrete sites, both effectseing individually reversed by proper reductants [26]. Pore open-

ng can also cause production of reactive oxygen species, as showny the occurrence of “superoxide flashes” triggered by transientpenings of the PTP in cardiomyocytes [27]. The permeability tran-ition is strictly modulated by matrix pH with an optimum atH 7.4, while the open probability decreases both below pH 7.4through reversible protonation of critical histidyl residues [28,29])nd above pH 7.4 (through an unknown mechanism). Openingf the PTP is inhibited by cyclosporin (Cs) A after binding of theatter to cyclophilin (CyP) D, a matrix peptidyl-prolyl cis-transsomerase encoded by the Ppif gene that facilitates PTP open-ng [30–32]; indeed, ablation of CyPD approximately doubles thehreshold Ca2+ load required to open the PTP, which becomes iden-ical to that of CsA-treated, strain-matched wild type mitochondria,hile no effect of CsA is observed in CyPD-null mitochondria

33–36]. Major membrane effectors are the inside-negative � m,

hich tends to stabilize the PTP in the closed conformation [28];

nd electron flux within respiratory chain complex I, with anncreased open probability when flux increases [37]. The latter find-ng led to the discovery that the PT is regulated by quinones [38],

Calcium 52 (2012) 22– 27 23

possibly through a specific binding site whose occupancy affectsthe open-closed transitions depending on the bound species [39].

The only primary consequence of PTP opening is mitochondrialdepolarization. Unless single channel events are being recordedopenings of short duration may not be detected by potentiometricprobes. Since opening events are not synchronized for individ-ual mitochondria [40,41] in population studies they may also bemissed due to probe redistribution. We refer the reader to a seriesof specific studies about the occurrence of PTP openings of differentdurations in mitochondria in situ, and their consequences on cellviability [40–44]. For openings of longer duration depolarizationcan be easily measured both in isolated mitochondria and intactcells. As long as the pore is open collapse of the � �H prevents ATPsynthesis, and ATP hydrolysis by the mitochondrial ATPase wors-ens ATP depletion, which together with altered Ca2+ homeostasisis a key factor in various paradigms of cell death [45]. PersistentPTP opening is also followed by loss of matrix pyridine nucleotideswith respiratory inhibition [46], by equilibration of ion gradients,and by diffusion of solutes with molecular masses lower than about1500 Da and possible occurrence of swelling, cristae unfolding andouter membrane rupture. This, however, is not an inevitable con-sequence of PTP opening. Cristae remodeling due to PTP openingcan also occur in the absence of outer membrane rupture [47]with mobilization of the large pool of cytochrome c usually com-partmentalized in the intracristal spaces [48] allowing increasedrelease of cytochrome c through BAX/BAK channels in an other-wise intact outer membrane [47]. Furthermore, osmotic swellingrequires the existence of a colloidosmotic gradient between thematrix and the intermembrane space that may not exist (or maynot be large enough) to cause outer membrane rupture in situ.

The possible role of the PTP in physiological Ca2+ homeosta-sis has not been studied thoroughly, in part because the natureof the PTP has remained elusive [23] and therefore modulationof the “channel” itself by genetic methods has not been possible;in part because true PTP blockers have not been developed. Greathopes were raised by the discovery that the PTP is inhibited by CsA[49–51], but it is now clear that CyPD is a key modulator but notan obligatory component of the PTP. It should therefore be bornein mind that Ppif−/− (CyPD-null) mice are not PTP-null mice, as apermeability transition can still occur, e.g. after an increased Ca2+

load [33–36]. This consideration has dramatic implications for theinterpretation of results obtained with CsA; similar to the absenceof CyPD, CsA can desensitize but not block the PTP, and thereforelack of sensitivity to CsA does not necessarily mean that the PTPis not involved in the event under study. Little attention has alsobeen paid to the fact that expression of CyPD can be modulated(e.g. by muscle denervation [52]), and that only CyPD-expressingmitochondria are expected to respond to CsA. A further element ofcomplexity is that CsA also affects mitochondria in situ by prevent-ing calcineurin-dependent dephosphorylation of the pro-fissionprotein DRP1, which is essential for its translocation to mitochon-dria [53]. The resulting inhibition of mitochondrial fission by CsAmay thus be a cytoprotective event that is independent of inhibitionof CyPD and desensitization of the PTP [53], a finding that shouldinduce some caution in evaluating in situ and in vivo studies basedonly on CsA rather than on genetic ablation of CyPD or on the useof CyP inhibitors devoid of effects on calcineurin [54–56].

3. The mitochondrial permeability transition pore as a Ca2+

release channel: is lack of selectivity a problem or anadvantage?

If a Ca2+ concentration gradient exists between the matrix andthe external medium (or the cytosol) PTP opening leads to Ca2+

release. In 1996 we have proposed that the PTP may serve as a

Page 3: The permeability transition pore as a Ca2+ release channel: New answers to an old question

2 / Cell Calcium 52 (2012) 22– 27

mnprvatcmrCimotCmss(CrteutrecCacmb

4m

byaYteo[naib(rDrs

csetteDt

Table 1Properties of mammalian and yeast PTP, and of the Drosophila mCRC.

Mammals Yeast Drosophila

Permeability tosolutes up to≈1500 Da

Yes Yes No

Selective Ca2+

releaseNo No Yes

Matrix Ca2+ Required May be required RequiredMatrix Pi Inducer Inhibitor Inhibitormt Cyp Yes Yes NoCsA Inhibitor No effect No effect

4 P. Bernardi, S. von Stockum

itochondrial Ca2+ release channel [18], and a specific point thateeds to be discussed is whether the lack of selectivity is a realroblem, or rather an essential feature that allows fast and effectiveelease of matrix Ca2+. Ca2+ efflux down its concentration gradientia a Ca2+-selective channel would be opposed by the buildup of

Ca2+ diffusion potential [57]. According to the Nernst equation,he magnitude of the Ca2+ diffusion potential when no charge-ompensating species are present is −30 mV per decade of theatrix/cytosol Ca2+ concentration ratio. This diffusion potential is

educed to less negative values (with corresponding increase of thea2+ efflux rate) by any charge-compensating current (influx of pos-

tive charges, efflux of negative charges, or both). Since the innerembrane has a very low permeability to charged species, the rate

f Ca2+ efflux would be extremely slow and essentially limited byhe H+ permeability. In other words, to obtain a significant rate ofa2+ efflux via a Ca2+-selective channel the inner membrane per-eability should be increased as well. An unselective pore of large

ize like the PTP confers the advantage of providing charge compen-ation within the channel itself, thus allowing maximal Ca2+ fluxi.e. Ca2+ release would occur at zero potential). This would makea2+ release possible even for vanishingly small [Ca2+] gradients,egulation being achieved through modulation of the pore openime. It should be noted that no K+ and Na+ concentration gradientsxist across the inner membrane because the slow electrophoreticptake of K+ and Na+ is compensated by the K+–H+ exchanger andhe NHE, respectively. Thus, PTP opening can lead to selective Ca2+

elease without perturbation of K+ and Na+ homeostasis, and novolutionary pressure may have existed for the development ofation selectivity. In this respect the PTP is strikingly similar to thea2+ release channel of the sarcoplasmic reticulum, which oper-tes as a Ca2+-selective channel despite its large size (>38 A), highonductance for monovalent cations (≈1 nS at saturating K+), per-eability to solutes like glucose and low permeability ratio when

oth K+ and Ca2+ are present (PCa/PK ≈ 6) [58].

. A Ca2+ release channel in Drosophila mitochondria: theissing link between the PTP of yeast and mammals?

We have recently argued that, in spite of clear differencesetween species, the PTP has been conserved in evolution fromeast to mammals [59]. Yeast mitochondria do possess a perme-bility pathway that resembles the mammalian PTP, also called theeast Mitochondrial Unselective Channel (YMUC) [60,61]. Based onhe features of the PTP of yeast and mammals as of 1998, Manont al. concluded that regulation of YMUC is too different from thatf the mammalian PTP for yeast to be a good model of the latter60]; in the light of new results we believe that the differences areot as fundamental as suspected earlier, and we refer the reader to

recent review specifically devoted to this aspect [59]. Other thann mammals and yeast, occurrence of a permeability transition haseen established in plants, amphibians and fish including zebrafishDanio rerio) [62] (see [59] for references). A mitochondrial Ca2+

elease channel (mCRC) that we recently identified in cells fromrosophila melanogaster [63] deserves a special mention, as it may

epresent an evolutionary variant of the PTP displaying remarkableelectivity toward Ca2+ and H+.

Mitochondria were studied in digitonin-permeabilized S2R+

ells [63], which are derived from Drosophila late embryonictages and represent a variety of tissue precursors [64]. Of inter-st, mitochondria of S2R+ cells possess all the classical Ca2+

ransport pathways found in mammalian mitochondria, i.e. (i)

he RR-sensitive Ca2+ uniporter (which is consistent with thexistence of orthologs of MCU [4,5] and of MICU1 [65] in therosophila genome); (ii) the Na+–Ca2+ antiporter recently iden-

ified as NCLX [12], whose ortholog also exists in Drosophila;

Tetracaine Inhibitor Not tested InhibitorRedox sites Yes Yes Yes

(iii) the RR-insensitive putative H+–Ca2+ antiporter mediating Ca2+

release at high membrane potential; (iv) a tetracaine-sensitive, RR-insensitive Ca2+ release pathway that opens in response to matrixCa2+ loading or to depolarization, and mediates Ca2+ release [63].The properties of the Drosophila mCRC appear to be intermediatebetween those of the PTP of mammals and yeast. Like the mam-malian PTP, the Drosophila Ca2+ release pathway is inhibited bytetracaine [66] and opens in response to matrix Ca2+ loading, innermembrane depolarization, thiol oxidation and treatment with rel-atively high concentrations of NEM [63]. Like the yeast PTP (and atstriking variance from the mammalian pore) the Drosophila mCRCis inhibited rather than stimulated by Pi; and is insensitive to ADP,quinones and to CsA [63], a finding that matches the absence ofa mitochondrial CyP in Drosophila. Together with the inhibitoryeffect of Pi, the most striking difference between the DrosophilamCRC and the PTP is the selectivity for the transported species. Atthe onset of Ca2+-dependent Ca2+ release Drosophila mitochondriaundergo depolarization, suggesting that the putative channel is alsopermeable to H+; yet no matrix swelling is observed even in KCl-based medium, indicating that the channel is not permeable to K+

(and to Cl−) in spite of the fact that the hydrated radius of Ca2+ islarger than that of K+ [63]. Lack of swelling was confirmed by lackof cytochrome c release and by ultrastructural analysis, and wasnot due to peculiar structural features of Drosophila mitochondriabecause matrix swelling and cytochrome c release readily followedthe addition of the pore-forming peptide alamethicin [63].

The key features of the yeast and mammalian PTP, and of theDrosophila mCRC are summarized in Table 1. The most remarkabledifferences are the permeability to solutes, the selectivity for Ca2+

and the effects of Pi. Since nearly all yeast strains lack an MCU,the Ca2+-dependence of the yeast PTP has not been easy to assessalthough it was known that the yeast PT is favored by added Ca2+

[67]. Recent work using the Ca2+ ionophore ETH129, which medi-ates electrophoretic Ca2+ transport, has demonstrated that the PTPis favored by Ca2+ uptake in mitochondria from Saccharomyces cere-visiae [68], and based on this finding we suggest that matrix Ca2+

may be required for onset of the PT in yeast as well (Table 1). Lack ofinhibition of Drosophila mCRC by CsA can be easily explained by thelack of a mitochondrial CyP. On the other hand, since a matrix CyPis found in yeast and its enzymatic activity is inhibited by CsA [69]we suspect that the CyP ability to interact with the PTP occurredlater in evolution.

5. The PTP as a mitochondrial Ca2+ release channel: only ahypothesis?

In 1992, Altschuld et al. demonstrated that CsA significantly2+ 2+

increases net Ca uptake and decreases Ca efflux in isolated

cardiomyocytes, as measured by radiolabeled 45Ca2+, without hav-ing any impact on cell morphology or viability [70]. The effect ofCsA was concentration-dependent and specific to mitochondria, as

Page 4: The permeability transition pore as a Ca2+ release channel: New answers to an old question

/ Cell

Anatpcmhnwwftir

p[wobriootnrfowNesctoo

a[pitndmduttawcooapddssciCl

[

[

[

[

[

[

[

[

[

P. Bernardi, S. von Stockum

TP-dependent Ca2+ uptake by the sarcoplasmatic reticulum wasot affected. The selectivity of the effects for mitochondrial Ca2+

nd the very short incubation time with CsA (15 min) suggest thathe PTP was being affected, and these data may represent the firstiece of evidence that the pore contributes to Ca2+ cycling in mito-hondria of living cardiomyocytes, and that reversible pore openingay be a physiological process in heart cells [70]. On the other

and, Eriksson et al. found that fluxes of Ca2+, Mg2+ and adenineucleotides in perfused rat livers following hormonal stimulationere unaffected by previous administration of CsA, even if the PTPas demonstrably desensitized by CsA in mitochondria isolated

rom the same CsA-perfused livers [71]. The Authors concludedhat regulation of mitochondrial ion and metabolite homeostasiss independent of the PTP [71], yet as discussed above a negativeesult is not as informative.

Two recent publications based on Ppif−/− cells and mice dorovide clear support for a role of the PTP in Ca2+ homeostasis72,73]. Adult cortical neurons from wild type and Ppif−/− miceere treated with either ATP (to activate P2Y purinergic receptors)

r with depolarizing concentrations of KCl (to open plasma mem-rane voltage-dependent Ca2+ channels), both stimuli causing aobust increase of both cytosolic and mitochondrial [Ca2+] that wasndistinguishable in neurons of the two genotypes [73]. Applicationf the two stimuli together, however, resulted in much higher levelsf mitochondrial [Ca2+] in the Ppif−/− neurons, suggesting that thehreshold for PTP activation had been reached in the wild type butot in the CyPD-null mitochondria in situ. Thus, it appears that theegulatory role of CyPD (and of PTP opening) becomes crucial onlyor relatively large mitochondrial Ca2+ loads exceeding the capacityf NCLX and of the Na+-insensitive release pathway [73]. In otherords, the PTP could be silent unless high Ca2+ loads saturate theCLX and the H+–Ca2+ exchanger, allowing matrix [Ca2+] to risenough to trigger pore opening. It should be noted that the tran-ient stimulation of these Ca2+-mobilizing pathways did not induceell death either in wild type or in CyPD-null neurons, suggestinghat the mitochondrial PTP-activating response was part of a physi-logical process that is consistent with occurrence of reversible PTPpening [42].

In adult mice ablation of CyPD increases resistance tocute ischemia-reperfusion injury both in the heart and brain33,35,36,74], while in neonatal mice, where the membrane-ermeabilizing effects of BAX predominate, Ppif−/− mice were

nstead remarkably sensitized suggesting age-related changes inhe mitochondrial response to injury [74]. An age-related phe-otype was also discovered in the hearts of Ppif−/− mice, whichisplayed an intriguing decrease of maximum contractile reserveatched by increased shortening and relaxation times with longer

ecay of cytosolic Ca2+ transients [72]. Ppif−/− mice were alsonable to compensate for the increase in afterload caused byransaortic constriction, displaying a larger reduction in frac-ional shortening, decompensation, ventricular dilation, fibrosisnd congestive heart failure; all consequences of CyPD ablationere cured by heart-selective reexpression of CyPD, strongly indi-

ating that the maladaptive phenotype of Ppif−/− mice dependsn a primary disturbance of myocyte mitochondria rather thann an underlying systemic response [72]. Metabolic in vivonalysis demonstrated a significant increase in the glucose toalmitate ratio, suggesting a metabolic shift from fatty acid oxi-ation to glycolysis associated with increased activity of pyruvateehydrogenase and �-ketoglutarate dehydrogenase. Direct mea-urement of total mitochondrial Ca2+ content of Ppif−/− heartshowed a 2.6-fold increase, which was matched by greater mito-

hondrial Ca2+ transients in myocytes treated with CsA. Verymportantly, under continuous pacing PTP desensitization withsA decreased the rise time in Ca2+ accumulation and pro-

onged the recovery time after pacing, findings that are entirely

[

Calcium 52 (2012) 22– 27 25

consistent with the PTP acting as a Ca2+ release channel to preventCa2+ overload [72].

6. Summary and conclusions

The hypothesis that transient opening of the PTP may servethe physiological function of regulating matrix Ca2+ by prevent-ing Ca2+ overload appears theoretically justified, and supported bya limited but extremely solid set of experimental results. Confu-sion may have arisen from the interpretation of results based onthe use of CsA as well as of Ppif−/− mice, which have too oftenbeen erroneously considered null for the PTP as well. Much morework is obviously needed, and we hope that this short reviewwill help rekindle interest and experimental testing of this sub-ject.

Acknowledgments

Work in our laboratory is supported by funds from the MIUR(FIRB and PRIN), Telethon Grants GPP10005 and GGP11082, AIRCInvestigator Grant 8722, the University of Padova and the Fon-dazione Cariparo.

References

[1] A. Scarpa, G.F. Azzone, The mechanism of ion translocation in mitochon-dria. 4. Coupling of K+ efflux with Ca2+ uptake, Eur. J. Biochem. 12 (1970)328–335.

[2] D.E. Wingrove, J.M. Amatruda, T.E. Gunter, Glucagon effects on the membranepotential and calcium uptake rate of rat liver mitochondria, J. Biol. Chem. 259(1984) 9390–9394.

[3] Y. Kirichok, G. Krapivinsky, D.E. Clapham, The mitochondrial calcium uniporteris a highly selective ion channel, Nature 427 (2004) 360–364.

[4] D. De Stefani, A. Raffaello, E. Teardo, I. Szabó, R. Rizzuto, A forty-kilodaltonprotein of the inner membrane is the mitochondrial calcium uniporter, Nature476 (2011) 336–340.

[5] J.M. Baughman, F. Perocchi, H.S. Girgis, M. Plovanich, C.A. Belcher-Timme,Y. Sancak, X.R. Bao, L. Strittmatter, O. Goldberger, R.L. Bogorad, V. Kotelian-sky, V.K. Mootha, Integrative genomics identifies MCU as an essentialcomponent of the mitochondrial calcium uniporter, Nature 476 (2011)341–345.

[6] G.F. Azzone, T. Pozzan, S. Massari, M. Bragadin, P. Dell’Antone, H+/site ratioand steady state distribution of divalent cations in mitochondria, FEBS Lett. 78(1977) 21–24.

[7] D.G. Nicholls, The regulation of extramitochondrial free calcium ion concen-tration by rat liver mitochondria, Biochem. J. 176 (1978) 463–474.

[8] P. Bernardi, D. Pietrobon, On the nature of Pi-induced, Mg2+-prevented Ca2+

release in rat liver mitochondria, FEBS Lett. 139 (1982) 9–12.[9] P. Mitchell, Keilin’s respiratory chain concept and its chemiosmotic conse-

quences, Science 206 (1979) 1148–1159.10] S. Chalmers, D.G. Nicholls, The relationship between free and total calcium

concentrations in the matrix of liver and brain mitochondria, J. Biol. Chem. 278(2003) 19062.

11] E.J. Harris, B. Zaba, The phosphate requirement for Ca2+-uptake by heart andliver mitochondria, FEBS Lett. 79 (1977) 284–290.

12] R. Palty, W.F. Silverman, M. Hershfinkel, T. Caporale, S.L. Sensi, J. Parnis, C. Nolte,D. Fishman, V. Shoshan-Barmatz, S. Herrmann, D. Khananshvili, I. Sekler, NCLXis an essential component of mitochondrial Na+/Ca2+ exchange, Proc. Natl. Acad.Sci. U.S.A. 107 (2010) 436–441.

13] P. Bernardi, G.F. Azzone, A membrane potential-modulated pathway for Ca2+

efflux in rat liver mitochondria, FEBS Lett. 139 (1982) 13–16.14] P. Bernardi, G.F. Azzone, Regulation of Ca2+ efflux in rat liver mitochondria. Role

of membrane potential, Eur. J. Biochem. 134 (1983) 377–383.15] M. Bragadin, T. Pozzan, G.F. Azzone, Kinetics of Ca2+ carrier in rat liver mito-

chondria, Biochemistry 18 (1979) 5972–5978.16] P. Bernardi, Mitochondrial transport of cations: channels, exchangers, and per-

meability transition, Physiol. Rev. 79 (1999) 1127–1155.17] G.M Heaton, D.G. Nicholls, The calcium conductance of the inner membrane of

rat liver mitochondria and the determination of the calcium electrochemicalgradient, Biochem. J. 156 (1976) 635–646.

18] P. Bernardi, V. Petronilli, The permeability transition pore as a mitochondrial

calcium release channel: a critical appraisal, J. Bioenerg. Biomembr. 28 (1996)131–138.

19] V. Petronilli, I. Szabó, M. Zoratti, The inner mitochondrial membrane containsion-conducting channels similar to those found in bacteria, FEBS Lett. 259(1989) 137–143.

Page 5: The permeability transition pore as a Ca2+ release channel: New answers to an old question

2 / Cell

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

6 P. Bernardi, S. von Stockum

20] K.W. Kinnally, M.L. Campo, H. Tedeschi, Mitochondrial channel activitystudied by patch-clamping mitoplasts, J. Bioenerg. Biomembr. 21 (1989)497–506.

21] I. Szabó, M. Zoratti, The giant channel of the inner mitochondrial membrane isinhibited by cyclosporin A, J. Biol. Chem. 266 (1991) 3376–3379.

22] S. Massari, G.F. Azzone, The equivalent pore radius of intact and damaged mito-chondria and the mechanism of active shrinkage, Biochim. Biophys. Acta 283(1972) 23–29.

23] P. Bernardi, A. Krauskopf, E. Basso, V. Petronilli, E. Blachly-Dyson, F. Di Lisa,M.A. Forte, The mitochondrial permeability transition from in vitro artifact todisease target, FEBS J. 273 (2006) 2077–2099.

24] A. Rasola, M. Sciacovelli, B. Pantic, P. Bernardi, Signal transduction to the per-meability transition pore, FEBS Lett. 584 (2010) 1989–1996.

25] A. Rasola, P. Bernardi, Mitochondrial permeability transition in Ca2+-dependentapoptosis and necrosis, Cell Calcium 50 (2011) 222–233.

26] P. Costantini, B.V. Chernyak, V. Petronilli, P. Bernardi, Modulation of the mito-chondrial permeability transition pore by pyridine nucleotides and dithioloxidation at two separate sites, J. Biol. Chem. 271 (1996) 6746–6751.

27] W. Wang, H. Fang, L. Groom, A. Cheng, W. Zhang, J. Liu, X. Wang, K. Li, P. Han, M.Zheng, J. Yin, W. Wang, M.P. Mattson, J.P. Kao, E.G. Lakatta, S.S. Sheu, K. Ouyang,J. Chen, R.T. Dirksen, H. Cheng, Superoxide flashes in single mitochondria, Cell134 (2008) 279–290.

28] P. Bernardi, Modulation of the mitochondrial cyclosporin A-sensitive perme-ability transition pore by the proton electrochemical gradient. Evidence thatthe pore can be opened by membrane depolarization, J. Biol. Chem. 267 (1992)8834–8839.

29] A. Nicolli, V. Petronilli, P. Bernardi, Modulation of the mitochondrial cyclosporinA-sensitive permeability transition pore by matrix pH. Evidence that the poreopen-closed probability is regulated by reversible histidine protonation, Bio-chemistry 32 (1993) 4461–4465.

30] A.P. Halestrap, A.M. Davidson, Inhibition of Ca2+-induced large-amplitudeswelling of liver and heart mitochondria by cyclosporin is probably causedby the inhibitor binding to mitochondrial-matrix peptidyl-prolyl cis-trans iso-merase and preventing it interacting with the adenine nucleotide translocase,Biochem. J. 268 (1990) 153–160.

31] E.J. Griffiths, A.P. Halestrap, Further evidence that cyclosporin A protectsmitochondria from calcium overload by inhibiting a matrix peptidyl-prolyl cis-trans isomerase. Implications for the immunosuppressive and toxic effects ofcyclosporin, Biochem. J. 274 (1991) 611–614.

32] A. Nicolli, E. Basso, V. Petronilli, R.M. Wenger, P. Bernardi, Interactions ofcyclophilin with the mitochondrial inner membrane and regulation of the per-meability transition pore, a cyclosporin A-sensitive channel, J. Biol. Chem. 271(1996) 2185–2192.

33] C.P. Baines, R.A. Kaiser, N.H. Purcell, N.S. Blair, H. Osinska, M.A. Hambleton, E.W.Brunskill, M.R. Sayen, R.A. Gottlieb, G.W. Dorn, J. Robbins, J.D. Molkentin, Lossof cyclophilin D reveals a critical role for mitochondrial permeability transitionin cell death, Nature 434 (2005) 658–662.

34] E. Basso, L. Fante, J. Fowlkes, V. Petronilli, M.A. Forte, P. Bernardi, Propertiesof the permeability transition pore in mitochondria devoid of cyclophilin D, J.Biol. Chem. 280 (2005) 18558–18561.

35] T. Nakagawa, S. Shimizu, T. Watanabe, O. Yamaguchi, K. Otsu, H. Yamagata, H.Inohara, T. Kubo, Y. Tsujimoto, Cyclophilin D-dependent mitochondrial perme-ability transition regulates some necrotic but not apoptotic cell death, Nature434 (2005) 652–658.

36] A.C. Schinzel, O. Takeuchi, Z. Huang, J.K. Fisher, Z. Zhou, J. Rubens, C. Hetz,N.N. Danial, M.A. Moskowitz, S.J. Korsmeyer, Cyclophilin D is a compo-nent of mitochondrial permeability transition and mediates neuronal celldeath after focal cerebral ischemia, Proc. Natl. Acad. Sci. U.S.A. 102 (2005)12005–12010.

37] E. Fontaine, O. Eriksson, F. Ichas, P. Bernardi, Regulation of the permeabil-ity transition pore in skeletal muscle mitochondria. Modulation by electronflow through the respiratory chain complex I, J. Biol. Chem. 273 (1998)12662–12668.

38] E. Fontaine, F. Ichas, P. Bernardi, A ubiquinone-binding site regulatesthe mitochondrial permeability transition pore, J. Biol. Chem. 273 (1998)25734–25740.

39] L. Walter, H. Miyoshi, X. Leverve, P. Bernardi, E. Fontaine, Regulation of themitochondrial permeability transition pore by ubiquinone analogs. A progressreport, Free Radic. Res. 36 (2002) 405–412.

40] J. Hüser, C.E. Rechenmacher, L.A. Blatter, Imaging the permeability pore tran-sition in single mitochondria, Biophys. J. 74 (1998) 2129–2137.

41] J. Hüser, L.A. Blatter, Fluctuations in mitochondrial membrane potential causedby repetitive gating of the permeability transition pore, Biochem. J. 343 (Pt 2)(1999) 311–317.

42] V. Petronilli, G. Miotto, M. Canton, M. Brini, R. Colonna, P. Bernardi, F. Di Lisa,Transient and long-lasting openings of the mitochondrial permeability transi-tion pore can be monitored directly in intact cells by changes in mitochondrialcalcein fluorescence, Biophys. J. 76 (1999) 725–734.

43] V. Petronilli, D. Penzo, L. Scorrano, P. Bernardi, F. Di Lisa, The mitochondrialpermeability transition, release of cytochrome c and cell death. Correla-tion with the duration of pore openings in situ, J. Biol. Chem. 276 (2001)

12030–12034.

44] L. Scorrano, D. Penzo, V. Petronilli, F. Pagano, P. Bernardi, Arachidonic acidcauses cell death through the mitochondrial permeability transition. Implica-tions for tumor necrosis factor-� apoptotic signaling, J. Biol. Chem. 276 (2001)12035–12040.

[

Calcium 52 (2012) 22– 27

45] P. Bernardi, V. Petronilli, F. Di Lisa, M. Forte, A mitochondrial perspective oncell death, Trends Biochem. Sci. 26 (2001) 112–117.

46] A. Vinogradov, A. Scarpa, B. Chance, Calcium and pyridine nucleotide inter-action in mitochondrial membranes, Arch. Biochem. Biophys. 152 (1972)646–654.

47] L. Scorrano, M. Ashiya, K. Buttle, S. Weiler, S.A. Oakes, C.A. Mannella, S.J.Korsmeyer, A distinct pathway remodels mitochondrial cristae and mobilizescytochrome c during apoptosis, Dev. Cell 2 (2002) 55–67.

48] P. Bernardi, G.F. Azzone, Cytochrome c as an electron shuttle between the outerand inner mitochondrial membranes, J. Biol. Chem. 256 (1981) 7187–7192.

49] N. Fournier, G. Ducet, A. Crevat, Action of cyclosporine on mitochondrial cal-cium fluxes, J. Bioenerg. Biomembr. 19 (1987) 297–303.

50] M. Crompton, H. Ellinger, A. Costi, Inhibition by cyclosporin A of a Ca2+-dependent pore in heart mitochondria activated by inorganic phosphate andoxidative stress, Biochem. J. 255 (1988) 357–360.

51] K.M. Broekemeier, M.E. Dempsey, D.R. Pfeiffer, Cyclosporin A is a potentinhibitor of the inner membrane permeability transition in liver mitochondria,J. Biol. Chem. 264 (1989) 7826–7830.

52] K. Csukly, A. Ascah, J. Matas, P.F. Gardiner, E. Fontaine, Y. Burelle, Muscle den-ervation promotes opening of the permeability transition pore and increasesthe expression of cyclophilin D, J. Physiol. 574 (2006) 319–327.

53] G.M. Cereghetti, A. Stangherlin, O. Martins de Brito, C.R. Chang, C. Black-stone, P. Bernardi, L. Scorrano, Dephosphorylation by calcineurin regulatestranslocation of Drp1 to mitochondria, Proc. Natl. Acad. Sci. U.S.A. 105 (2008)15803–15808.

54] P.C. Waldmeier, J.J. Feldtrauer, T. Qian, J.J. Lemasters, Inhibition of the mito-chondrial permeability transition by the nonimmunosuppressive cyclosporinderivative NIM811, Mol. Pharmacol. 62 (2002) 22–29.

55] M.J. Hansson, G. Mattiasson, R. Mansson, J. Karlsson, M.F. Keep, P. Waldmeier,U.T. Ruegg, J.M. Dumont, K. Besseghir, E. Elmer, The nonimmunosuppressivecyclosporin analogs NIM811 and UNIL025 display nanomolar potencies on per-meability transition in brain-derived mitochondria, J. Bioenerg. Biomembr. 36(2004) 407–413.

56] T. Tiepolo, A. Angelin, E. Palma, P. Sabatelli, L. Merlini, L. Nicolosi, F. Finetti,P. Braghetta, G. Vuagniaux, J.M. Dumont, C.T. Baldari, P. Bonaldo, P. Bernardi,The cyclophilin inhibitor Debio 025 normalizes mitochondrial function, mus-cle apoptosis and ultrastructural defects in Col6a1−/− myopathic mice, Br. J.Pharmacol. 157 (2009) 1045–1052.

57] K.E.O. Akerman, Changes in membrane potential during calcium ion influx andefflux across the mitochondrial membrane, Biochim. Biophys. Acta: Bioenerg502 (1978) 359–366.

58] J.S. Smith, T. Imagawa, J. Ma, M. Fill, K.P. Campbell, R. Coronado, Purified ryan-odine receptor from rabbit skeletal muscle is the calcium-release channel ofsarcoplasmic reticulum, J. Gen. Physiol. 92 (1988) 1–26.

59] L. Azzolin, S. von Stockum, E. Basso, V. Petronilli, M.A. Forte, P. Bernardi, Themitochondrial permeability transition from yeast to mammals, FEBS Lett. 584(2010) 2504–2509.

60] S. Manon, X. Roucou, M. Guerin, M. Rigoulet, B. Guerin, Characterizationof the yeast mitochondria unselective channel: a counterpart to the mam-malian permeability transition pore? J. Bioenerg. Biomembr. 30 (1998)419–429.

61] S. Uribe-Carvajal, L.A. Luévano-Martìnez, S. Guerrero-Castillo, A. Cabrera-Orefice, N.A. Corona-de-la-Pena, M. Gutiérrez-Aguilar, Mitochondrial unselec-tive channels throughout the eukaryotic domain, Mitochondrion 11 (2011)382–390.

62] L. Azzolin, E. Basso, F. Argenton, P. Bernardi, Mitochondrial Ca2+ transport andpermeability transition in zebrafish (Danio rerio), Biochim. Biophys. Acta 1797(2010) 1775–1779.

63] S. von Stockum, E. Basso, V. Petronilli, P. Sabatelli, M.A. Forte, P. Bernardi, Prop-erties of Ca2+ Transport in Mitochondria of Drosophila melanogaster, J. Biol.Chem. 286 (2011) 41163–41170.

64] I. Schneider, Cell lines derived from late embryonic stages of Drosophilamelanogaster, J. Embryol. Exp. Morphol. 27 (1972) 353–365.

65] F. Perocchi, V.M. Gohil, H.S. Girgis, X.R. Bao, J.E. McCombs, A.E. Palmer, V.K.Mootha, MICU1 encodes a mitochondrial EF hand protein required for Ca2+

uptake, Nature 467 (2010) 291–296.66] A.P. Dawson, M.J. Selwyn, D.V. Fulton, Inhibition of Ca2+ efflux from mitochon-

dria by nupercaine and tetracaine, Nature 277 (1979) 484–486.67] A.J. Kowaltowski, A.E. Vercesi, S.G. Rhee, L.E. Netto, Catalases and thioredoxin

peroxidase protect Saccharomyces cerevisiae against Ca2+-induced mitochon-drial membrane permeabilization and cell death, FEBS Lett. 473 (2000)177–182.

68] A. Yamada, T. Yamamoto, Y. Yoshimura, S. Gouda, S. Kawashima, N. Yamazaki,K. Yamashita, M. Kataoka, T. Nagata, H. Terada, D.R. Pfeiffer, Y. Shinohara, Ca2+-induced permeability transition can be observed even in yeast mitochondriaunder optimized experimental conditions, Biochim. Biophys. Acta 1787 (2009)1486–1491.

69] M. Tropschug, I.B. Barthelmess, W. Neupert, Sensitivity to cyclosporin A is medi-ated by cyclophilin in Neurospora crassa and Saccharomyces cerevisiae, Nature342 (1989) 953–955.

70] R.A. Altschuld, C.M. Hohl, L.C. Castillo, A.A. Garleb, R.C. Starling, G.P. Brierley,

Cyclosporin inhibits mitochondrial calcium efflux in isolated adult rat ventric-ular cardiomyocytes, Am. J. Physiol. 262 (1992) H1699–H1704.

71] O. Eriksson, P. Pollesello, E. Geimonen, Regulation of total mitochondrial Ca2+ inperfused liver is independent of the permeability transition pore, Am. J. Physiol.276 (1999) C1297–C1302.

Page 6: The permeability transition pore as a Ca2+ release channel: New answers to an old question

/ Cell

[

[

831–842.

P. Bernardi, S. von Stockum

72] J.W. Elrod, R. Wong, S. Mishra, R.J. Vagnozzi, B. Sakthievel, S.A. Goonasek-era, J. Karch, S. Gabel, J. Farber, T. Force, J.H. Brown, E. Murphy,

J.D. Molkentin, Cyclophilin D controls mitochondrial pore-dependent Ca2+

exchange, metabolic flexibility, and propensity for heart failure in mice, J. Clin.Investig. 120 (2010) 3680–3687.

73] A. Barsukova, A. Komarov, G. Hajnoczky, P. Bernardi, D. Bourdette, M. Forte,Activation of the mitochondrial permeability transition pore modulates Ca2+

[

Calcium 52 (2012) 22– 27 27

responses to physiological stimuli in adult neurons, Eur. J. Neurosci. 33 (2011)

74] X. Wang, Y. Carlsson, E. Basso, C. Zhu, C.I. Rousset, A. Rasola, B.R. Johansson, K.Blomgren, C. Mallard, P. Bernardi, M.A. Forte, H. Hagberg, Developmental shiftof cyclophilin D contribution to hypoxic-ischemic brain injury, J. Neurosci. 29(2009) 2588–2596.