You are on page 1of 42

Physiol Rev 94: 909 –950, 2014

doi:10.1152/physrev.00026.2013

MITOCHONDRIAL REACTIVE OXYGEN SPECIES


(ROS) AND ROS-INDUCED ROS RELEASE
Dmitry B. Zorov, Magdalena Juhaszova, and Steven J. Sollott

A. N. Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, Moscow, Russia;
and Laboratory of Cardiovascular Science, National Institute on Aging, National Institutes of Health, Baltimore,
Maryland

Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial Reactive Oxygen Species

L
(ROS) and ROS-Induced ROS Release. Physiol Rev 94: 909 –950, 2014;
doi:10.1152/physrev.00026.2013.—Byproducts of normal mitochondrial metabo-
lism and homeostasis include the buildup of potentially damaging levels of reactive
oxygen species (ROS), Ca2⫹, etc., which must be normalized. Evidence suggests that
brief mitochondrial permeability transition pore (mPTP) openings play an important physiological
role maintaining healthy mitochondria homeostasis. Adaptive and maladaptive responses to redox
stress may involve mitochondrial channels such as mPTP and inner membrane anion channel
(IMAC). Their activation causes intra- and intermitochondrial redox-environment changes leading to
ROS release. This regenerative cycle of mitochondrial ROS formation and release was named
ROS-induced ROS release (RIRR). Brief, reversible mPTP opening-associated ROS release appar-
ently constitutes an adaptive housekeeping function by the timely release from mitochondria of
accumulated potentially toxic levels of ROS (and Ca2⫹). At higher ROS levels, longer mPTP openings
may release a ROS burst leading to destruction of mitochondria, and if propagated from mitochon-
drion to mitochondrion, of the cell itself. The destructive function of RIRR may serve a physiological
role by removal of unwanted cells or damaged mitochondria, or cause the pathological elimination
of vital and essential mitochondria and cells. The adaptive release of sufficient ROS into the vicinity
of mitochondria may also activate local pools of redox-sensitive enzymes involved in protective signaling
pathways that limit ischemic damage to mitochondria and cells in that area. Maladaptive mPTP- or
IMAC-related RIRR may also be playing a role in aging. Because the mechanism of mitochondrial RIRR
highlights the central role of mitochondria-formed ROS, we discuss all of the known ROS-producing sites
(shown in vitro) and their relevance to the mitochondrial ROS production in vivo.

I. INTRODUCTION 909 opening is a mitochondrial response to an oxidative chal-


II. ROS: GENERAL DEFINITIONS 910 lenge resulting in an amplified ROS signal, which depending
III. ROS: FROM SIGNALING TO... 910 on ROS levels may result in different outcomes. In addition
IV. ROS: REDOX STRESS 911 to ROS effects in those mitochondria (where the RIRR orig-
V. ROS GENERATION IN MITOCHONDRIA 911 inated), ROS released into cytosol could trigger a complex
VI. ROS-INDUCED ROS RELEASE... 922 cellular signaling response and/or RIRR in the neighboring
VII. RIRR AND Ca2ⴙ 927 mitochondria. In the latter case, ROS trafficking between
VIII. MITOCHONDRIAL COMPARTMENTATION 928 mitochondria could constitute a positive-feedback mecha-
IX. MITOCHONDRIAL RIRR IN... 929 nism resulting in an elevated production of ROS that could
X. IMAC-ASSOCIATED RIRR 930 be propagated throughout the cell and may cause percepti-
XI. RIRR-RELATED PATHOLOGIES: THE... 931 ble mitochondrial and cellular injury. Although photo-in-
XII. PRACTICAL ASPECTS OF USING... 934 duced formation of ROS could be initially used in the ex-
XIII. CONCLUDING COMMENTS 935 perimental setting as a trigger for more massive, avalanche-
like ROS release, this phenomenon is representing a more
I. INTRODUCTION fundamental mechanism, e.g., light-independent spontane-
ous redox transitions associated with the induction of
Photo-activated reactive oxygen species (ROS) may trigger mPTP or other mitochondrial channel(s) that may occur
mitochondrial permeability transition pore (mPTP) induc- under different physiological or pathological conditions
tion within individual mitochondria in intact cell systems. with corresponding impacts on mitochondrial and cellular
The phenomenon of ROS-triggering of the mPTP associ- physiology. This review will cover the spectrum of RIRR-
ated with further stimulation of ROS formation has been related phenomena, both physiological and pathological in-
termed “ROS-induced ROS release” (RIRR) (491). mPTP cluding the processes of mitochondrial ROS production

0031-9333/14 Copyright © 2014 the American Physiological Society 909


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

and scavenging. Ultimately, the imbalance between the in- and metal-oxygen complexes. These radicals combined
flow, neutralization, and outflow of ROS with correspond- with the previously mentioned ROS form a large and im-
ing triggers in specific cell signaling pathways may result in portant group of active redox agents playing critical role in
extreme situations such as oxidative and reductive stresses a number of intra- and extracellular processes.
with the consequent onset of numerous pathologies or even
the cell and organismal death.
III. ROS: FROM SIGNALING TO
PATHOLOGICAL
II. ROS: GENERAL DEFINITIONS
One of our specific aims is to give an overview of a spectrum
Eleven years ago this journal published an excellent and of phenomena associated with different ROS serving a “sig-
comprehensive review by Droge (117) on free radicals and naling” role (discussed below) which is essential for a large
their beneficial and detrimental roles in cell physiology and number of biochemical reactions. In the following sections
pathology. Since then, the general interest surrounding the we address the significance of reductive and oxidative
roles of these species has constantly increased, shifting the stress. Under physiological conditions, the balance between
main focus to highly potent oxidants containing oxygen, ROS generation and ROS scavenging is highly controlled.
called ROS. The term ROS encompasses oxygen free radi- Depending on circumstances, regulated oxidative stress
cals, such as superoxide anion radical (O2·⫺) and hydroxyl could initiate diverse cellular responses ranging from trig-
radical (·OH), and nonradical oxidants, such as hydrogen gering signaling pathways involved in cell protection, initi-
peroxide (H2O2) and singlet oxygen (1O2). ating coordinated activation of mitochondrial fission and
autophagy to optimize clearance of abnormal mitochondria
ROS can be interconverted from one to another (depending and cells to protect spreading the damage to the neighbor-
on ⌬G of relevant processes) by enzymatic and nonenzy- ing mitochondria and cells (100, 117, 490). On the other
matic mechanisms. The primary and most abundant ROS hand, unregulated oxidative and reductive stresses could
is the superoxide anion radical that has a comparatively result in severe cellular damage, unwanted cell death, and
high oxidative capacity [standard redox potential of the consequently whole organ and organism failure (102, 242,
oxygen/superoxide couple ⫽ ⫺0.137 V (337) allowing 498). Therefore, adaptive physiological redox stresses, such
single-electron reduction of molecular oxygen by certain as those occurring under the process of a programmed re-
mitochondrial oxidoreductases]. H2O2 is generated moval of damaged biological systems including mitochon-
through spontaneous or superoxide dismutase (SOD)- dria and other cellular components (“physiological”), must
catalyzed dismutation of O2·⫺ (143). In mammals, three be differentiated from maladaptive unwanted (“pathologi-
SOD isoforms were found in the living cell with precise cal”) oxidative damage.
compartmentalization: the Cu,Zn-dependent isoform
(Cu,Zn SOD, SOD1) (142) is located in the mitochondrial Under normal physiological conditions, ROS emission (es-
intermembrane space and cytosol; the Mn-dependent iso- sentially, production minus scavenging) was considered to
form (Mn SOD, SOD2) (358, 468) is located in the mito- account for ⬃2% of the total oxygen consumed by mito-
chondrial matrix; and Cu,Zn SOD is located in the extra- chondria (81). [A recent measurement of ROS production
cellular space (ecSOD, SOD3) (285). in mitochondria with disabled antioxidant systems revealed
values fluctuating from 0.25 to 11% depending on the an-
The most potent and aggressive oxidant primarily respon- imal species and respiration rates (21).] A lower, as well as
sible for oxidative damage of DNA bases is the hydroxyl a higher, percentage may have deleterious consequences
radical, which has a relatively short half-life. ·OH can be since ROS, when low, are unable to provide proper cellular
generated through a variety of mechanisms. It is well known functioning through regulation of a great number of bio-
that ·OH is generated from H2O2 and O2·⫺ which is cata- chemical reactions. When high, they are unable to provide a
lyzed by iron ions through the Haber-Weiss reaction (169) controlled regulation. It will require conditions when the
with a specific case of Fe2⫹-mediated decomposition of flux of metabolic regulators of ROS level is finely tuned to
H2O2 [the Fenton reaction (130), reviewed in Ref. 237]. respond to the cell demands with mitochondria playing a
Ionizing radiation causes decomposition of H2O, which critical role. As for many cellular signaling elements, the
also results in forming ·OH and hydrogen atoms. ·OH could principle “multet nocem” (excess is harmful) may become a
be also generated by photolytic decomposition of alkylhy- key element in a switch operating between “signaling”
droperoxides (447). (meaning as survival-promoting which is physiologically
required for renovating a biological component) and path-
In addition, a number of other oxygen-containing free rad- ological (meaning undesired death-promoting) modes (216,
icals are capable of causing oxidation of essential cell com- 218, 219, 232, 491, 493, 497). For example, the ROS-
ponents: nitric oxide (NO), peroxynitrite, lipid hydroper- mediated ignition of a death of cells designed for long-term
oxides (LOOH), alkoxyl radical (RO·), peroxyl radical use (postmitotic cells such as cardiac myocytes or neurons),
(·OOH), nitrogen-centered radical, sulfate radical (SO4·⫺) i.e., occurring under severe ischemic conditions causing pa-

910 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

thologies such as myocardial infarct or stroke, under any ation, differentiation, migration, immune response, cell se-
conditions may be considered a pathological event. In other nescence and death, and number of inherited or acquired
cases, apart from their pro-survival signaling role, ROS are pathologies such as ischemia-related disease, atherosclero-
apparently involved in a designated physiological function, sis, neurodegenerative disease, malignant transformation,
eliminating unwanted mitotic cells or mitochondria and a diabetes mellitus, rheumatoid arthritis, aging, etc., is de-
significant rise of local ROS level within might be an effi- scribed elsewhere (25, 117, 185, 203, 263, 348, 352, 497).
cient means to fulfill such functions. Futhermore, ROS have However, under stress, when the ROS levels remain outside
been shown to play a central role in regulation of the cell the normal range (either under conditions of enhanced an-
cycle progression (453). Whether the ROS burst can serve a tioxidative pathways associated with reductive stress or of
signaling (survival) function will be discussed in a section those characterized by the rise of uncompensated ROS as-
describing the oscillatory behavior of mitochondria. For sociated with oxidative stress), resultant instability of the
clarity, we will stay within definitions of signaling ROS as redox environment may develop that could be harmful (un-
of those serving essential pro-survival functions and anti- wanted loss of function) if it is not compensated by the
survival functions including required elimination of un- feedback control mechanism.
wanted cells (quality control, maintenance of function),
while pathological ROS are considered as those causing In summary, considering these scenarios, both high levels of
oxidant-induced unwanted changes including unwanted ROS (oxidative stress) and excessively low levels of ROS
cell death (loss of function). (reductive stress) are deleterious and apparently play a caus-
ative role in the pathologies caused by malfunctioning pro-
cesses related to the dramatic change of redox environment.
IV. ROS: REDOX STRESS Underlying all these arguments, redox homeostasis seems to
be a critical factor for normal functioning of the mitochon-
There is an apparent heterogeneity in ROS levels and types drion, cells, and organisms (174). Previously, it has been
when comparing different cells and organs (54, 173, 271, argued that the cell normally maintains cytosolic thiols in a
342, 373, 401, 477). This is largely due to a heterogeneous highly reduced redox state, thus not supporting the exis-
distribution of activities of ROS producing and utilizing tence of reductive stress (154). More recently, however, a
machineries (11, 78, 103, 290, 346). The general consensus profound increase of reduced glutathione concentration
is that overwhelming ROS production when not compen- and the ratio of GSH/GSSG in cardiomyopathic animals
sated for with their scavenging by endogenous antioxidants carrying the R120G mutation in the ␣B-crystallin molecule
will lead to the rise of ROS beyond the “normal” or “phys- was detected (349). The elevated level of reduced equiva-
iological” threshold level. This results in a process conven- lents in these animals was accompanied by the augmented
tionally called “oxidative stress.” Apparently, the definition expression together with increased antioxidative enzymatic
of this widely used term (up to the year 2012 this term yields activity of glutathione peroxidase, glutathione reductase,
over 100,000 citations in PubMed) is quite broad and has and catalase, which supports the implication of deleterious
“soft borders” considering the scenarios presented above. reductive stress (349). The presence of reductive stress in
This is due to the fact that physiological levels and types of yeast was also confirmed (440). A more precise term, redox
ROS in different tissues and in different parts of the same stress, might be introduced reflecting both the incidence of
tissue under different physiological conditions are hetero- oxidative and reductive stresses; however, this review is
geneous and highly dependent on the energy load that is met focused primarily on the circumstances related to oxidative
by the cell response. Even within the confines of a single cell, stress.
there are at least eight distinct organellar compartments
(mitochondrial matrix, lysosomes, smooth ER/SR, rough
ER, the Golgi, peroxisomes, the nucleus, the cytosol), each V. ROS GENERATION IN MITOCHONDRIA
with its own redox poise (315). Accordingly, the term oxi-
dative stress is often used in the literature in a very general In 1961, Jensen was among the first investigators to dem-
term to define a state when the levels and types of oxidants onstrate that mitochondria produce ROS (209). He ob-
in the cell or the organelle on average significantly exceed served that a small portion of the oxygen consumed by
the ground/resting/steady-state level associated with nor- submitochondrial particles oxidizing NADH or succinate
mal homeostatic function. At the opposite end of the redox was converted to H2O2 since this consumption was catalase
spectrum, when the reduced glutathione levels are too high, sensitive. Later, in 1972–1973, a classic, more general
“reductive stress” occurs and demonstrates potentially det- study, was done at the Johnson Research Foundation in
rimental consequences for the cell (154, 349). Within nor- Philadelphia by Britton Chance and co-workers (57, 58)
mal fluctuation of energy load, the productions of ROS and who initiated the modern era of the mitochondrial ROS
the ROS levels in mitochondria, cells, and the tissue are safe research. Since that time, scientists debated the physiologi-
to perform normal activity (maintenance of function) of the cal relevance of data obtained using “artificial” systems
particular biological system. ROS signaling and the role of such as isolated mitochondria, inside-out submitochondrial
ROS in vital cellular functions associated with cell prolifer- particles, reconstituted respiratory complexes, and pure en-

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 911


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

zymes. According to the critics, these are not adequate sys- detected in mitochondria oxidizing succinate in vitro at ex-
tems to extrapolate data to the cell, organ, and organism perimental ambient (5–10 mM) concentrations, they do not
levels (reviewed in Refs. 340, 480, 492). However, some produce significant amounts of peroxide at low, more phys-
counter-arguments support the relevance of these model iologically relevant succinate concentrations (175).
systems. We are still lacking a detailed mechanistic knowl-
edge of the architecture of mitochondrial ROS-producing Some studies point to flavins (flavin adenine nucleotides,
systems such as of complex I or complex III and detailed FAD) of SDH rather than other electron-carrying compo-
insights on the mechanisms controlling their activities. We nents (such as iron-sulfur clusters or quinones) as the site of
will make an attempt to partially address and clarify this autoxidation responsible for generating ROS (200, 293,
scientific debate and to present the arguments in support of, 294). Others implicate ubisemiquinone and iron sulfur cen-
and against, the importance and physiological relevance of ters as these sites, although under normal and steady-state
those specific proposed mitochondrial ROS-producing conditions these components are only partially reduced and
components. The primary goal of this review is not a com- short-lived (167, 192, 267), thus giving a low probability to
prehensive coverage of this specific issue, but the back- transfer single electrons directly to oxygen.
ground that needs to be addressed here to provide a basis
for understanding those cases when an “innocent” molec- 2. Redox regulation of ROS production and redox
ular site (i.e., normally associated with moderate and phys- buffering
iological ROS production) becomes a “killer,” producing
ROS levels leading to the destruction of the biological sys- During oxidation of succinate in isolated respiring mito-
tem (perhaps through some poorly understood amplifica- chondria, electron flow can bifurcate forming direct (to-
tion mechanism). Good reviews on the current mechanisms wards cytochrome oxidase) and reverse (toward NAD; ro-
of ROS production in mitochondria are available elsewhere tenone-blocked) transport with the latter requiring energy
for the reader interested in a general background and for input (79, 80, 187). The succinate-driven ROS generation
those interested in substantially detailed mechanistic depth during reverse electron transport from succinate to NAD
(3, 13, 201, 202, 308, 397, 416). resulting in the formation of NADH is higher when com-
pared with that forming under direct oxidation of NAD-
dependent substrates (456). The observed relationship be-
A. Complex II tween ROS formation and the redox state of the couple
NADH/NAD resulted in the proposition that the ROS for-
1. Under normal conditions mation is directly proportional to the level of reduction of
NAD. Possibly, a more generalized rule might be formu-
We begin this review of mitochondrial ROS-producing sites lated that the more reduced the mitochondrial interior is,
from complex II since succinate is a more frequently used the more probable there will be primary ROS formation.
oxidative substrate to explore the functioning of isolated
mitochondria which became a classical object to study ROS The redox state of the cellular milieu is mainly determined
production. by the ratios of reduced/oxidized cofactors and proteins
which carry the bulk of redox-sensitive amino acid residues
Complex II, succinate-ubiquinone oxidoreductase (EC and functional groups, NAD(P)H/NAD(P)⫹ and GSH/
1.3.5.1), commonly known as succinate dehydrogenase GSSG, which all together form a compartmentalized redox
(SDH), is a tetrameric iron-sulfur flavoprotein of the inner buffer where all components are in a redox equilibrium
mitochondrial membrane and acts as part of the Krebs cycle under cellular steady-state conditions (reviewed in Ref.
and respiratory chain. SDH catalyzes the conversion of suc- 309). This buffer may be an important factor in determining
cinate into fumarate, yielding reduced equivalents in the ROS levels in the compartments such as the mitochondrial
form of reduced flavin adenine nucleotide (FADH2). This is matrix or cytosol (174). High intramitochondrial redox
followed by a reduction of ubiquinone to ubiquinol. Mam- buffering capacity, only partially represented by 3–5 mM
malian Complex II, as well as that from yeast, harbors a NAD(P)H and 2–14 mM GSH (416), would resist the short-
covalently bound FAD, three iron-sulfur clusters, a b-type term exposure to ROS, while profound sustained ROS ex-
heme, and two quinone-binding sites termed Qp and Qd, posure would eventually exhaust this buffer, resulting in the
standing for proximal or distal sites correspondingly. elevation of intramitochondrial ROS levels (353, 416, 460).
Later, we discuss in greater detail the redox dependence of
Typically, complex II is excluded from the list of potential ROS formation and the role of reducing equivalents and
candidates for important physiological contributors of mitochondrial membrane potential on the net ROS produc-
ROS (347, 348, 358). It is partially due to fact that the tion (see sect. VB5).
succinate level in the tissue is low (in a range of hundreds of
micromoles), while in in vitro experiments with isolated The role of complex II in maintaining and modulating the
mitochondria millimolar concentrations are used. Hans- mitochondrial/cellular redox environment remains unde-
ford et al. (175) found that while H2O2 production can be termined. It is unknown whether in in vivo mitochondria

912 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

reverse electron transfer from complex II to complex I oc- ciation (see FIGURE 1) might result in a direct single-electron
curs, and whether under physiological conditions the re- reduction of oxygen by a reduced iron-sulfur cluster of com-
verse electron transport could result in substantial ROS plex II (113). Consequently, it has been proposed that com-
production considering that physiological concentrations plex II may function as a general sensor for apoptosis (162,
of NADH would significantly attenuate O2·⫺ production 260). This is an example of a pathological, conformation-
under conditions where reverse electron transport could be induced ROS production which does not happen in intact
observed in in vitro model systems (165). Thus it remains complex II.
questionable under normal conditions if there is a signifi-
cant contribution of ROS generated in complex II to the net In a catalytic mechanism, ubiquinone receives electrons
ROS production. from the [3Fe-4S] center being bound at the Qp site. Since
the ubiquinone is a two-electron acceptor receiving these
3. Under pathological conditions electrons in separate steps, the intermediate state of SDH
exists where after receiving a single electron the ubisemiqui-
As we discussed previously, the question about complex II none radical must be stabilized to prevent the escape of the
contributing to the net ROS production remains controver- electron to an inappropriate acceptor such as molecular
sial. Although the tissue level of succinate is as low as 200 – oxygen. Mutations in the vicinity of the Qp site were shown
500 ␮M, under oxygen deficiency (hypoxia/ischemia) it (167) to compromise the ability to stabilize ubisemiqui-
may rise 5- to 10-fold (44, 371, 471). Recently, significantly none, and thus its unpaired electron may become more
increased levels of succinate (to a few millimolar range) readily available to react with ambient oxygen producing its
were detected in macrophages exposed to lipopolysaccha- derivative, superoxide anion radical followed by dismuta-
ride, ultimately identifying succinate as a metabolite in in- tion to form hydrogen peroxide.
nate immune signaling, which stabilizes HIF-1␣ and en-
hances interleukin-1␤ production during inflammation There is multiple evidence that impaired electron transport
(433). Activation of macrophages is known to be associated in SDH, as well as its effect on the levels of NAD(P)H
with elevation of their ROS production, but whether succi- through the impairment of the Krebs cycle, are the source
nate triggers this production remains unexplored. and the cause of a substantial amount of ROS determining
the onset of numerous pathologies (e.g., Refs. 32, 204, 481,
Under some circumstances of drug-induced apoptosis when 482). Malfunctioning of respiratory complexes, including
intracellular pH becomes significantly acidified, impair- complex II in brain mitochondria, is a hallmark of Hunting-
ment of complex II could correlate with ROS generation ton disease (HD), a neurodegenerative genetic disorder that
without changes in the SDH enzymatic activity which is a elicits progressive motor, cognitive, and emotional deficits.
part of complex II activity (260). This process has been 3-Nitropropionic acid, an irreversible inhibitor of SDH,
accompanied by dissociation of the SDHA (flavoprotein mimics HD-like pathology and symptoms (68) and evokes
subunit)/SDHB (iron-sulfur protein-containing part) sub- an ROS increase in neurons (270). Leigh syndrome, an in-
units, which encompass the SDH activity, from the mem- fantile-onset progressive neurodegenerative human disease
brane-bound components of complex II that are required is suggested to be caused by mutations in the SDHA gene. It
for the SQR activity (for details, see FIGURE 1). Such disso- appears that mutations in the SDHB, SDHC, or SDHD

FIGURE 1. Model for the role of specific complex II inhi-


bition for apoptosis induction by various proapoptotic com-
pounds. Mitochondrial matrix, inner membrane (IM), inter-
membrane space, outer membrane (OM), and cytosolic
space are indicated. A: in healthy cells, complex II serves to
funnel electrons derived from the Krebs cycle to the respi-
ratory chain. SDHA-mediated oxidation of succinate to fu-
marate by the succinate dehydrogenase activity (SDH), as
part of the Krebs cycle, provides electrons to complex II.
They are transferred to the iron-sulfur centers of the
SDHB subunit and finally to the CoQ reduction site the
succinate CoQ oxidoreductase (SQR). B: proapoptotic
compounds, such as various anticancer drugs, FasL, or
tumor necrosis factor (TNF)-␣, induce cytosolic (pHc) and
mitochondrial (pHM) acidification. These pH changes lead
to the dissociation of the SDHA/B subunits from complex
II and finally to the partial inhibition of the SQR activity
without any impairment of the SDH reaction. This specific
inhibition leads to complex II uncoupling, superoxide pro-
duction, and apoptosis. [From Lemarie et al. (260). Re-
printed by permission from Macmillan Publishers, Ltd.]

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 913


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

genes can cause paraganglioma (a neuroendocrine, highly There is debate on the critical role of the components of
vascularized neoplasm developing tumors in the head, neck, complex I involved in superoxide production. Some con-
thorax, or abdomen) (37) or pheochromocytoma (a cate- sider FMN (247, 256, 287, 345, 455), while others claim
cholamine-secreting neuroendocrine tumor occurring in the that iron-sulfur clusters N1a and N2 (142, 152, 186, 254),
medulla of the adrenal glands) (37, 155; reviewed in Ref. NAD radical (245), or ubisemiquinone (241) are responsi-
305). Unfortunately, it is impractical so far to estimate the ble for O2·⫺ generation in complex I. The last point was
contribution of the impaired Krebs cycle and reverse elec- actively challenged by Lenaz (264) who considered only
tron transport occurring under pathological SDH impair- hydrophilic quinones to be prooxidants, while physiologi-
ment to the modulating ROS production. cal hydrophobic ubiquinones (such as CoQ10) behave more
as antioxidants rather than prooxidants (264). Therefore,
In parasitic worms residing in an anaerobic environment in the question regarding the major source of superoxide in
a host intestine, the energy partially is obtained from so- complex I under physiological conditions remains unre-
called fumarate respiration reflecting a reverse activity of solved.
succinate-ubiquinone reductase of complex II. Apparently,
in their fumarate reductase reaction, ROS are produced in a Recently, the physiological relevance and thus the impor-
FAD site and quinone-binding site as well. Since in the adult tance of the production of ROS by complex I was ques-
stage, these worms do not have either complex III or IV, tioned on the basis that NADH-supported complex I-cata-
their respiratory chain could serve as a good model to study lyzed superoxide production by submitochondrial particles
the production of ROS in complex II in mitochondria (364, shows maximal activity at low NADH concentrations (⬃50
436). It is noteworthy that this model with a missing cyto- ␮M) while at physiological concentrations of NADH (in the
chrome c oxidase may somehow simulate either hypoxic millimolar range) this reaction is severely inhibited (164,
conditions or those induced by defective electron transfer 165).
downstream of Complex II.
2. Under pathological conditions

B. Complex I It has been noticed that at least 40% of all mitochondrial


disorders are associated with mutations in subunits of com-
1. Under normal conditions plex I (402). Defects in complex I are associated with a wide
diversity of neurodegenerative pathologies, including Par-
The association of complex I deficiency with a wide spec- kinson’s disease (PD) which is characterized by a substan-
trum of pathologies such as cardiomyopathies, cataracts, tial loss of the dopaminergic neurons and cell bodies of
Leigh disease, exercise intolerance, mitochondrial encepha- which are in the substantia nigra pars compacta and nerve
lomyopathy, lactic acidosis, strokelike episodes (MELAS), terminals in the striatum. ROS are thought to be highly
hepatopathy, and tubulopathy has been suggested. Prevail- involved in PD pathogenesis, triggering the loss of redox
ing dogma holds that complex I (NADH-ubiquinone oxi- buffers (GSH and proteinaceous thiols) (336) at least partly
doreductase) is the main source of ROS in mitochondria. caused by dopamine oxidation-related metabolic pathways.
However, the ROS production at complex I depends on
circumstances; consequently, complex I becomes a major Dopamine in the central nervous system, apart from being a
ROS source under pathological conditions rather than be- neuronal neurotransmitter, serves as a precursor of norepi-
ing a dominant source under resting and healthy conditions. nephrine and epinephrine, and is a regulator of movement
(nigrostriatal pathway), and a behavior motivator (me-
When submitochondrial particles or isolated mitochondria solimbic pathway) (425). While under normal conditions
oxidize NAD(P)H or glutamate plus malate, correspond- oxidative deamination of dopamine by monoamine oxidase
ingly, complex I production of superoxide is negligible. produces hydrogen peroxide (282), it could generate toxic
However, supplementation with the inhibitor of complex I, oxidants through alternative ways of oxidation wherein
rotenone, results in robust production of O2·⫺ (350, 456). mitochondria play a role. In this pathway, dopamine is
This implies that the major site of ROS production in com- oxidized nonenzymatically by superoxide forming dopa-
plex I is either upstream of a rotenone-binding site or it is mine quinone which can be reduced by mitochondrial com-
tightly coupled to the increased level of NAD(P)H after plex I generating semiquinone followed by a transfer of its
rotenone supplementation during oxidation of NAD-de- electron to molecular oxygen to form superoxide (488),
pendent substrates (175, 444). According to the first alter- completing a vicious oxidative cycle. In addition, PD is
native, rotenone would induce progressive reduction of the hallmarked by elevated iron levels that may catalyze pro-
upstream redox groups (432) including Fe-S clusters, flavin duction of deadly oxidants, possibly in a self-amplifying
mononucleotide (FMN), and the tightly bound pool of mode (411).
ubiquinone (62, 325), which can supply the oxygen mole-
cule with a single electron yielding superoxide anion radi- PD could be mimicked by the action of complex I inhibitors
cal. such as rotenone, paraquat, and 1-methyl-4-phenyl-1,2,

914 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

3,6-tetrahydropyridine (49, 94). Exposure to the latter drug of oxygen is limiting its utilization. Note that this paradox
has shown to produce permanent parkinsonism in humans, is absent in the isolated mitochondrial system (FIGURE 2),
non-human primates, and rodents, by exerting an effect free from extramitochondrial signaling pathways which
primarily on the function of mitochondrial complex I. In confirms that elevated mitochondrial ROS generation in the
patients with Friedreich ataxia, a deficient activity of the cell in response to hypoxia is not intrinsic to the mitochon-
Fe-S cluster-containing subunits of mitochondrial respira- drial respiratory chain alone but can be attributed to some
tory complexes I, II, III, and aconitase was found (361). involvement of extramitochondrial factors (189). Marshal
Kushnareva et al. (254) claimed that the ratio of NAD(P)H/ et al. (286) indicated that hypoxia-induced superoxide pro-
NAD(P)⫹, rather than the level of NADH, determines re- duction occurs through activation of NADPH oxidase lo-
duction of ROS-producing sites in complex I. cated in the cell membrane. In addition, under moderate
hypoxia, NO synthesis in mitochondria continues although
Generation of ROS associated with hypoxia/reoxygenation being only 5–10% of the normal steady-state level (Km for
is known as one of the most deleterious causes of oxidative oxygen of the mitochondrial NO synthase is 30 – 40 ␮M;
damage. Three potential sources of ROS have been pro- Ref. 8). In turn, NO can partially block cytochrome oxidase
posed to be responsible for this release: mitochondrial com- (69, 70, 378), thus reducing mitochondrial electron carri-
plex I, xanthine oxidase, and NADPH oxidase (2, 471, ers, increasing its Km for oxygen (91) and favoring genera-
491). However, the latter two are probably not involved tion of superoxide at hypoxic conditions (444).
since inhibition of these complexes in vivo did not afford
cell protection (84, 129). In highly metabolizing tissues, the areas surrounding mito-
chondria in the cell may have higher ROS levels than remote
One of the very specific features of the mammalian NADH- areas. Without considerable mitochondrial ROS-quenching
ubiquinone oxidoreductase is the slow active/deactive state activities, intramitochondrial levels of ROS may potentially
transition, suggesting gross conformational rearrangements reach very high levels. That may happen in case of an im-
of complex I, at least in that part which is involved in balance between the oxygen supply and demand, for exam-
rotenone-sensitive ubiquinone reduction [which may be in- ple, under conditions of high metabolic needs.
volved in the superoxide production (150, 454)]. It was
found that complex I isolated from the heart which was The nature of the tissue oxygen gradient between the source
exposed to a normoxic perfusion is in a fully active state, of oxygen (blood capillary) and the site of its utilization (the
while 30-min anoxic perfusion results in a significant trans- mitochondrion) can be explained by the Krogh cylinder
formation of the enzyme into a deactive state which returns model which generally serves to analyze capillary tissue
back to normal after reoxygenation (283). It has been pro- exchange kinetics (246, 266) (FIGURE 3). The effective ra-
posed that these conformational transitions can be relevant dius of this cylinder beyond which the tissue hypothetically
to producing ROS by complex I after cardiac tissue is re- is experiencing hypoxia depends in part on O2 consump-
oxygenated following a coronary occlusion (283). Using tion. In a tissue with high metabolic rate (such as heart
EPR spectroscopy, DeJong et al. (104) showed that NADH- muscle), capillary density during maximum or moderate
coenzyme Q oxidoreductase undergoes energy-dependent exercise would not be sufficient to supply tissue with oxy-
structural changes in parts determining ubisemiquinone gen and might potentially produce more ROS in the vicinity
production (iron-sulfur cluster 2) (104). Thus, under path-
ological conditions, conformational rearrangements may 120 80
be involved in the changes of the efficiency of ROS-produc-
Respiration rate, nmol o2/min/mg

ing machinery in complex I. 100 H2O2 emission, pmol/min/mg


60
3. ROS and hypoxia 80

The reaction of formation of a primary ROS (superoxide 60 40


only) generated in the respiratory chain from molecular
oxygen is of a first order with respect to oxygen concentra- 40
tion. However, paradoxically, generating ROS in mito- 20
chondria in the cell remains constant or even increases when 20
PO2 drops dramatically (i.e., under moderate hypoxic con- State 3
ditions). Robust ROS production under 1.5% of O2 has 0 0
been recorded also (314, 374, 462). 0 1 3 5 7 9 11
[O2], µM
Interestingly, in the cell the affinity of molecular oxygen to
FIGURE 2. Effect of the oxygen tension on the rates of phosphor-
ROS-generating modules is higher than to cytochrome ox- ylating respiration and H2O2 emission in rat liver mitochondria. (Note
idase. This obviously takes place under conditions of partial that there is no ROS production increase under low PO2.) [From
reduction of cytochrome oxidase, i.e., when the availability Starkov (416), with permission from John Wiley & Sons.]

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 915


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

tion by the mitochondrial respiratory chain. On the other


hand, the same reduced redox conditions provide more
Rt buffering capacity to quench ROS activity, so it is unclear
Z under which conditions (more reduced or more oxidized)
Rc
Flow Direction the net level of ROS will be higher. It appears that the
steady-state level of ROS in the compartment rather than
the ROS-producing activity determines the level of oxidant-
induced biological modifications, many of which are impor-
tant because of their biological effects, both physiological
FIGURE 3. Geometry of the Krogh cylinder-type model. Inner cyl- and pathological.
inder represents the capillary. Outer cylinder corresponds to tissue
cylinder. Shaded area: example of hypoxic region under conditions of Redox steady states of respiratory components responsible
high demand. Rt, tissue cylinder radius; Rc, capillary radius; z, dis-
tance along the capillary. [From McGuire and Secomb (289).]
for ROS production should be in redox equilibrium with
adjacent mobile and immobile redox carriers such as
and inside of mitochondria/mitochondrial clusters (214). NAD(P)H, GSH, and the thiol groups of the proteins occu-
However, basic theoretical assumptions of the Krogh cylin- pying the same compartment.
der model do not consider that the diffusion coefficient for
O2 in muscle tissue may be higher due to the possibility of In 1996 Liu and Huang (275) and in 1997 Korshunov et al.
facilitated O2 transport by the mitochondrial network (10, (240) (FIGURE 5A) found in isolated mitochondria a strong
396) and/or by myoglobin molecules (475). In non-muscle dependence of mitochondrial ROS production on the level
tissues lacking myoglobin, the cytoglobins and neuroglo- of the transmembrane potential (⌬␺), supported by succi-
bins may potentially serve as facilitators of oxygen trans- nate oxidation (reviewed in Refs. 272–274). [It should be
port (220, 327, 354, 428). Both the extended mitochondrial emphasized that mitochondrial matrix alkalinization could
network and these various heme-containing proteins may be the cause of increased generation of ROS due to the fact
be responsible for a less steep O2 gradient in the normal that ⌬pH is an integral part of the ⌬p (381).] While H2O2
active cell, thus also likely flattening the ROS gradient as production was low at ⌬␺ values at and below the phos-
well. Cytoglobin and neuroglobin have been reported to not phorylating membrane potential [the membrane potential
only exert an oxygen carrier function, but possibly to be reached under the state 3 respiration according to B.
oxygen sensors and ROS scavengers (220). Chance’s terminology when mitochondria are supple-
mented with an excess of ADP (82), i.e., that thermody-
Under experimental conditions acutely restricting the oxy- namically required to generate ATP from ADP and Pi], it
gen supply to an organ, such as the heart, very steep re-
gional redox transitions have been observed across the bor-
derline of the ischemic area (33), apparently reflecting a
similarly steep oxygen gradient and thus probably the ROS
gradient as well (FIGURE 4). On the basis of the data that
ROS production is directly linked to reduced equivalents
such as NADH (discussed above), hypoxic regions mani-
festing the highest level of NAD reduction would likely
achieve much higher ROS levels than normoxic regions.
The brain is the organ most vulnerable to the lack of oxy-
gen, immediately responding by the reduction of NAD (78),
thus potentially eliciting ROS within local areas adjacent to
ischemic zones.

4. Is the ROS production proton motive force


sensitive?

Proton motive force (pmf, ⌬p) across the inner mitochon-


drial membrane, with ⌬␺ as a main component, is driving
ATP production (268, 298, 299, 424). Whether ROS pro-
duction is dependent on this proton motive force/trans-
membrane potential in mitochondria has become a crucial
question.
FIGURE 4. NADH fluorescence emission from perfused rat heart
with a local ischemic area near the apex (seen as white areas),
As discussed above, a reduced redox intramitochondrial caused by ligation of a coronary artery. [From Korshunov et al. (33).
environment is a prerequisite for high primary ROS forma- Reprinted with permission from AAAS.]

916 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

A B
100 110

H2O2 production, pmol x min-1 x mg-1


glutamate + malate
90 100 α-ketoglutarate
80 + ADP
90

H2O2 production, %
70
80

State 3
60
70
50
60
40
50
30
40
20

10 30 State 3

0 20
50 60 70 80 90 100 100 120 140 160 180
ΔΨ, % ΔΨ, %
FIGURE 5. Relationship between mitochondrial ⌬⌿ and H2O2 production supported by succinate or NADH-
linked respiratory substrates. A: rat heart mitochondria oxidizing succinate. The ⌬␺ level was varied by adding
different concentrations of uncoupler SF6847 (black squares and solid line), malonate (white squares), or 100
␮M ADP and 5 mM P (triangle). Dashed line, the state 3 ⌬␺ level. [From Korshunov et al. (240), with
permission from Elsevier.] B: rat brain mitochondria oxidizing glutamate ⫹ malate or ␣-ketoglutarate. Differ-
ences in ⌬⌿ were generated by adding various concentrations of uncoupler FCCP ranging from 0 to 80 nM.
Alternatively, a decrease in ⌬␺ was induced by adding 0.8 mM ADP to mitochondria. [From Starkov and Fiskum
(419), with permission from John Wiley & Sons.]

rises dramatically above these values proportionally to the ological state. As discussed previously, regulated moderate
⌬␺ elevation (FIGURE 5). Accordingly, an 18% decrease in (“mild”) uncoupling of mitochondrial oxidative phosphor-
the value of the transmembrane potential inhibits 90% of ylation has been suggested as a feasible therapeutic strategy
mitochondrial ROS production (240) above the phosphor- (397, 398, 415) for regulation of the intracellular and in-
ylating membrane potential. tramitochondrial ROS level (99).

Subsequently, very steep dependence of H2O2 production Mitochondrial uncoupling proteins (UCPs) have been con-
on the values of ⌬␺ exceeding the phosphorylation poten- sidered as potential mild uncouplers. The relationship be-
tial was confirmed in isolated mitochondria oxidizing tween ROS production and UCPs activity was revealed in
NADH-dependent substrates (419) (FIGURE 5B). However, 1997 in experiments where GDP, an inhibitor of UCP1,
in another study using NADH-dependent substrates, mito- caused an increase of ⌬␺ and ROS production (316). Later,
chondrial respiration produced ROS in a membrane poten- it was demonstrated that superoxide directly activates
tial-independent mode (456). UCPs resulting in a negative feedback controlling both ROS
production and their levels (120).
5. Can ROS production be decreased in mitochondria
without jeopardizing ATP production? Mild uncoupling Mild uncoupling may be protective against excitotoxic in-
as a possible downregulator of ROS production jury (469) and against injury of dopaminergic neurons in
substantia nigra from mitochondrial poisons such as rote-
As was shown in the preceding section (FIGURE 5), moder- none (478). Decreasing ROS generation by uncoupling mi-
ate lowering of ⌬␺ could result in a lower ROS production tochondria increases longevity in healthy animals (74).
in mitochondria without a significant effect on ATP produc-
tion. This could be achieved by inducing a small proton leak Of all the possible mild uncouplers, fatty acids are probably
through the inner mitochondrial membrane which would the most natural ones (240, 398, 476). In their protonated
both stimulate oxygen consumption and, in parallel, shift form they can cross the mitochondrial inner membrane fol-
the level of reduction of mitochondrial ROS-producing sites lowed by deprotonation in the matrix side, and then the
to a more oxidized state lowering the probability of ROS anionic form of the fatty acid completes the cycle by return-
production in the mitochondria. While higher potential ing back to the cytosolic side. The rate-limiting step of this
could drive increased ATP production, the higher ⌬␺ could cycle is the transport of anionic form. Different proteins
also result in the production of increased ROS levels and such as the adenine nucleotide transporter (ANT) and glu-
potentially unwanted oxidative consequences. Therefore, tamate/aspartate transporter are involved in fatty acid-me-
achieving a reasonable balance between ROS and ATP pro- diated uncoupling through facilitation of the transport of
duction by mitochondria is crucial since this reflects the the anionic form (12, 368). Thyroid hormones may also be
current energy needs of the cell under the particular physi- considered as natural mild uncouplers (177, 397).

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 917


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

Among artificial uncouplers, 2,4-dinitrophenol (DNP) has


been tested as an anti-obesity drug, but it was found to be
too toxic for practical use (90, 99). Additionally, DNP was
found to limit the experimental infarct size in the heart and
brain, and this was interpreted to occur through diminish-
ing the ROS level (238, 359). Recently, mild uncoupling
activity was ascribed to a series of derivatives of cationic
rhodamine (15).

C. Complex III

Complex III (ubiquinol-cytochrome c oxidoreductase) ac-


cepts reducing equivalents formed in complexes I and II and
processes them by the Q-cycle operating mechanism. Oper-
ation of this cycle is initialized by ubiquinol, which releases
its proton to the intermembrane space and donates one
electron to the Riske iron-sulfur protein (which can bind to,
and be inhibited by, myxothiazol) producing unstable
semiquinone on the outer side of the inner mitochondrial
membrane. The semiquinone serves as an electron donor
for hemes of cytochrome bL, and then of cytochrome bH
which is located close to the inner side of the membrane.
Cytochrome bH reduces ubiquinone in an antimycin A-sen- FIGURE 6. Q-cycle model. The mechanism of superoxide forma-
sitive way producing ubisemiquinone followed by its fur- tion in complex III (bcl1 complex). The reaction starts from the
ther reduction with a second electron and protonation oxidation of the CoQ quinol (QH2) in a bifurcated electron transfer
(442) (FIGURE 6). reaction at the Qo site of the complex III. The first electron is trans-
ferred to a high reduction potential chain consisting of the iron-sulfur
protein (ISP, aka Rieske protein), cytochrome c1 (cyt c1), cyto-
Under normal conditions, the probability of existence of chrome c (cyt c), and further to cytochrome c oxidase (not shown).
unstable semiquinone (Q_.) is low due to its fast oxidation; The remaining semiquinone (Qo⫺·) is unstable. It donates the sec-
therefore, the probability of donation of one electron to ond electron to the low reduction potential chain consisting of two
molecular oxygen in this system is relatively low. Only the cytochromes b, cyt bL and cyt bH, which serve as a pathway con-
block of the electron flow by antimycin A results in a high ducting electrons to the Qi site. There these electrons reduce an-
other CoQ molecule. To provide two electrons required for the com-
superoxide release apparently due to the reduction of both plete reduction of CoQ quinone at the Qi-site, the Qo-site oxidizes two
hemes of the cytochrome c in parallel with the elevation of QH2 molecules in two successive steps. The first electron at the
the steady-state level of semiquinone, thus giving a higher Qi-site generates a stable semiquinone (Qi⫺) that is reduced to a
chance of one-electron reduction of oxygen (FIGURE 6). quinol (QH2) by the second electron. Most frequently used inhibitors
Among all of the mechanisms presented in the scheme, in- of complex III, stigmatellin and myxothiazol, prevent the transfer of
the first electron to ISP and the binding of the quinol at the Qo site
hibitors of bc1 complex (antimycin A, myxothiazol, and correspondingly. [Modified from Starkov and Fiskum (418), with
stigmatellin), antimycin is the only effective ROS inducer, permission from Elsevier.]
although some low level of superoxide production could be
detected in the presence of other inhibitors (418) in the site
of bc1 complex different from that induced by antimycin A. ratory chain in close proximity to an antimycin A-bind-
Therefore, the potential role of complex III as a cause of ing site. B. Chance’s lab performed a study on a patient
gross mitochondrial ROS production under the physiolog- who was diagnosed with a deficiency of cytochrome b in
ical steady-state mode of operation remains uncertain con- complex III, resulting in muscle weakness associated with
sidering that substantial mitochondrial ROS release occurs a ragged-red fiber myopathy and lactic acidosis (122).
only after application of a drug having no natural analogs in The total succinate-cytochrome c reductase activity in
animal physiology. Noteworthy, as in the case of ROS pro- skeletal muscle of this patient was only ⬃5% of normal.
duction in complex I, conformational changes detected in It was already known that menadione can shuttle at least
bcl1 complex after antimycin A binding (45, 95, 193, 355) a portion of electrons over an antimycin-sensitive site
may be a prerequisite for dramatic molecular rearrange- (319). The treatment of this patient with menadione
ments in the complex resulting in a marked amount of ROS bridging electrons from coenzyme Q directly to cyto-
production. chrome c, thus bypassing the defective cytochrome b,
resulted in a significant therapeutic effect with partially
Overproduction of ROS by complex III may result from normalized muscle exercise tests. This approach, named
acquired and genetic defects in the mitochondrial respi- “redox therapy,” demonstrates the importance of de-

918 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

tailed understanding of mitochondrial redox-related drial autofluorescence signal originating from cellular fla-
pathogenesis. vins (178, 250)] with the redox state in equilibrium with the
environmental NAD(P)H/NAD(P)⫹. It has been found that
E3 is responsible for superoxide and hydrogen peroxide
D. Importance of Redox State of NAD(P)H/ generation in purified KGDHC and PDHC as well as in
NAD(P)ⴙ in Mitochondrial ROS KGDHC operating in mitochondria in vitro (71, 151, 417,
Production 420, 438).

A comprehensive analysis of the relevance of NADH in It is not surprising that ammonium has an effect on the
managing mitochondrial production of ROS has been per- component of ␣-KGDHC since ␣-ketoglutarate, instead of
formed by Vinogradov’s group (163, 165). With the use of converting into succinyl CoA in the citric acid cycle (the
submitochondrial particles oxidizing NADH, it has been Krebs cycle), can be transformed into glutamate by gluta-
found that a substantial amount of superoxide production mate dehydrogenase. Typically, this reaction does not take
took place only when 50 ␮M NADH was used, while in the place in mammals, since the equilibrium of the reaction is
presence of 1 mM NADH, the production was remarkably shifted toward the reverse direction, but it may occur in
suppressed. NAD⫹ revealed the same superoxide suppres- toxic levels of ammonia. Ammonia metabolism is impor-
sive ability (165). Considering that physiological concen- tant in all tissues. However, in the brain for which a high
tration of the couple NAD⫹ ⫹ NADH in the mitochondrial level of ammonium is extremely toxic (59), it becomes a
matrix is in range of a few millimolar (474), with a signifi- critical element involved not only in the detoxification pro-
cant fraction of it existing in a free form, the gross genera- cess (by astrocytic glutamine synthase and the all-mito-
tion of ROS mediated by complex I may be almost negligi- chondria-located urea cycle) but also in a number of essen-
ble. Similar experiments with isolated permeabilized mito- tial biochemical reactions in the cell as part of the brain
chondria and the soluble protein fraction of the signaling modules (i.e., glutaminase reaction to maintain
mitochondrial matrix showed the same result. The authors optimal cycling of glutamine/glutamate; Refs. 421, 422,
concluded that complex I was not a primary source of ROS 448). With the consideration of results mentioned above,
in mitochondria under physiological conditions. Instead, the ammonium toxicity might be at least partially mediated
they hypothesized that some oxidoreductases poised in by the mitochondria-formed ROS.
equilibrium with NAD(P)/NAD(P)H may be that primary
mitochondrial source of ROS (165).
E. Other Mitochondrial ROS-Producing Sites
In isolated permeabilized mitochondria, the same authors
detected quite high NADH-dependent H2O2 production 1. NADPH-oxidase
when they supplemented the system with ammonium salts
(163). The mitochondrial H2O2 release was insensitive to The prototypic NADPH-oxidase (Nox) has been found in
dicumarol (inhibitor of NADH-quinone oxidoreductase, the plasma membrane of phagocytes and B lymphocytes,
D,T diaphorase) and NAD-OH (inhibitor of complex I), and it is involved in the phagocytic activity by ejecting su-
suggesting the matrix localization of H2O2-producing ac- peroxide radical which is a primary element igniting anti-
tivity. This ROS-generating activity depended on the ratio bacterial defense. Its membrane domain is represented by a
of NAD(P)⫹/NAD(P)H. It was concluded that a specific protein gp91PHOX (PHOX for phagocyte oxidase) which
ammonium-sensitive NADH oxidase activity in the mito- can be organized as a heterodimer in combination with
chondrial matrix is responsible for this H2O2 production, other cytosolic proteins from the PHOS family which all
but the in vivo relevance of this process is still unknown. together form flavocytochrome b558 complex (6, 24).
[An alternative explanation for an ammonium effect could
be the hypothesis that mitochondrial matrix alkalinization So far, seven isoforms of Nox (Nox1–5, Duox1–2) have
(caused by ammonium entry) increases superoxide produc- been identified, with all having distinct catalytic domains.
tion by stabilizing semiquinone radical (381).] The above-mentioned form is called Nox2. It has been
found not only in the cell membrane, but also in the cell
Further analysis of the nature of this ammonium-stimulated interior. Other isoforms reside in different specialized tis-
enzyme producing primary mitochondrial ROS revealed sues and different intracellular loci. Nox4 is the only mem-
that the questioned enzyme possessed NADH:lipoamide ber of the family found to have a mitochondrial localization
oxidoreductase activity and later was identified as dihydro- [in cultured mesangial cells (55), rat kidney cortex (55) and
lipoyl dehydrogenase (224). Dihydrolipoyl dehydrogenase cardiac myocytes (251)]. Nox4 differs from the other mem-
is an essential component (called E3 component) of two bers of the Nox family in that it preferentially produces
mitochondrial redox complexes: ␣-ketoglutarate dehydro- H2O2 rather than O2·⫺ (356). The discovery of its mito-
genase complex (KGDHC) and pyruvate dehydrogenase chondrial localization conflicts with data that shows that
complex (PDHC). This mitochondrial enzyme contains the natural cytosolic partner of Nox4 complex,
FAD [which contributes mostly to the overall mitochon- p22PHOX, whose presence affords NADPH oxidation

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 919


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

was not found in mitochondria, and furthermore, that abolished by MAO inhibitors (179). Compared with nor-
the specific activity of Nox in mitochondria is not mea- mal conditions, MAO produces much more H2O2 during
surable (111). Considering the high importance of NA- ischemia/reperfusion of the brain (389), the kidney (249),
DPH-oxidase, specifically Nox4, in angiogenesis (483) and and the heart (50, 221). MAO activity in cardiac mitochon-
pathogenesis of atherosclerosis, diabetic injury (279) and dria of 24-mo-old rats was about eight times higher than
other pathologies (280) including aging (4), further experi- that in 1-mo-old rats, demonstrating that MAO may be an
mental research is needed to resolve this apparent conflict important source of ROS in the aging heart (109, 288).
and gain a better understanding of the relevant aspects of However, the basic MAO activity in normal tissue is quite
Nox-mediated redox signaling. low due to the limitation of the availability of endogenous
substrates of oxidation (such as serotonin, epinephrine,
An interesting mode of interaction between Nox and mito- norepinephrine, dopamine, and others present in the brain
chondria was shown recently in cancer cells where glucose in only nanomolar concentrations). On the other hand,
deprivation provoked a signaling-based positive-feedback chemical inhibitors of MAO elevate ROS production in
loop that amplifies ROS levels above a toxicity threshold cells (50 –52). Paradoxically, the ablation of MAO causes a
resulting in cell death (159). This positive-feedback loop very slight rise of its endogenous substrates in the brain
involved the complex integration of homeostatic control tissue (386). Because the levels remain in the nanomolar
mechanisms for metabolism (particularly, redox balance range, it is doubtful that these substrates could contribute
established by Nox and mitochondria) and tyrosine kinase significantly to overall ROS production in tissue. However,
signaling through regulation of protein tyrosine phospha- close proximity to the sites where biologically active amines
tases. According to the authors, glucose withdrawal acti- are formed and released (such as synapses and extraneuro-
vates supraphysiological phosphotyrosine signaling and nal compartments such as astrocytes and glial cells) and the
ROS-mediated cell death. In cancer cells that are highly high mitochondrial density in and around these areas, may
dependent on glucose for survival, glucose and pyruvate render mitochondrial MAO an important component in
deprivation induces oxidative stress driven by Nox and mi- both inactivation of amines and the local rather than overall
tochondria. This oxidative stress provokes a positive-feed- ROS production in such areas. It has been speculated that
back loop in which Nox and mitochondria generate ROS under normal physiological activity, ROS produced by
and inhibit tyrosine phosphatases by oxidation. With the MAO in these areas performs metabolic and signaling func-
negative regulators turned off, tyrosine kinase activates tions in the brain (31). In addition to ROS, the by-products
Nox, further amplifying ROS generation and provoking of biologically active amine conversion by MAO may play a
cell death. direct role in degenerative processes, e.g., the dopamine
molecule after entering MAO reaction produces a reactive
2. Monoaminoxidase quinone that could modify and damage cellular compo-
nents (425).
Monoaminoxidase (MAO) resides in the outer mitochon-
drial membrane and serves as a marker there. This flavoen- 3. p66shc
zyme has two isoforms A and B with different substrate
specificity and sensitivity to inhibitors (121). Their sub- On the basis of the observation that p66shc-deficient mice
strates are biogenic amines whose oxidation yields in the display extended life span, remarkably reduced levels of
generation of corresponding aldehydes, H2O2, and ammo- ROS, and increased tolerance to oxidative stress, it has been
nium base. MAO activity has special importance in the suggested that p66shc could have an important role in ROS
brain where peroxidase and catalase activities are low to production and aging (295, 328, 439). Normally p66shc
fully decompose H2O2 formed during oxidative deamina- resides in the cytosol, while under oxidative stress (e.g.,
tion of neurotransmitters (e.g., dopamine, serotonin), thus under ischemia/reperfusion insult), it could be translocated
significantly depleting the endogenous pool of reduced glu- in the mitochondria in a PKC␤-dependent way (341) where
tathione (370). it serves as an important source of ROS (109). In mitochon-
dria, this adapter molecule has been suggested to function
The H2O2-generating activity of MAO might be the highest as a redox enzyme possibly oxidizing cytochrome c and
among all mitochondrial ROS generators. Isolated rat brain generating H2O2 in the amino-terminal portion of p66shc
mitochondria produce H2O2 during oxidation of the exog- containing sequence similar to that of certain redox en-
enous amine, tyramine (at supraphysiological 2 mM con- zymes (156).
centration) at a rate 45.2 ␮M/s (179), while H2O2 produc-
tion during succinate oxidation in the presence of antimycin It was hypothesized that p66shc in mitochondria exists
(considered to be the “gold standard” method for mito- within a high-molecular-weight complex which includes
chondrial ROS production) is 0.95 ␮M/s (179, 334), i.e., mtHSP70 and TIM-TOM complex. Inside of such a com-
MAO activity is 48 times more H2O2-generating than com- plex, p66shc is inactive. After propagation of apoptotic sig-
plex III. Oxidation of tyramine by brain mitochondria re- nal, the complex is dissociated resulting in the release of free
sults in oxidative damage of mitochondrial DNA which is p66shc, which becomes activated and capable of participat-

920 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

ing in the electron transport which generates H2O2. Muta- comes from ETF which accepts electrons from different
tions in the redox active sequence of p66shc abolish its dehydrogenases including those involved in ␤-oxidation
pro-apoptotic activity apparently through the inability to (41) and transfers them to the ubiquinone pool in the inner
interact with cytochrome c (156). Recent data show that mitochondrial membrane by a reaction catalyzed by ETF-
p66shc-generated ROS regulate insulin signaling (48), T-cell ubiquinone oxidoreductase (ETF-QO) residing in the ma-
and B-cell signaling pathways (133), and expression and trix side of the inner mitochondrial membrane (153). In
activity of the ROS-generating enzyme NADPH oxidase ␤-oxidation, the sequence of reactions is as follows: acyl
(NOX4) as well as activate NF-␬B (226, 291), thereby am- CoA dehydrogenase ¡ ETF ¡ ETF-QO ¡ Ubiquinone ¡
plifying oxidative stress and inflammation. complex III.

4. ␣-Glycerophosphate dehydrogenase ETF-QO contains flavin (FAD) and a [4Fe-4S] cluster which
makes it vulnerable to ROS (363), and it serves as a conver-
gence point for electrons flowing from nine flavoprotein
Another potential source of ROS in mitochondria could be
acyl-CoA dehydrogenases and two N-methyl dehydroge-
␣-glycerophosphate dehydrogenase that occupies the outer
nases (153, 227). Recently, it was found that oxidation in
surface of the inner mitochondrial membrane, but its activ-
muscle mitochondria of long-chain fatty acids in physiolog-
ity is relatively low in the liver, heart, and brain but high in
ical (low) concentration is associated with higher rates of
brown adipose tissue (304). However, isolated mitochon-
ROS formation than oxidation of NADH-linked substrates
dria supplemented with ␣-glycerophosphate in the presence
while exhibiting relatively low dependence on the mito-
of antimycin A produce hydrogen peroxide which, when
chondrial membrane potential (380). The authors suggest
normalized to an enzymatic activity, exceeds that originat-
that this enzymatic activity may be responsible for ⌬␺-in-
ing from complexes I or II (116). This source could repre-
dependent ROS production.
sent one of the most efficient ROS generators in mitochon-
dria. It is almost insensitive to the presence of an uncoupler
Deficiency of ETF-QO in most cases is caused by single
or rotenone, implying that ROS generation by ␣-glycero-
point mutations around the FAD-ubiquinone interface
phosphate oxidation is not dependent on mitochondrial ⌬␺
(326) and results in a human genetic disorder known as
contrary to, e.g., succinate oxidation (304). The detailed
multiple acyl-CoA dehydrogenase deficiency (MADD) or
mechanism of ROS production by this enzyme is not de-
glutaric acidemia type II (141). This is characterized by
fined yet.
impaired fat and protein metabolism. It may be associated
with acidosis or hypoglycemia and accompanied by other
In addition to previously mentioned sources of mitochon-
symptoms such as general weakness, liver enlargement, in-
drial ROS production, it was shown that cytochrome b5
creased risk of heart failure, and carnitine deficiency, and
reductase (470) and dihydroorotate dehydrogenase (112,
could result in a fatal metabolic crisis (144, 390, 443).
140, 276) produce ROS on the outer surface of the inner
membrane. However, the significance of these enzymes in
6. Aconitase
the total ROS production remains questionable.
Aconitase catalyzes transformation of citrate to isocitrate in
5. Electron transfer flavoprotein (ETF) and ETF the Krebs cycle. It contains a cubane-type [4Fe-4S] center
quinone oxidoreductase (ETF dehydrogenase) with three iron atoms interacting with cysteine residues and
inorganic sulfur atoms, while the fourth iron, Fe-␣, is ex-
In 1972 Boveris et al. (58) found that rat liver mitochondria posed to the solvent that allows the catalytic dehydration of
can produce a substantial amount of hydrogen peroxide citrate to form the intermediate cis-aconitate, as well as the
when oxidizing palmitoyl carnitine or octanoate. This pro- subsequent hydration of cis-aconitate to form isocitrate
duction was ceased in the presence of an uncoupler. The (42, 138). The prosthetic group of aconitase is highly sus-
fatty acid-induced ROS generation was comparable to that ceptible to inactivation by superoxide anion radical yielding
in the presence of glutamate plus malate and was slightly an inactive [3Fe-4S] form, Fe2⫹-␣ and H2O2 (138, 267,
lower than with succinate (58). Thirty years later, similar 452). Interaction of the latter two ignites Fenton’s reaction
results were obtained with skeletal muscle and heart mito- resulting in the release of ·OH radical (452). It has been
chondria (414). Increased lipid metabolism was correlated proposed that aconitase would be an ideal sensor for ROS
with upregulated UCPs expression/activities (73, 369), pos- in cells (147, 148). Subsequently, superoxide toxicity in
sibly to salvage the system from excessive ROS production mitochondria could be explained by enhanced aconitase
(414). Since ROS generation was almost insensitive to ex- inactivation and related processes. For example, doxorubi-
ternal SOD, it has been proposed that the fatty acid ␤-oxi- cin cardiotoxicity was explained mainly by an aconitase
dation may result in ROS generation at a distinct mitochon- inactivation accompanied by hydroxyl radical release
drial matrix site which is different from o-center of complex (296). Aconitase inactivation is an example of a “ROS
III. In addition, it has been proposed that a significant con- cross-talk” where one type of ROS (superoxide) is inducing
tribution to ROS production during fatty acid oxidation the release of another, potentially more damaging, ROS

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 921


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

(the hydroxyl radical). An alternative point of view is that Although mPTP induction is typically referred to as a path-
aconitase inactivation may serve a protective role by dimin- ological event very often resulting in the degradation of
ishing electron flow through the ROS-generating respira- mitochondria or the cell (which will be discussed later),
tory chain. In addition, the accumulation of citrate as a there is multiple evidence and assumptions that in fact it can
result of a decreased Krebs cycle flux promotes chelation of also serve physiological functions. It has been postulated
Fe2⫹ which is then irreversibly oxidized to a more stable that mPTP could serve as a release valve for quick release of
complex citrate-Fe3⫹, thus preventing catalysis of Fenton’s cations constantly leaking into the mitochondrial matrix
reaction by free Fe2⫹ (400). due to the mitochondrial membrane potential. A flickering
mode of mPTP may serve this purpose with the mPTP open-
ing for a time not sufficient for the onset of complete mito-
VI. ROS-INDUCED ROS RELEASE chondrial depolarization (239). Another support of a phys-
ASSOCIATED WITH THE mPTP iological role of the mPTP was obtained in CyPD knockout
mice which demonstrated an obvious maladaptive pheno-
type in their hearts (124). While the role of CyPD as a core
A. Fundamentals of the Discovery of the mPTP has been dismissed, there is a general con-
sensus that CyPD can serve a modulatory role in the
The importance of cellular redox homeostasis in progres- process of the mPTP induction. It was shown that CyPD
sion of inherited and acquired pathologies, including those knockout mice exhibit substantially greater cardiac hy-
associated with an aggressive oxidative environment, was pertrophy, fibrosis, and reduction in myocardial function
postulated elsewhere (reviewed in Refs. 72, 174). As we in response to pressure overload stimulation than control
discussed earlier, the redox homeostasis is determined by mice while cardiomyocyte-specific transgene expression
the balance between ROS generation matching metabolic of CyPD in these mice helped to rescue from the named
needs and ROS quenching capacity. Undoubtedly, tipping pathologies. Also, in mice lacking CyPD ischemic pre-
the balance in favor of increased ROS production within the conditioning was augmented (239) while mPTP openings
cellular microenvironment can severely alter the cellular in wild-type mitochondria were much more frequent
redox equilibrium, potentially resulting in oxidative stress than in mitochondria of knockout mice. This supports
which when mild can cause oxidation of essential mito- the notion that the mPTP might have an important phys-
chondrial components. In extreme cases it can irreversibly iological role, possibly through regulation of intramito-
damage these components resulting in a cell death. Within chondrial Ca2⫹.
the great diversity of types of cell death, which to date
comprise as much as 13 types (145), at least some of them The role of Ca2⫹ in the induction of the mPTP has been
could be associated with induction of the mPTP. already mentioned above; similarly, the role of oxidants in
generation of the mPTP pore is essential too, although both
mPTP opening is a phenomenon known in the field of mi- resulting in the same phenomenological outcome (255,
tochondrial research for many decades, but for the first time 277). It has been recognized that the mPTP induction rep-
described in details in a set of three consecutive papers by resents a highly complex phenomenon (322, 409). Particu-
Haworth and Hunter in 1979 (182, 195, 196). This phe- larly, isolated mitochondria exposed to Ca2⫹ plus Pi
nomenon [also recognized as an opening of a megachannel demonstrate the collapse of ⌬␺ preceding mitochondrial
(228, 229, 430)], originally studied in isolated mitochon- swelling, suggesting that the generation of a smaller, low-
dria, represents a sudden change in the permeability of the conductance pore occurs with permeability for ions but not
inner mitochondrial membrane allowing not only protons for solutes prior to induction of the mPTP (full size) pore (5,
but also other ions and solutes of a size up to ⬃1.5 kDa to 67, 244, 339). Oxidants such as hydrogen peroxides, or-
go through this membrane. There are many reviews on the ganic peroxides, and some other inducers generate both
tentative nature and identity of the mPTP (64, 489, 494) pores, but the insensitivity of induction of low-conductance
with details which are far beyond the scope of the present pore to Ca2⫹ and insensitivity of fully induced mPTP to
review. Previously, many candidates were considered to conventional inhibitor, cyclosporine A and sometimes to
serve as the core of the pore [i.e., mitochondrial VDAC, EGTA (65, 243, 262, 284), makes them different from clas-
cyclophilin D (cyPD), ANT (108, 170, 217, 351, 457, 489, sical pore inducers (255; reviewed in Ref. 489). Therefore,
494)], but largely they have been all dismissed because of the mitochondrial pore phenomenon appears to be multi-
various reasons, but still leaving for them important pore- faceted.
modulating functions. Recent evidence suggests that a
dimer of mitochondrial ATP-synthase is essential to form a Light is known to be a potent oxidant inducer when it
core of the mitochondrial pore (157) and c subunit of the interacts with photosensitizing agents, and such a property
mitochondrial ATP synthase complex may be required for is successfully utilized in photodynamic therapy (114, 317).
mPTP-dependent mitochondrial fragmentation and cell It is based on the ability of the excited fluorophore to gen-
death (56). erate primary ROS which after release inside of the biolog-

922 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

ical sample may become destructive for cellular compo- pound becomes reactive toward ROS, and after oxidation it
nents. In principle, many biological molecules and cell com- is transformed into fluorescing DCF with maximum excita-
partments can be fluorescently labeled and oxidatively tion in the blue region of the spectrum. The first scan of a
modified after interaction with excitation light. The mito- cardiac myocyte previously unexposed to light frequently
chondrion is an easy target for such modification since it reveals the area(s) in which mitochondria are expected to
can accumulate fluorescent cations to a great extent using occupy, but unstained with TMRM (suggesting that mito-
its intrinsic proton motive force, namely, its ⌬␺ (211, 298). chondrial ⌬␺ had collapsed in these mitochondria) but also
Mitochondrial ⌬␺ provides selectivity for photodynamic demonstrating high DCF fluorescence. The size of the area
action localized exclusively in mitochondria without poten- is dependent on the physiological status of the cell (FIGURE 7),
tial impact on other intracellular compartments (301, 303, possibly reflecting the cellular level of tolerance to ROS.
366, 367, 403). After exposure to light, the photosensitizer, Therefore, isolated adult rat cardiac myocytes are a conve-
e.g., tetramethyl rhodamine methyl ester (TMRM) which is nient model to study cellular oxidative stress.
a conventional probe for mitochondrial membrane poten-
tial widely used for visualization of energized mitochondria
in the cell, generates various ROS in water including very B. mPTP and Ischemia/Reperfusion Injury
strong oxidants such as superoxide anion radical and hy-
drogen anion radical (491). For the last 50 years it has been well recognized that coro-
nary reperfusion of infarcted myocardium is associated
Because mitochondria are critical intracellular loci of ROS with increased necrotic death of irreversibly injured cardiac
production, together with the fact that ROS exposure can myocytes. Jennings et al. (208) were first to report harmful,
lead to the mPTP, it was hypothesized that under certain both structural and functional, changes associated with rep-
circumstances the mPTP could become self-amplifying and erfusion. Therefore, early reperfusion while crucial for pre-
unstable (491). This hypothesis was tested in cardiac myo- serving ventricular function, preventing infarct expansion
cytes, taking advantage of the unique organization of mito- and potential development of heart failure, may also con-
chondria between myofilaments in an ordered three-dimen- tribute to the pathogenesis of reperfusion arrhythmias and
sional latticelike array forming straight lines thus allowing myocardial stunning manifested by reversible contractile
specific applications of line-scan confocal microscopy to dysfunction. The biological basis of reperfusion injury has
address this question (491, 493). For this purpose, rat car- been extensively studied since then and consequently the
diac myocytes are double stained with the probe for mito- notion of reperfusion as a double-edged sword was formu-
chondrial membrane potential (TMRM or ethyl derivative, lated in 1985 by Braunwald and Kloner (63). In 1986
TMRE having excitation maximum in a green region of the ischemic preconditioning was described as a means to ren-
spectrum) and the ROS probe, 2,7-dichlorodihydorfluores- der the heart more resistant to ischemia/reperfusion injury
cein diacetate (DCF-H2). DCF-H2 itself is nonfluorescing (311). The importance and the clinical potential of these
and unreactive toward oxidants; however, after base hydro- discoveries has prompted the research community to focus
lysis of the ester bonds, the resulting nonfluorescing com- on deciphering the molecular mechanisms that underlie rep-

CONTROL DIAZOXIDE HYPOXIC-PC

20 µm

FIGURE 7. ⌬␺ loss in a significant fraction


of mitochondria, caused by hypoxia/reoxy-
genation. Depolarized mitochondria (red-flu-
orescence ”holes“; bottom panels) are asso-
ciated with increased ROS (green; bottom
left panel). Hypoxic PC or pharmacological
(no PC) (no PC) HYPOXIC-PC + 5HD
preconditioning (PC), represented by diazox-
ide (Dz), prevents mitochondrial depolariza-
tion, and 5-hydroxydecanoate (5HD) accen-
tuates the loss. [From Hoffman et al. (189).]

TMRM ΔΨ
DCF (ROS)

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 923


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

erfusion-induced cellular injury as well as cardioprotection. mitochondrial Ca2⫹ uptake and consequent induction of
Thousands of research papers have already been published, mPTP (96, 97). This results in mitochondrial uncoupling
and a detailed picture has emerged, e.g., recently reviewed and activation of ATP hydrolysis by F1Fo-ATP-synthase
in References 181, 205, 207, 375. The significant role of (96, 97, 118).
so-called oxygen paradox was recognized (183) early on
and was based on the finding that substantial injury occurs The cardiomyocytes exposed to the hypoxia-reoxygenation
when molecular oxygen is reintroduced into ischemic tis- cycle have much larger areas occupied with fully deener-
sue. A burst of ROS production on reperfusion has been gized mitochondria and high levels of ROS, and these levels
detected (149, 500). Immediately after exposure of the were diminished and ⌬␺ was regained in most mitochon-
ischemic organ to oxygen, superoxide anion radicals dom- dria after ischemic or pharmacological preconditioning of
inate among the emerging ROS in the effluent perfusate, the cell (FIGURE 7) (218). Based on the previously expressed
with further formation of hydroxyl radical (499) pointing assumption that the cardiac ischemia-reperfusion injury is
to the occurrence of iron-mediated Fenton chemistry. This associated with the induction of the mPTP (97, 161), the
can be explained by the likelihood that in oxygen-depleted speculation was put forth that the found abnormal subcel-
medium the iron ions are mostly reduced (redox potential of lular loci in cardiac myocytes were indeed occupied by mi-
the couple Fe2⫹/Fe3⫹ ⫽ ⫹0.77 V), but since hydrogen per- tochondria which had undergone permeability transition.
oxide is absent, the Fenton reaction does not yet occur. The most striking detail was that mitochondrial deenergi-
Reperfusion ignites the formation of superoxide anions zation, if caused by the mPTP opening, was associated with
which dismutate resulting in formation of H2O2 and imme- higher, rather than lower, ROS production as follows from
diately reacting with iron ion [still in a reduced form FIGURE 5.
(Fe2⫹)], thus generating highly reactive ·OH (400).

Mitochondria have been implicated as a potential source as C. RIRR: Experimental Demonstration


well as a target of the generated ROS resulting in an ob-
served loss in mitochondrial function during ischemia/rep- To further explore this puzzling phenomenon, confocal mi-
erfusion and consequent irreversible cellular injury (9, 101, croscopy was employed in line-scan mode allowing re-
329). It has been noted that not only these radicals play a peated excitation of a row of mitochondria along the se-
significant role in the tissue damage observed following lected line in cardiac myocyte (FIGURE 8, A–C). Within the
ischemia/reperfusion but that this injury can be mitigated light-exposed mitochondria, a sudden loss of ⌬␺ with cor-
by oxygen radical scavengers (e.g., Ref. 213). In the early responding rise of ROS generation was revealed, suggest-
1990s, Crompton and colleagues demonstrated in isolated ing mPTP induction. Although these transitions were
mitochondria that the derangement of mitochondrial bio- modestly sensitive to cyclosporine A, they were dramat-
energetics that develop on reoxygenation, when resting cy- ically delayed by another mPTP inhibitor, bongkrekic
tosolic Ca2⫹ is high and ATP low, could lead to excessive acid, and by a ROS scavenger, Trolox. Furthermore, re-

A D 200

CALCEIN FLUORESCENCE
100
ΔΨm
TMRM FLUORESCENCE

175

B CALCEIN
150

C
50

125
MPT
50 sec

0 100
0 5 10 15
TIME (sec)
2 µm

FIGURE 8. RIRR and MPT induction. A: confocal microscope fluorescence imaging of cardiac myocytes
loaded with TMRM (red) and DCF-H2 (green). A, top: fluorescence image of TMRM-loaded cardiac myocyte.
B: enlarged portion of the TMRM-loaded cardiac myocyte. Line drawn on image shows position scanned for
experiment in bottom panel. C: row of mitochondria were line-scan imaged at 2 Hz with excitation at both 488
for DCF and 543 nm for TMRM and collecting simultaneous fluorescence emission at 510 –550 nm and ⬎560
nm, respectively. The sudden loss of ⌬␺ and rise of ROS generation in individual mitochondria are indicated by
the loss of TMRM fluorescence and increase in DCF fluorescence. B: cells dual-loaded with TMRM (⌬␺) and
calcein-AM (the latter loaded under conditions that results in a cytosolic distribution initially in excess over that
in mitochondria); line-scan imaging at 2 Hz. For details, see Refs. 491, 493.

924 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

distribution of the inert fluorescent probe calcein (mol wt The paradoxical aberration of established relations be-
620) from cytosol to mitochondria at the moment of tween ROS production and mitochondrial transmembrane
mPTP induction also confirmed opening of a pore perme- potential (240, 275) may have the same nature that we
able to a 620-Da compound (FIGURE 8B). Thus it was suggested when discussing the ROS production in com-
concluded that the collapse of the mitochondrial ⌬␺ oc- plexes I and II. We concluded that, at least partially, it is
curs due to the mPTP pore opening. determined by unusual conformational changes within
these complexes. It is known that at least some components
Fine analysis of the ROS burst kinetics associated with the demonstrate conformational changes within mitochondria
mPTP induction showed that ROS generation in the af- undergoing mPTP (171, 426, 427, 463, 464). Such confor-
fected mitochondrion proceeds in two distinct phases: the mational changes have been described, e.g., for ANT. It was
initial, slow rise due to the accumulation of a photochem- shown that the c-conformation of ANT (when ADP is
istry-associated production of ROS (called “trigger ROS”) bound to the cytosolic side) is more preferred for the open
and the subsequent ROS burst associated with dissipation state of the mPTP than the m-conformation (when ATP is
of the mitochondrial membrane potential. This biphasic bound to the matrix side), and it corresponds to the closed
process was named “ROS-induced ROS release” (abbrevi- state of the pore. Accordingly, atractylates stabilize the c-
ated RIRR). In addition to the ROS burst and the mPTP conformation of the ANT and promote the mPTP while
induction, it was accompanied by the burst of nitric oxide bongkrekic acid (184) stabilizes ANT in the m-conforma-
production; thus the term ROS-induced ROS/RNS release tion and inhibits the mPTP (172, 233, 320). However, ANT
where RNS stands for reactive nitrogen species is also rele- was proposed to be a pore modulator (494) rather than a
vant (491). component of the core of the pore (235).

The identical phenomenon using the same instrumentation Mitochondrial ROS rise simultaneously with the mPTP-
as in original RIRR study (491) was later described in cells induced drop of ⌬⌿. However, quite often it has been pos-
infected by adenovirus carrying mitochondria targeted cir- sible to observe a very brief phase of mitochondrial hyper-
cularly permuted yellow fluorescent protein (cpYFP) as the polarization coinciding with the onset of the mPTP open-
ROS sensor. The term superoxide flashes was used to de- ing. The analysis proved that the hyperpolarization phase
scribe the RIRR in transfected cardiac myocytes (461). Re- also coincided with the start of the excessive ROS genera-
cently, this approach was reexamined and challenged by tion (493) (FIGURE 9). Although a hyperpolarization spike
observation that the cpYFP is very sensitive to pH changes, has not been observed in all recorded mPTP opening events
therefore rendering this interpretation inconclusive (376, in the cardiac myocyte, the flicker could be sufficiently brief
377, 466). and be below the kinetic resolution of the fluorescent signal
from TMRM belonging to the class of “slow-response”
The complex I inhibitor rotenone (at low concentrations (distributed or accumulated) probes for the mitochondrial
not affecting TMRM sequestration) significantly decreased ⌬⌿ (458, 459). The flickering mitochondrial hyperpolarization
the ROS burst magnitude, confirming the mitochondrial mode was found to be a frequent phenomenon during pho-
nature of the ROS burst during mPTP-associated RIRR to-excitation of TMRM in loaded cardiac myocytes ob-
(491). The few second delay of NADH oxidation following served in the confocal microscope during line-scanning.
the collapse of the mitochondrial ⌬⌿ suggests that NADH Not every flicker could ignite the mPTP opening; never-
is the redox-energy store driving the electron donor neces- theless, each flicker initiated a small burst of ROS pro-
sary to support the single-electron reduction of molecular duction that was insensitive to the mPTP inhibitor bong-
oxygen that produces superoxide as an initiating radical for krekic acid. Also these hyperpolarization flickers were
generation of primary ROS followed by the mPTP induc- not associated with the entry of the inert probe calcein
tion. As we have already mentioned, the ROS burst accom- from the cytosol into the mitochondria, suggesting these
panying ⌬⌿ collapse is against the existing “dogma” that flickers do not involve a long opening of the mPTP. So,
ROS production is inversely related to the mitochondrial apparently there were at least two different modes of
membrane potential magnitude (FIGURE 5). This dogma RIRR, one of which was accompanied by the large-con-
seems to be true when the classical uncoupling is consid- ductor mPTP opening, while another was too brief to
ered, i.e., the drop of ⌬⌿ due to the enhanced proton con- establish the conductance. It is noteworthy that the flick-
ductance of the inner mitochondrial membrane. It seems to er-induced ROS production associated with mitochon-
be irrelevant to the extreme dysequilibrium situation when drial hyperpolarization theoretically obeyed the conven-
a megachannel is opened in this membrane. The enhanced tionally established relations between ROS production
ROS production in mitochondria undergoing mPTP open- and mitochondrial transmembrane potential (FIGURE 5)
ing has been confirmed in vitro when isolated mitochondria (240, 275), while the mPTP-associated ROS production
were supplemented with NADH to compensate for its loss did not. Both were caused by the triggering ROS formed
as a result of opening a megachannel (35). during the photoexcitation process.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 925


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

A B C D
Transient
hyperpolarization

* 300

250

TMRM Intensity (arb)


200

150

100
MPT
50

0
5 µm

0 5 10 15 20 25
Time (sec)
10 sec

ΔΨ
E H

F ROS
2 µm

1 sec
1 2 3 4 56 7
240 I
70
G 80

60 70 190

50
60
TMRM

1 2 3
TMRM

DCF

40 140
50
30
4
40
20 90
30 56 7
10 TMRM
DCF
0 20 40
0 1 2 3 50 100 150 200 250 300 350
Time (sec) Time (sec)
FIGURE 9. Transient ⌬␺ hyperpolarization and depolarization (flickering) preceding MPT induction can be
observed. A: overlay showing the 20-Hz linescan image in a cardiomyocyte loaded with TMRM (red) and DCF
(green). B: the TMRM channel from A; arrows indicate examples of transient hyperpolarization immediately
preceding ⌬␺ loss (MPT induction). C: the directional derivative of the TMRM intensity with respect to time, to
enhance the identification of intensity transients (transient hyperpolarization prior to MPT induction); arrows
indicate examples of transient hyperpolarization (positive slope maxima) immediately preceding MPT. D: intensity
plot of the TMRM responses of the mitochondrial pair indicated by the asterisk in B; inset shows an enlargement
of the area inside the red box and shows transient (relative) hyperpolarization immediately preceding MPT. [From
Zorov et al. (493), with permission from Elsevier.] E–G: ROS production (green) occurs during transient mitochon-
drial hyperpolarization flickering (red image, arrows in E). H–I: 2-Hz line scan image in a cardiomyocyte loaded with
TMRM. Transient hyperpolarization and depolarization spikes are seen as white and black vertical lines (shown by
the arrows from 1 to 7). TMRM fluorescence intensity (⌬␺) in the region of the cardiomyocyte denoted by the
arrow-bracket in H (⬃6 ␮m in width, consisting of 3 sarcomere-associated pairs of mitochondria) within which
mitochondria display concerted flickerings of the membrane potential. Note that these transient depolarizations,
during the time interval between arrows 1–7, are accompanied by progressive mitochondrial swelling (as seen by
the lateral displacement of adjacent mitochondria) consistent with repeated transient MPT-induction episodes.

926 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

VII. RIRR AND Ca2ⴙ Isolated mitochondria possess an ability to sequester very
large amounts of Ca2⫹ from the medium (85, 106, 259).
RIRR was named for the analogy with the process of Ca2⫹- Although proton motive force potentially drives cations to
induced Ca2⫹ release (with acronym CICR) which has been negatively charged mitochondrial matrix forcing cations to
known since the 1970s (125, 127, 128, 139) and which be accumulated inside of mitochondria, it has been found
plays a critical role in excitation-contraction coupling. In that at resting conditions in rat cardiac myocytes the in-
this process, an action potential depolarizes the cell mem- tramitochondrial free Ca2⫹ level is ⬍100 nM. Mitochon-
brane providing a small influx of Ca2⫹ through the plasma drial Ca2⫹ concentration is maintained by inward and out-
membrane, which by itself is insufficient to provide contrac- ward Ca2⫹-transporting systems in the inner mitochondrial
tile activation. Moreover, since the contractile apparatus is membrane. One of these, the uniporter (231, 372, 450,
at a distance from the sites of inward Ca2⫹ transport, Ca2⫹- 451), has low affinity but high capacity for transporting
induced muscle fibers contraction would be retarded if it Ca2⫹, whereas Ca2⫹/2H⫹ and Ca2⫹/2Na⫹ exchangers (60,
depended only on Ca2⫹ diffusion within the muscle cell. 61, 75, 98, 136, 137) have much lower capacity. In stimu-
CICR serves two goals. First, it significantly facilitates the lated cells at higher pacing rates, the increased cytosolic
Ca2⫹ will steadily raise the [Ca2⫹]m. Over the course of
propagation of the Ca2⫹ signal in the cell, and second, it
many contractions, [Ca2⫹]m can rise up to 600 nM (300)
amplifies Ca2⫹ concentration to reach Ca2⫹ levels in the
and potentially results in activation of mitochondrial dehy-
vicinity of contractile apparatus essential to provide a con-
drogenases. Pyruvate and ␣-ketoglutarate dehydrogenases
traction. Then, both RIRR and CICR contain signaling am-
show the greatest dependence on [Ca2⫹]m (107, 176, 278,
plification loops to reach the threshold necessary for the
302). Such activation of the key mitochondrial dehydroge-
transition. CICR and RIRR are also similar in spatial terms;
nases results in activation of respiration and ATP synthesis
CICR includes different intracellular compartments and
to meet rising energy demands under increased work load.
covers the space between the plasma membrane Ca2⫹-chan-
nels and ER/SR and fibers, and RIRR can span between
Interestingly, calcium ions target and activate those mito-
mitochondria and thus spread across and between cells.
chondrial enzymes (pyruvate dehydrogenase and ␣-keto-
glutarate dehydrogenase) that are considered to be a key
Although there have been some attempts in the past to
source of ROS in mitochondria (see above). However, while
ascribe mitochondrial mPTP to the existence of mitochon-
the crucial role of Ca2⫹ in mitochondrial metabolism is
drial CICR (194) or even strontium-induced strontium re-
established (reviewed in Refs. 166, 434) with recent molec-
lease (190), eventually the essential role of ROS has been ular identification of the Ca2⫹-uniporter (36, 105) and
established in this process. In fact, mitochondria have quite Ca2⫹/2Na⫹ exchanger (332), the role of Ca2⫹ in ROS pro-
a high Ca2⫹-buffering capacity (maximal calcium uptake), duction in mitochondria remains controversial. In isolated
explaining the fact that a number of consecutive Ca2⫹ mitochondria, the mPTP is opened by elevated Ca2⫹. This
pulses could be sequestered and tolerated by isolated mito- led to the conventional assumption that mPTP induction by
chondria or permeabilized cells until mPTP opening occurs mitochondrial Ca2⫹ loading may be the cause of many
resulting in a robust release from mitochondria of both
sequestered and intrinsic amounts of Ca2⫹ (or Sr2⫹) (77, 8

166, 321, 331). Such mitochondrial ability to sequester 50 7 30


EVENT RATE
(100 µm · s)-1

Ca2⫹ could be modified by a number of factors (7, 30, 132, Ca2+ Sparks
5
4

FLUO-3 Fluorescence
3
(FLUO-3)
306, 465). 40
TMRM Fluorescence

2
1 25
0
Control MPT-
Proximity
Consequently, “X-induced X release” may be a general 30

mechanism for any determined (such as RIRR and CICR) 20

and still undetermined signaling amplification loops. Mean- 20

while, reciprocal amplification of ROS and Ca2⫹ signals or 15


any other “X-induced Y” release processes representing soli- 10 ΔΨm
tary events or additional complex processes within a signal- (TMRM)

ing cascade, probably also occur, e.g., in the case when ROS 0 10
0 0.5 1 1.5 2 2.5 3 3.5
generated in mitochondria may be a source of Ca2⫹ unitary Time (sec)
releases from ryanodine receptor of SR (Ca2⫹-sparks) (491)
FIGURE 10. Induction of Ca2⫹ sparks after the mPTP. Cell is
(see FIGURE 10). The process of the spontaneous ROS-in- dual-loaded with TMRM (⌬␺) and fluo-3 (Ca2⫹) and line-scan imaged
duced Ca2⫹ release by the SR ryanodine receptor described at 230 Hz. Representative example showing the dissipation of
in Reference 491 has been recently analyzed in a model of TMRM fluorescence from a single mitochondrion and a cluster of
sustained mitochondrial ⌬⌿ oscillations driven by the mi- Ca2⫹ sparks in the immediate vicinity, within seconds of MPT induc-
tion. Inset: comparison of Ca2⫹ spark rate in proximity of the mPTP
tochondrially produced ROS. Dynamic changes in mito- occurrence, i.e., within the sarcomere containing the involved mito-
chondrial energy state resulted in altered frequency and chondria and within 3 s after the mPTP occurrence. [From Zorov et
properties of the Ca2⫹ sparks (485). al. (491).]

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 927


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

types of cell death, such as those induced by ischemia-rep- A 1.4


erfusion injury and chemical toxins, in the heart and brain.
1.2
In intact cardiomyocytes and neurons, however, normal

tMPT (normalized)
excitability achieves Ca2⫹ elevations to levels comparable 1.0

0.8
to those believed to induce the mPTP based on in vitro data.
This suggests a serious paradox that will have to be recon- 0.6

ciled regarding the role of Ca2⫹ in mPTP induction in vitro 0.4

with the empirical evidence of health and longevity of these 0.2

postmitotic cells in vivo (218, 494). Although Ca2⫹ is a 0.0


1 Ca2+ (rest) 5 Ca2+ 100 nM Ca2+ 500 nM Ca2+
typical tool for induction of permeability transition in iso- (10 HzTet)
lated mitochondria (182, 195, 196; reviewed in Ref. 489),
Intact SL Perm SL
experiments with intact cardiac myocytes and neurons
demonstrated that the mPTP is largely insensitive to in-
B
creased cytosolic Ca2⫹ (see supplements to Ref. 218). In 1.2

these experiments, however, after cell permeabilization, the 1.0


Ca2⫹ sensitivity of the mPTP is unmasked and is similar to

tMPT (normalized)
0.8
that observed in mitochondria isolated from these cells (FIG-
URE 11, A AND B). Probably, some cytosolic factors that are 0.6
lost after permeabilization and mitochondrial isolation are
responsible for Ca2⫹ insensitivity of the mPTP in the intact 0.4

excitable cell such as the cardiac myocyte or neuron. These 0.2


facts themselves put into question the role of Ca2⫹ in cell
0.0
death correlated with mPTP induction after, e.g., hypoxia/ 1 Ca2+ (pre Sr2+) 1 Sr2+ (1 hr) 1 Sr2+ (6 hr) 1 Ca2+ (post Sr2+)
reoxygenation which may be overestimated due to limita-
tion of the biological model (212) (see FIGURE 11C). C 100
90
There is also a puzzling discrepancy between a very high 80
% CELLS DEAD

apparent Km (250 –350 ␮M) for endogenous ADP in the 70


control of mitochondrial respiration in permeabilized mus- 60
cle cells and that in isolated mitochondria where the appar- 50
ent Km equals 15–20 ␮M (248). One of the possible expla- 40
nations for this inconsistency is that isolated mitochondria 30
are stripped from cytoskeletal proteins that maintain mito- 20
chondrial integrity within the intracellular energetic unit, 10
which might be involved in fine regulation of enzymatic 0
O2 -- Bapta O Ca2+
parameters in the live cell and be responsible for observed
differences (22). N2
2⫹
FIGURE 11. Ca dependence of the mPTP in rat adult cardiac
myocytes. A: mPTP ROS threshold during a “clamp” of intracellular
VIII. MITOCHONDRIAL Ca2⫹ ⬎500nM for ⬎20 min (using 10-Hz electrical tetanization in 5
COMPARTMENTATION mM bathing Ca2⫹ in the presence of thapsigargin or cyclopiazonic
acid to disable the sarcoplasmic reticulum) is identical to that in
unstimulated cells (with a cytoplasmic Ca2⫹ of ⬃100 nM). B: com-
The reviewed data raise the question about the adequacy of plete equimolar replacement of Ca2⫹ for Sr2⫹ (for 6 h) in intact
isolated mitochondria or permeabilized cells to the actual cardiac myocytes resulted in the same mPTP ROS threshold as seen
in cells with normal Ca2⫹. C: buffering intracellular Ca2⫹ (with
processes taking place in an intact live cell, as all three
BAPTA) or limiting Ca2⫹ influx (in nominally Ca2⫹-free buffer) does
objects have different levels of structural and organizational not limit cardiac myocyte death after hypoxia/reoxygenation. [From
complexity. One of the structural parameters missing in Juhaszova et al. (218). Republished with permission from the Amer-
isolated mitochondria is a close proximity of mitochondria ican Society for Clinical Investigation; permission conveyed through
Copyright Clearance Center, Inc.]
to ER/SR found in the cell (119, 379, 384), thus making
easier shuttling of small signaling elements (e.g., ROS, H⫹,
or Ca2⫹) and tightening the spatial and functional compart- greatly diminish the diffusion-controlled step(s) when the
mentation of two organelles (357). The loss of spatial com- product of one enzymatic reaction almost immediately be-
partmentation due to the isolation procedure may dramat- comes a substrate of the second enzyme without the step of
ically change kinetics and, thus, enzymatic parameters of a product release in the bulk phase followed by extraction
interactive systems. The compartmentalized system might from this phase.

928 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

It is beyond the scope of this review, but the dramatic dif-


ference between two theories of oxidative phosphorylation,
one proposed by Peter Mitchell and another by Robert
Williams, deserves mentioning. Both theories were
named chemiosmotic with one principal difference: while
Mitchell’s theory considered the delocalized proton ejected
from a proton pump into the bulk phase to be a substrate
for ATP synthase complex (298, 299), Williams considered
a localized proton to be consumed by ATP synthase without
an intermediate step of going into the bulk phase (472,
473). Although there are some more or less sketchy obser-
vations supporting the second mechanism which appar-
ently requires organization of supramolecular protein com-
plexes (reviewed in Ref. 131), the first mechanism was gen-
erally accepted, and its founder was awarded the Nobel
prize in 1978.

The problem of intramitochondrial compartmentation has


many supporters (e.g., Refs. 412, 413; reviewed in Ref.
265). However, recent data has led to an interpretation
highly critical of the organization of mitochondrial respira-
tory chain into supramolecular complexes (organized pro-
tein assemblies) (441), leaving the compartmentation issue FIGURE 12. Oscillations of ⌬␺ in individual mitochondria over time
still rather controversial. in a portion of a cardiac myocyte induced by 10 ␮M clotrimazole.
Panels of consecutive fluorescence frames with a different interval
indicated in each plane 10 s acquired from either a 18 ⫻ 15 ␮m
Another important issue is the intermitochondrial compart- region of an adult cardiac myocyte [A; from Zorov et al. (493), with
mentation. The possibility that mitochondrial units may permission from Elsevier.] or neonatal cardiac myocytes loaded with
work in concert unifying energy transmission and signaling TMRM (B). Arrows indicate individual mitochondria or mitochondrial
has been highly explored. Concerted work of individual clusters displaying periodic changes in the membrane potential.
mitochondria in striated muscles interconnected by electri-
cally permeable junctions (28) was proposed (394) and later sealed and ⌬␺ regained with many repetitive cycles of the
confirmed for cardiac myocytes (10). Such electrical unifi- mPTP induction and closure. The first demonstration in
cation of mitochondria was postulated to adequately allow intact cells of such an oscillatory mode of the mPTP trig-
fast transportation of energy along the mitochondrial retic- gered by photodynamically induced ROS accompanied by
ulum to all cellular regions remote from the initial energy the RIRR was given in 2000 (491) and further explored in
source. Later the mitochondrial matrix lumen continuity 2006 (493).
was confirmed by using photoactivated GFP (445), sup-
porting the idea of mitochondrial organization in networks The regulation of mitochondrial ROS production and their
(115, 199, 395) which play an important role in intracellu- levels is exerted by a number of factors, such as the redox
lar signaling (46, 215, 395). Cooperation of mitochondria state of respiratory components, oxygen tension, ionic en-
in terms of synchronous response to oxidative challenge as vironment and the activity of redox buffers, etc. However,
part of the RIRR process seems to open a new door for the mitochondrial environment may become unstable,
exploring alternative roles of mitochondria in the cell apart which may cause the instability of mitochondrial function-
from their energetic function (491, 493). ing, and potential episodes of fluctuating mitochondrial
ROS production are not an exception. Several researchers
described synchronized mitochondrial oscillations includ-
IX. MITOCHONDRIAL RIRR IN ing rhythmic changes of ion fluxes, respiration, and mito-
OSCILLATING MODE chondrial volume in vitro in suspensions of isolated mito-
chondria (particularly, under conditions of energized ion
It is noteworthy that fluctuations of ROS within the cell in transport) (23, 83, 158, 168, 198, 312, 330). Typically,
the vicinity of mitochondria can elicit instability of the in- these oscillations had the form of damped sinusoidal rhyth-
tracellular redox state which, as we pointed out earlier, is mic changes. Under these conditions, within each cycle, the
greatly maintained by homeostatic mitochondrial functioning. mitochondrial shrinkage phase was associated with in-
FIGURE 12 demonstrates ROS-induced fluctuations of the creased respiration, a more highly oxidized state of pyridine
mitochondrial ⌬␺ (one recorded in adult and another in nucleotides, a stimulation of ATP hydrolysis, an inhibition
neonatal cardiac myocyte) showing that mPTP in these cells of proton release, and stimulation of cation release. The
can be reversed and the mitochondrial inner membrane be shrinkage phase is followed by a swelling phase, and when

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 929


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

these changes were reversed, it also showed evidence of fore, the mPTP is either too transient for sufficient move-
damping. It was concluded that the oscillatory states of ment of calcein to be measured, or of low conductance, or
electron and energy transfer pathways might be under the possibly even not involved in this type of oscillatory mode.
control of mitochondrial swelling and ion transport by a In a comprehensive search for ion channels responsible for
feedback mechanism. the various types of non-mPTP-related mitochondrial oscil-
lations, the focus moved to the so-called IMAC because the
Later, such oscillatory behavior of different mitochon- oscillations were blocked by the IMAC inhibitor DIDS (38 –
drial parameters including those induced by ROS has 40). In addition, these oscillations were significantly atten-
been detected in intact cells and tissues (18, 92, 126, 281, uated by ligands of the mitochondrial (peripheral) benzodi-
297, 360, 491, 493) showing that structural peculiarities azepine receptor (PBR) such as Ro5– 4864 and PK11195
in organization of mitochondria within the cell do not (18, 323). It has been speculated that PBR (which presum-
play the most important role in induction and propaga- ably consists of ANT, VDAC, and 18-kDa protein of the
tion of oscillations. outer mitochondrial membrane, TSPO) could be responsi-
ble for IMAC (20) since some ligands of PBR can block the
mitochondrial inner membrane 107-pS channel (230)
X. IMAC-ASSOCIATED RIRR
which has moderate anion selectivity (410). However, PBR
ligands (230) as well as the mPTP inhibitor cyclosporine A
Another model of RIRR was generated and explored in B. (430) both block mitochondrial giant MCC (multiple con-
O’Rourke’s lab at Johns Hopkins University. The study was ductance channel), implying that mPTP might be related to
preceded by an investigation of oscillations in the sarcolem-
the functional state of PBR. However, this speculation
mal current in cardiac myocytes caused by substrate depri-
needs additional experimental support with elucidation of
vation resulting in spontaneous fluctuations of ATP-sensi-
the molecular identity of IMAC.
tive K⫹ current in parallel with changes of the redox state of
NADH/NAD (324). This implies the critical role of energy
Thus, in addition to mPTP-associated RIRR, another in-
in generating oscillations. Since metabolic oscillations were
dependent mode was proposed for RIRR which involves
in concert with the shortening and suppression of the action
IMAC. According to this model, mitochondrial respira-
potential, it has been concluded that these metabolic oscil-
tory chain is the main oscillatory source of ROS which
lations may be relevant to pathologies such as arrhythmias
can be released to the cytosol through IMAC due to its
as a result of ischemia/reperfusion insult (reviewed in Refs.
permeability for superoxide (449) produced in complex
20, 323). Later, it was found that a local two-photon exci-
tation of cardiac myocyte covering a few mitochondria III (FIGURE 13).
loaded with a mitochondrial membrane potential-sensitive
fluorescent dye triggered oscillation of the membrane po- Recently, the ⌬␺ oscillations and RIRR induced by local
tential within these mitochondria. It also triggered the os- oxidative stress (401) and by perfusion with hydrogen per-
cillations in intramitochondrial redox potential measured oxide (53) have been detected in live, intact myocardium.
by a ratio of reduced-to-oxidized flavins. The igniting im- Perfusion of rat hearts with hydrogen peroxide, depending
pulse was triggered by ROS generated by the local excita- on the concentration of the oxidant, elicited two peaks of
tion resulting in production of primary ROS. The oscilla- superoxide in live myocardium measured by optical map-
tions spread in three dimensions along all interconnected ping of the fluorescent ROS probe, with the second peak
mitochondria forming a lattice with the primary oscillators being much more intense than the first one (53). Spatiotem-
consisting of only a few mitochondria. These could later poral ROS mapping during the second ROS peak revealed a
spread over the entire mitochondrial network and possibly propagation of superoxide signal within the cardiac tissue
extend beyond the boundaries of the single cell. with a velocity ⬃20 ␮m/s. The hearts with the second peak
displayed a much greater arrhythmia compared with those
Mitochondrial inhibitors such as rotenone and bongkrekic where the second peak was absent. PBR ligand, Ro5– 4864,
acid suppressed mPTP-associated RIRR and ROS to the and superoxide dismutase mimetic but not cyclosporine A
level below the threshold implicating the source of ROS as abolished RIRR and ventricular tachycardia and ventricu-
mitochondrial (491). However, these inhibitors as well as lar fibrillation, suggesting IMAC involvement. These find-
others including cyclosporine A, cyanide, myxothiazol, ni- ings further extended the concept of RIRR as a key factor in
gericin, and oligomycin did not block RIRR associated with the incidence of postischemic arrhythmias.
transient mitochondrial hyperpolarizations (mitochondrial
membrane potential flickers described above) (FIGURE 9) A computational model with superoxide as a trigger of
(493). In contrast to mPTP-associated RIRR, in the RIRR mitochondrial ⌬␺ depolarization has been developed (484,
mode associated with hyperpolarization flickers, the brief 486). The model is based on the percolation theory to ex-
changes of the membrane potential were not accompanied plain how neighbor-to-neighbor interaction defines propa-
by significant redistribution of an inert fluorescent agent of gation of the signals along the spatially organized excitable
620 Da (calcein) between mitochondria and cytosol. There- matrix, arising from a spanning cluster of oxidatively

930 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

stressed mitochondria united in the network (19, 20). In chondrial instability caused by the introduction of oxygen
addition, the dynamic spatiotemporal properties of individ- into the system. In these cells, transient mitochondrial hy-
ual mitochondrial oscillators within stress-induced syn- perpolarization was followed by a progressive rise of the
chronized oscillatory clusters of cardiac mitochondrial net- ROS level in deteriorating mitochondria to the point of
work have been characterized by applying wavelet-based mPTP introduction and accompanying large ROS burst
analysis (252, 253). Individual oscillating mitochondria (FIGURE 14A), i.e., typical of RIRR. The incidence of cardiac
within clusters have been identified and their frequency cor- cell death depended on the proportion of damaged mito-
related with the size of the cluster. As the cluster grew chondria having developed mPTP to their total number
larger, they required more time to achieve full synchroniza- in the cell and on the physiological status of the cell (218)
tion of all mitochondria within the cluster. Additionally, the (FIGURE 14B).
percentage of cluster size has been inversely correlated with
the percentage of amplitude. Nonlinear characteristics of To assess the degree of system readiness to resist an oxida-
the mitochondrial network, particularly in the heart, make tive challenge, a simple approach was developed using time
mitochondria prone to the local disturbances including as a factor. Time quantifies the titrated amount of ROS
those involving oxidative stress. Once these local distur- (delivered incrementally by successive photoactivated expo-
bances are formed, they spread over the network causing sures during linescan imaging) required to achieve the
loss of its normal functioning. They cause a mismatch of the threshold for mPTP induction in a particular mitochon-
finely tuned balance of available energy production to en- drion (218, 491). Apparently, mitochondria with higher
ergy expenditures (93, 365, 435, 479), resulting in an in- resistance to oxidative stress require a longer time (i.e.,
ability of mitochondrial network to completely repolarize higher total oxidant exposure level) to open the pore while
between oscillations. Finally, it reaches a point of no return those with a low tolerance require shorter time (i.e., lower
that results in pathologies and ultimately cell death. total oxidant exposure level) for the mPTP induction. The
timing of mPTP opening (tmPTP) is an integral factor, de-
XI. RIRR-RELATED PATHOLOGIES: THE pending on the rate of oxidant quenching (the level of intra-
IMPORTANCE OF ⌬␺ HOMEOSTASIS and extramitochondrial antioxidants, e.g., glutathione) and
the physicochemical state of the system responsible for the
Mitochondria are not only the energy source but also one of generation of the mPTP as an event. While the molecular
the primary sources of a vast number of deadly pathologies. identity of the mPTP needs to be validated in detail (157,
As to cardiovascular problems, a striking example of “mis- 217, 494), the presence of bcl-2/bcl-xL or phosphorylated
behavior” of a mitochondrial network extending over the glycogen synthase kinase 3␤ (P-GSK-3␤) in the vicinity of
whole heart and affecting the whole organism is ischemia/ the pore or attachment of hexokinase II desensitize the pore
reperfusion-induced oxidative injury which in extreme to oxidants (86, 217, 218) and extend the time necessary to
cases results not only in cell death, but also organ failure, overcome the desensitization and open the pore. The con-
and eventually organismal death. Experiments with single ventional line-scan confocal microscopy has been employed
cardiac myocytes exposed to transient hypoxia, followed by as a model to determine the ROS threshold for inducing the
a reoxygenation phase, demonstrated substantial mito- mPTP. For this purpose, an ideal experimental object is a

FIGURE 13. Hypothetical scheme of the


IMAC-associated ROS-induced ROS release
mechanism underlying the synchronization
of mitochondrial activity during the oscilla-
tions. [Modified from Aon et al. (18), with
permission from The American Society for
Biochemistry and Molecular Biology.]

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 931


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

FIGURE 14. ROS are involved in mitochondria and cell deterioration. A: hypoxia/reoxygenation-induced
mitochondrial hyperpolarization leads to increased ROS. Mitochondria stained with TMRM (⌬␺, red) and DCF
(ROS, green) were laser line-scanned (2 Hz) during hypoxia and the reoxygenation phase. The ROS burst is
delayed after reoxygenation and starts at the maximum ⌬␺. Mitochondrial hyperpolarization lasts for ⬃2 min,
followed by loss of ⌬␺. B: cell survival after constant-energy photoexcitation of a 25 ⫻ 25 ␮m2 region. Right
panels show TMRM-stained cardiomyocyte (red) immediately after, and 1 h after, irradiation. Survival is
inversely related to the fraction of mitochondria (mito) undergoing mPTP induction and is improved by ROS
scavenger (Trolox), NO donor (SNAP), and the KATP channel opener diazoxide (Dz) and is impaired by 5-hy-
droxydecanoate (5HD), the blocker of KATP channel. [From Juhaszova et al. (218). Republished with permission
from the American Society for Clinical Investigation; permission conveyed through Copyright Clearance Center,
Inc.]

cardiac myocyte with mitochondria arranged in regular lat- bated in media with low O2, tmPTP was much greater than
ticelike arrays rendering the line-scan instrumentation in mitochondria incubated in normoxic or hyperoxic me-
straightforward (491, 493, 496) (FIGURE 15A). Further- dia (FIGURE 15B). Consequently, adequate oxygenation
more, the same approach has been successfully applied to of cellular mitochondria is important for unambiguous
neuronal mitochondria in situ and cultured neonatal car- interpretation of tmPTP.
diac myocytes (217) and potentially could be applied to
mitochondria in other cell types with cell motility or intra- The value of tmPTP determined by the above-mentioned
cellular mitochondrial motion during the line-can proce- approach, the inherent resistance of mitochondria to
dure being carefully considered. mPTP induction by ROS, and can be significantly pro-
longed by ischemic or pharmacological preconditioning.
It is not surprising that the mPTP threshold is highly This approach has been applied for the detailed elucida-
dependent on ambient molecular oxygen, since ROS pro- tion of the architecture of the signaling pathways en-
duction in the system is a reaction of the first order with gaged by ischemic/pharmacological preconditioning
respect to molecular oxygen. In cardiac myocytes incu- (218, 385).

FIGURE 15. Linescan technique applied to mitochondria in rat cardiac myocytes. A: definition of the mPTP
ROS threshold as an average time over population of mitochondria (shown by a dashed red line) in myocyte
stained with fluorescent mitochondrial probe for the membrane potential to induce the mPTP (shown by arrow)
(491, 493). B: dependence of the mPTP induction in the cell on ambient molecular oxygen.

932 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

MPT with long


mean open time

A
tMPT

FIGURE 16. Basic mitochondrial redox states


MPT with short MPT in oscillation
(see explanation in the text). A: unstressed mito-
mean open time mode
chondria with high tolerance to ROS (high sur-
tMPT B C vival). B: stressed mitochondria with low tolerance
to ROS (low survival). C: stressed mitochondria
with high tolerance to ROS (unknown survival).

DEATH

The term mitochondrial criticality was coined (16, 17) for mPTP induction, mitochondria in state C are characterized
the situation when the mitochondrial network in the car- by dynamic instability revealed by ⌬␺ oscillations which
diac myocyte becomes highly sensitive to the changes of progressively deteriorate and ultimately cease. These differ-
ambient/intramitochondrial ROS, for example, by respond- ences probably reflect not only differences in the state of the
ing as synchronized oscillatory depolarizations of the mito- antioxidant defense system in B versus C but also intrinsic
chondrial ⌬␺ which could be scaled to the whole cell or differences in the function/activities of the pore complex
cluster of myocytes (401). When the mitochondrial system subunits itself triggered in response to posttranslational
reaches the point of criticality, oscillations of ⌬␺ are ignited, modifications, the signaling status of the immediate pore
eventually creating spatial and temporal heterogeneity of modulators, and/or changes in the pore complex arrange-
excitability in a heart exposed to oxidative challenge, e.g., ment at the protein level (FIGURE 16).
under hypoxia/reoxygenation. This inflicts dramatic path-
ological changes to the heart involving lethal cardiac ar- A preponderance of data point to the fact that mitochon-
rhythmias possibly followed by cardiac arrest and an indi- drial membrane potential is an essential attribute of mito-
vidual’s death. chondria, and its homeostasis is a prerequisite for the health
of mitochondria and the preservation of normal cell and
Based on the discussion above, we would tentatively suggest tissue function. Under hypoxic conditions, when mitochon-
the existence of specific in vivo mitochondrial states that dria are incapable of sustaining ⌬⌿ driven by respiration,
may contribute to the fate of the cell (see FIGURE 16). The the mitochondrial membrane potential is maintained at the
relatively stable state (state A) has a comparatively high expense of cytosolic ATP hydrolysis by mitochondrial
tolerance to oxidative stress and is able to sustain the ⌬␺ for ATPase (110). Sufficient buffering of the mitochondrial ⌬␺
a longer time (with longer tmPTP). This is characteristic of by cytosolic ATP may explain why the intracellular concen-
the healthy cell where the mPTP formation is delayed. States tration of ATP (a few mM) exceeds the Km for intracellular
B and C characterize mitochondria which have been ATP-utilizing enzymes by orders of magnitude. Thus, under
stressed (prior to experimental procedure of the mPTP-in- physiological conditions, these enzymes are fully saturated
duction protocol has been employed) and thus are unable to and do not depend on ATP concentration for their function.
sustain the ROS challenge and to maintain their ⌬␺ for an
extended period. In both cases (B and C), the first mean The homeostasis of mitochondrial ⌬␺ is critical for the mi-
tmPTP value is similar, characterized by a shorter tmPTP com- tochondrial function. Only small vital variations in a range
pared with mitochondria in state A, but there are some (approximately between 160 and 220 mV) were tolerated in
obvious and significant differences between the states B and isolated mitochondria at states 3 and 4 according to the
C. While in state B mitochondria remain depolarized after terminology of B. Chance (82). These small variations pre-

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 933


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

FIGURE 17. Tentative model of regulation of the mitochondrial quality control by ⌬␺. Parkin is an E3 ubiquitin
ligase located in the cytosol, while PINK1 is a mitochondrial kinase carrying mitochondrial import sequence to
be accumulated inside of mitochondria followed by degradation by PARL. ⌬␺ drives PINK1 into mitochondria
where it is processed by mitochondrial processing protease (MMP). When mitochondrial membrane potential
is dissipated, PINK1 accumulates on the outer mitochondrial membrane, where it can recruit Parkin to
impaired mitochondria followed by its ubiquitination as a starting point for mitophagy. OM, outer membrane;
IM, inner membrane. TOM and TIM are transport outer and inner membrane proteins correspondingly. (From
Ref. 210: Copyright 2010 Jin et al. Originally published in Journal of Cell Biology. 191: 933–942.
doi:10.1083/jcb.201008084.)

serve the minimal energy that may be sufficient for phos- remain unknown. One possible candidate for such a process
phorylation of ADP by mitochondrial ATP synthase but is the energy (⌬␺)-dependent transport of peptides into mi-
may dramatically influence ROS production within mito- tochondria (reviewed in Ref. 338). This machinery is highly
chondria (see FIGURE 5). In case of a drop of the membrane important for the quality control of mitochondria (362)
potential below the phosphorylating membrane potential, through the ⌬␺-dependent ubiquitin labeling of the im-
especially when it is nulled after induction of mPTP, mito- paired mitochondrion resulting in the initiation of mi-
chondria could initiate a process of self-destruction involv- tophagy (see explanations in the FIGURE 17; Ref. 210).
ing Ca2⫹-independent degrading hydrolases and phospho-
lipases (66, 261, 495, 496). This process is called mitophagy
(reviewed in Refs. 88, 234, 258) and further substantiates XII. PRACTICAL ASPECTS OF USING
the vital importance of maintaining ⌬␺ homeostasis to keep MITOCHONDRIAL ⌬␺ AS A DRIVER
mitochondria structurally and functionally intact. The AND REPORTER
questions regarding a mitochondrial ⌬␺ sensor, or which
processes in mitochondria could be considered as vital, the The mitochondrial membrane potential drives natural cat-
cessation of which will result in mitochondrial degradation, ionic substances into mitochondria, and this property could

934 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

be utilized for intramitochondrial transport of man-made received support from experiments wherein 1) the primary
drugs. extramitochondrial ROS source induced a secondary in-
tramitochondrial ROS release, and 2) both of these pro-
In principle, the delivery of drugs to the mitochondrial in- cesses were prevented by specific mitochondria-targeted an-
terior can be executed by using two targeting vehicles: 1) tioxidants (344).
mitochondrial targeting signal sequences and 2) mitochon-
driaphillic compounds. While the first approach implies
specific transport of the vehicle through the TIM/TOM mi- XIII. CONCLUDING COMMENTS
tochondrial protein transport system (318), the second uses
positively charged lipophilic compounds, for example, con- ROS-induced ROS release is a fundamental process origi-
structed in a way to delocalize the charge over the set of nating in mitochondria (18, 491, 493) responding to an
coupled double bonds (268). Due to negative charge of the increased oxidative stress by a positive feedback loop result-
mitochondrial interior, the latter nonspecifically accumu- ing in a regenerative, autocatalytic cascade which in a vast
lates these cations to the thermodynamically permissive majority of cases is terminated, presumably by the local
level which theoretically exceeds extramitochondrial levels redox environment together with a change of functional
by about three orders of magnitude. The idea of driving properties of the “release valve.” Indeed, for this mecha-
such charged compounds by virtue of the electrogenic na- nism to serve a physiological role, its underlying properties
ture of mitochondrial energetics is fully consistent with of chain-reaction-like propagation must be efficiently ter-
Mitchell’s chemiosmotic theory (268, 298, 299). Nowa- minated to prevent the unwanted spread and widespread
days, after establishment of this mechanism (1), the ap- destruction of mitochondria risking unintended and un-
wanted loss of the cell. These considerations are entirely
proach started to serve two experimental missions: to de-
analogous to the control of a nuclear chain reaction in
liver to the mitochondrial interior a desired compound
power plants (e.g., by control rods) or that of CICR in
which is bound to a mitochondriatropic vehicle (225, 393,
excitable cells by intrinsic ryanodine receptor- and SR-re-
407) and, on the other hand, to discriminate between the
lated mechanisms (423). When RIRR is appropriately not
mitochondrial and nonmitochondrial nature of some pro-
terminated, it may lead to the wanted removal of a mito-
cess under study. In this context, mitochondriatropic anti-
chondrion or cell, e.g., damaged beyond repair. When
oxidants are not only potentially beneficially targeted drugs
RIRR is inappropriately not terminated, it may lead to un-
to quench unwanted excessive ROS and other reactive spe-
wanted cell loss such as after stroke or myocardial infarc-
cies in mitochondria, but they may also serve as excellent
tion.
reporters providing an answer as to whether a given mech-
anism involves mitochondrial ROS. In the first case, in vitro We propose that RIRR may result in a cellular response that
data and animal experiments have suggested the possible can be considered either adaptive (e.g., promoting signaling
beneficial effect of mitochondria-targeted antioxidants resulting in the elevation of the tolerance to oxidative stress
(such as MitoQ, SkQ1, SkQR1, and others) in treatment of or the removal of unwanted organelles and cells) or mal-
a great number of model systems associated with oxidative adaptive (e.g., the pathological destruction and death of
stress. These include stroke, arrhythmias, ischemia/reoxy- functional mitochondria and vital, essential cells). Adaptive
genation, chemical toxicity, infection, inflammation, and cell elimination may involve cell death executed by necrotic
some inherited and acquired age-related and unrelated dis- or apoptotic mechanisms (for example, elimination of non-
eases (such as Alzheimer’s, diabetes, hepatitis, metabolic functional or damaged cells). It remains an open question
syndrome, hearing loss, etc.) (14, 27, 146, 206, 223, 236, whether RIRR mechanisms would be recruited to serve the
269, 292, 307, 343, 388, 391, 399, 404 – 406, 408). adaptive elimination of cells necessary to achieve proper
organ size, shape and functions related to embryogenesis,
Apart from the practical issue of using mitochondria-tar- self-antigen-recognizing immune cells that would be dan-
geted chimeric compounds as potential therapeutic agents, gerous, cells no longer needed (e.g., nonlactating breast in-
their specific effects uncovered the fundamental role of mi- volution after pregnancy), activated lymphocytes after in-
tochondria and mitochondrial ROS in the onset and prop- fection is resolved, or pre-cancer and cancer cells.
agation of different pathologies. What we have learned is
that ROS originating from mitochondrial rather than from The RIRR phenomenon provides an opportunity to see the
other intracellular sources, when they exceed the homeo- mitochondrion as machinery designed not only to trigger
static level, are among the most pathogenic factors (222, but also to facilitate important signaling mechanisms, fur-
382, 383, 437, 467). In this context, targeted normalization ther extending the concept of the mitochondrion as an ex-
of ROS levels (specifically in mitochondria) may be benefi- citable organelle capable of generating and conveying re-
cial in preventing a number of ROS-related pathologies in- dox, electrical (10), and Ca2⫹ (199) signals. Therefore,
cluding that involving RIRR (310, 392). The requirement of mitochondria are not only efficient cellular energy power-
mitochondria-derived ROS for the propagation of RIRR houses, they are significant signal transmitters and signal
together with its concomitant pathological cell death has resonators, by which we mean the mitochondrial ability to

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 935


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

respond to and to enhance, internal and external signals, in The third mode constitutes a persistent, generally irrevers-
a manner analogous to the propagation, and Ca2⫹-acti- ible mPTP-associated large conductance of the inner mito-
vated reaction triggering of CICR (125, 127, 128, 139). chondrial membrane sufficient to activate the mechanisms
of destruction of mitochondria destined for elimination (for
We propose that the mPTP may be involved in at least four example, by mitophagy, because they do not meet the nec-
conceptual modes of pathophysiological function. The first essary mitochondrial “quality control” standards). Appar-
role is to serve as a potential release valve which transiently ently, the existence of impaired mitochondria could be a
opens a gate for controlled release of accumulated ions or threat for the proper functioning of the whole mitochon-
newly formed low-molecular-weight toxic substances in- drial population (because they are connected in a functional
cluding elevated ROS and Ca2⫹. This likely maintains mi- network), and opening of the mPTP may be interpreted as
tochondrial (and cellular) redox homeostasis, which is a the switch leading to programmed mitochondrial elimina-
critical point in normal cell functioning. The characteristics tion (mitophagy) (66, 76, 88, 210, 234, 258, 261, 446, 495,
of this mode include brief and reversible episodes of the 496). The signal initiating the process of mitochondrial ex-
mPTP pore opening or activation of some other mitochon- ecution may arise from the consistent and long-lasting drop
drial pores/channels/leaks constituting a low-conductance of the mitochondrial membrane potential (210) (FIGURE
state of the mPTP which have been demonstrated in in vitro 17), which may reinforce the transition to irreversible
systems (5, 29, 244, 255). We conceive that this serves a mPTP opening (i.e., the point of no return) (88, 188). The
housekeeping function, and at low energy demand (rest), it timing to reach this point is controlled in part by the deple-
is likely relatively inactive. At high stress/energy demand tion of the intramitochondrial redox buffer caused by the
conditions, this function plays a significant role (123). This flux of external (or internally produced) oxidants (491).
is borne out by the study (124) whereby the genetic loss of This point may correspond to the term mitochondrial crit-
CyPD (Ppif⫺/⫺) and acute cyclosporine A treatment (pre- icality (17) by reaching a redox crisis in an oxidatively
sumably causing mPTP-desensitization to ROS-induction) stressed cell after which the behavior of mitochondrial en-
result in elevated resting levels of Ca2⫹ in the mitochondrial ergetics becomes unstable. By this principle, criticality de-
termines ROS-induced organization of an excitable mito-
matrix of cardiomyocytes and an increase in the time of
chondrial cluster spreading oscillations of ⌬␺ over the cell.
decay to baseline following continuous pacing. Also,
Ppif⫺/⫺ mice display increased mortality and heart failure in
The fourth mode of the mPTP is the terminal destructive
pathological and physiological models of hypertrophy
step resulting in the elimination of the cell (which could be
(124) presumably due to an inappropriate high “set-point”
wanted or unwanted). As for mitochondria, impaired cells
for the physiological activation of transient mPTP-related
can be programmed for the destruction with the mPTP play-
RIRR events leading to the inappropriate mitochondrial
ing a critical role in that process. In general, this mechanism
accumulation of toxic levels of Ca2⫹ and ROS.
of cell elimination would encompass a large fraction of
mitochondria, undergoing the process described above, in-
The second mode may also include mPTP openings that
volving significant releases of ROS from mitochondria un-
prevent the accumulation of ROS and Ca2⫹ (etc.) inside dergoing mPTP (218).
mitochondria that would cause damage to that organelle
if not released into the cytosol where a large buffer pool Schematically, FIGURE 18 represents a hierarchical model of
might better deal with that stress factor. In this scenario, the proposed pathophysiological roles and functions of the
mPTP openings, although much more frequent, are still mPTP. According to this model, the stable state with the
brief enough not to cause irreversible mitochondrial or closed mPTP can undergo a transition to a metastable (tran-
cellular changes. These brief episodes of mPTP [or some sient) state of the mPTP. By analogy to an energy barrier
other channel(s)] opening may coincide with the flicker- (threshold) necessary to overcome in order to undergo a
ing mode of the mitochondrial membrane potential physico-chemical transition, the reaction probability and
which when integrated may cause a small albeit signifi- rate is determined by a redox threshold (process I) that
cant rise of ROS in the immediate vicinity of mitochon- normally is lower than the threshold of process II determin-
dria, thus activating local pools of redox-sensitive en- ing the transition to the fully open, stable mPTP. Such a
zymes, such as protein kinase C (218). In this mode, for metastable state of the pore (preceding its potential transi-
example, the levels of released ROS are enough to pre- tion to a long open state) occurs as a result of intra- and
condition the cell to upcoming more drastic oxidative extramitochondrial changes of the redox buffer still com-
challenge (e.g., long-lasting ischemic insult; Refs. 180, promising the dynamic balance between the influx/produc-
218, 335). The potential protective significance of the second tion of oxidants and antioxidants, and the stability of the
mode of the mPTP is to adaptively desensitize the mPTP pore “gating architecture” of the mPTP complex. As discussed
complex and to delay the transition to the third and fourth above, the RIRR-mPTP process may serve to provide an
destructive modes of mPTP associated with much higher levels adaptive release of excess Ca2⫹ and ROS from mitochon-
of oxidative stress. dria. The latter may be responsible for redox-activated pro-

936 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

ROS-Threshold
(“Energy Barrier”)

Protection
signaling
I. II.
Stable Transient- Stable
Closed state Open
mPTP mPTP mPTP

I. ΔΨm (repeated. . .)
Physiology
ROS

mPTP state

ROS
II. ΔΨm Burst

Pathology ROS

mPTP state

FIGURE 18. Two-ROS thresholds scheme for the induction of the mPTP. Resting state (shown by a pink ball)
is a state with a stable closed mPTP. After depletion of a redox buffer, the pore may undergo a transition to a
metastable flickering state (I) (shown by yellow color). The flickering mode of the pore is accompanied by a small
release of ROS from mitochondria, and when low it serves a physiological housekeeping purposes as a release
valve and when higher it is enough to precondition the system for better output after the following, more
extensive oxidative challenge (180). If the oxidative stress is severe, the phase I is changed by a state II (shown
by red) characterized by a stable open mPTP accompanied by a ROS burst. The pathological ROS burst results
in a programmed elimination of unwanted organelles or cells. The critical height of the ROS threshold is
determined by a balance between ROS production and quenching within mitochondria. It can be raised by
protective signaling [through preconditioning (218), cyclosporin A (180, 218), sanglofehrin (89), acidosis
(47), cyclophilin D acetylation or its knockout (26, 34, 313, 387)] shown by a green dotted line. Low redox
buffering (491), low SIRT3 activity (387), hexokinase II detachment from mitochondria (86), as well as high age
(216, 487) drive the threshold II down (shown by a brown dotted line). Note that one of the most important
elements of the scheme is that the long occupancy in state I hardens the onset of the state II.

tective signaling. The stable-open mPTP state is a result of pendent potassium channels, inhibition of glycogen syn-
the redox transition to the point of no return (long-lasting thase kinase-3␤, and other effectors; described in Ref. 218)
mPTP-opening) (188) leading to the destruction of mito- as well as more enhanced redox buffering (491), high
chondria (210), and when involving a sufficient fraction of deacetylase activity, specifically resulting in deacetylation
the entire mitochondrial population, it results in the death of CyPD (387), attachment of hexokinase II to the mPTP
of the cell. The height of both ROS thresholds (FIGURE 18) complex (86, 333) restore or even drive this threshold up,
is under redox control as well as the sensitivity of the mPTP- but this protective signaling also fails with aging (reviewed
gating mechanism (which is regulated by certain enzymatic in Ref. 216). Therefore, it is possible that the beneficial
posttranslational modifications), both of which may be tar- “anti-aging” effect of increased deacetylation (191) (par-
gets for pharmacological intervention. It has been shown ticularly by mitochondrial SIRT3–5; Refs. 43, 160, 429)
that the height of the second threshold is age dependent, of critical mPTP-regulatory elements such as CyPD (172)
being higher in mitochondria of young, and lower in mito- may partially be explained by a relative desensitization of
chondria of old animals (216, 257, 487). Various mPTP- the mPTP complex to ROS induction which could restore
related cytoprotective signaling cascades (e.g., through pre- the age-dependent decline of mPTP ROS threshold (i.e.,
conditioning, activation of PKC, receptor triggers including “aged” state) (257) back towards to the basal level
adenosine, opioid and ␤-receptors, mitochondrial ATP-de- (“young” state).

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 937


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

There is new information concerning the identification of tion of ROS which may activate the mPTP and/or IMAC
ATP synthase complex as the core structure necessary to and, if not appropriately terminated, is involved in the con-
form the mPTP that is certain to be relevant to the RIRR version of signaling to pathological ROS. The primary ROS
mechanism. The important role of mitochondrial ATP syn- signal to ignite RIRR may originate in as well as outside of
thase complex (complex V) in the mPTP functioning has mitochondria and may in turn be amplified by the RIRR
been proposed in a number of studies over more than two mechanism. The ROS amplification site may coincide with
decades (56, 87, 197, 322), in part based on the fact that one or more already known mitochondrial sites of ROS
oligomycin prevents mPTP in isolated mitochondria (321). production. The full understanding of the RIRR mechanism
However, this fact was explained by a probable indirect will require a more comprehensive and detailed mechanistic
effect of changed levels of adenine nucleotides rather than knowledge of the ROS-producing sites operating in situ and
by a direct effect on the pore structure itself. Recent evi- in vivo. It is certain that at least some of the mechanisms of
dence indicates that a dimer of complex V is necessary and mitochondrial ROS production presented in this review are
sufficient to form the mitochondrial pore with the apparent relevant to both physiological and pathological ROS pro-
modulating role of CyPD (157). It is tempting to speculate duction in the cell.
how this could potentially confer the gating mechanism of
the mPTP and explain bimodal behavior of ATP synthase in ACKNOWLEDGMENTS
terms of induction of the mPTP. Such bimodality (opened/
closed pore) may correlate with a bimodality of ATP syn- D. B. Zorov and S. J. Sollott contributed equally to this
thase (dimer/monomer and higher order oligomers). Quick paper.
and sometimes reversible transitions from closed to open
Address for reprint requests and other correspondence:
state of the pore (mPTP oscillations or flickers) may be
D. B. Zorov, A. N. Belozersky Institute of Physico-Chemi-
speculated to stem from a bimodal output of CyPD enzy-
cal Biology, Lomonosov Moscow State University, Mos-
matic activity which is a rotamase (peptidyl prolyl cis-trans
cow, Russia 119992 (e-mail: zorov@genebee.msu.su).
isomerase; Refs. 135, 431) determining two modes of con-
formation of proline residues in the targeted protein(s)
GRANTS
(134) (e.g., such as ATP synthase complex dimer), forming
the mitochondrial pore. Additionally, this may also explain
This work was supported by the Intramural Research Pro-
why CyPD acetylation critically appears to be affecting this
gram of the National Institutes of Health, National Institute
process, for example, during aging (387).
on Aging, Russian Foundation of Basic Research (14-04-
00542) and Russian Science Foundation (14-15-00147, 14-
In addition to the mPTP-associated RIRR, the important 24-00107). The funders had no role in data collection and
contribution of the IMAC-associated mode of RIRR func- analysis, decision to publish, or preparation of the manu-
tioning (20, 323), especially in the mitochondrial network script.
propagation properties, and its possible role in arrhythmo-
genesis, needs to be established. Resolving the molecular
identity of the IMAC will enable us to better appreciate its DISCLOSURES
role in physiology and pathology. It seems that IMAC-as-
sociated redox instability during pathological stress is quite No conflicts of interest, financial or otherwise, are declared
different from a safe, physiological flickering mode of the by the authors.
mPTP, and it is possible that IMAC-related mechanisms
may contribute to a transitional state between mPTP-asso-
ciated signaling and pathological modes of RIRR. REFERENCES

1. A man driven by proticity: Peter Mitchell Nobel prize for Chemistry 1978. Nature 276:
In conclusion, mPTP and IMAC are the main known play- 8 –9, 1978.
ers in the mechanism of RIRR which is a multifaceted pro-
2. Abramov AY, Scorziello A, Duchen MR. Three distinct mechanisms generate oxygen
cess important in the physiology and pathology of the cell. free radicals in neurons and contribute to cell death during anoxia and reoxygenation.
The mitochondrial role in this mechanism is above and J Neurosci 27: 1129 –1138, 2007.
beyond its well-recognized energy-producing activity, but 3. Adam-Vizi V, Chinopoulos C. Bioenergetics and the formation of mitochondrial re-
includes signaling potentially regulating both positive and active oxygen species. Trends Pharmacol Sci 27: 639 – 645, 2006.
negative destructive roles. Mitochondrially produced ROS
4. Ago T, Matsushima S, Kuroda J, Zablocki D, Kitazono T, Sadoshima J. The NADPH
are elements that may activate/deactivate a diverse set of oxidase Nox4 and aging in the heart. Aging 2: 1012–1016, 2010.
signals igniting as well as terminating signaling cascades.
5. Al-Nasser I, Crompton M. The reversible Ca2⫹-induced permeabilization of rat liver
Apparently, the “fine-tuning” regulation of ROS levels in mitochondria. Biochem J 239: 19 –29, 1986.
mitochondria is an essential function of RIRR, and mito-
6. Al Ghouleh I, Khoo NK, Knaus UG, Griendling KK, Touyz RM, Thannickal VJ, Bar-
chondria are important players in the propagation of ROS chowsky A, Nauseef WM, Kelley EE, Bauer PM, Darley-Usmar V, Shiva S, Cifuentes-
signals within a cell. Hypoxia/ischemia raises basal produc- Pagano E, Freeman BA, Gladwin MT, Pagano PJ. Oxidases and peroxidases in cardio-

938 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

vascular and lung disease: new concepts in reactive oxygen species signaling. Free runova VB, Serebryakova LI, Skulachev MV, Stelmashook EV, Studneva IM, Tskitishvili
Radic Biol Med 51: 1271–1288, 2011. OV, Vasilyeva AK, Victorov IV, Zorov DB, Skulachev VP. Mitochondria-targeted plas-
toquinone derivatives as tools to interrupt execution of the aging program. 2. Treat-
7. Altschuld RA, Hohl CM, Castillo LC, Garleb AA, Starling RC, Brierley GP. Cyclosporin ment of some ROS- and age-related diseases (heart arrhythmia, heart infarctions,
inhibits mitochondrial calcium efflux in isolated adult rat ventricular cardiomyocytes. kidney ischemia, and stroke). Biochemistry 73: 1288 –1299, 2008.
Am J Physiol Heart Circ Physiol 262: H1699 –H1704, 1992.
28. Bakeeva LE, Chentsov Yu S, Skulachev VP. Mitochondrial framework (reticulum mi-
8. Alvarez S, Valdez LB, Zaobornyj T, Boveris A. Oxygen dependence of mitochondrial tochondriale) in rat diaphragm muscle. Biochim Biophys Acta 501: 349 –369, 1978.
nitric oxide synthase activity. Biochem Biophys Res Commun 305: 771–775, 2003.
29. Balakirev MY, Zimmer G. Gradual changes in permeability of inner mitochondrial
9. Ambrosio G, Zweier JL, Duilio C, Kuppusamy P, Santoro G, Elia PP, Tritto I, Cirillo P, membrane precede the mitochondrial permeability transition. Arch Biochem Biophys
Condorelli M, Chiariello M. Evidence that mitochondrial respiration is a source of 356: 46 –54, 1998.
potentially toxic oxygen free radicals in intact rabbit hearts subjected to ischemia and
reflow. J Biol Chem 268: 18532–18541, 1993. 30. Bambrick LL, Chandrasekaran K, Mehrabian Z, Wright C, Krueger BK, Fiskum G.
Cyclosporin A increases mitochondrial calcium uptake capacity in cortical astrocytes
10. Amchenkova AA, Bakeeva LE, Chentsov YS, Skulachev VP, Zorov DB. Coupling but not cerebellar granule neurons. J Bioenerg Biomembr 38: 43– 47, 2006.
membranes as energy-transmitting cables. I. Filamentous mitochondria in fibroblasts
and mitochondrial clusters in cardiomyocytes. J Cell Biol 107: 481– 495, 1988. 31. Bao L, Avshalumov MV, Patel JC, Lee CR, Miller EW, Chang CJ, Rice ME. Mitochon-
dria are the source of hydrogen peroxide for dynamic brain-cell signaling. J Neurosci
11. Anderson RE, Meyer FB. In vivo fluorescent imaging of NADH redox state in brain. 29: 9002–9010, 2009.
Methods Enzymol 352: 482– 494, 2002.
32. Bardella C, Pollard PJ, Tomlinson I. SDH mutations in cancer. Biochim Biophys Acta
12. Andreyev A, Bondareva TO, Dedukhova VI, Mokhova EN, Skulachev VP, Tsofina LM, 1807: 1432–1443, 2011.
Volkov NI, Vygodina TV. The ATP/ADP-antiporter is involved in the uncoupling effect
of fatty acids on mitochondria. Eur J Biochem 182: 585–592, 1989. 33. Barlow CH, Chance B. Ischemic areas in perfused rat hearts: measurement by NADH
fluorescence photography. Science 193: 909 –910, 1976.
13. Andreyev AY, Kushnareva YE, Starkov AA. Mitochondrial metabolism of reactive
oxygen species. Biochemistry 70: 200 –214, 2005. 34. Basso E, Fante L, Fowlkes J, Petronilli V, Forte MA, Bernardi P. Properties of the
permeability transition pore in mitochondria devoid of cyclophilin D. J Biol Chem 280:
14. Anisimov VN, Egorov MV, Krasilshchikova MS, Lyamzaev KG, Manskikh VN, Moshkin 18558 –18561, 2005.
MP, Novikov EA, Popovich IG, Rogovin KA, Shabalina IG, Shekarova ON, Skulachev
MV, Titova TV, Vygodin VA, Vyssokikh MY, Yurova MN, Zabezhinsky MA, Skulachev 35. Batandier C, Leverve X, Fontaine E. Opening of the mitochondrial permeability tran-
VP. Effects of the mitochondria-targeted antioxidant SkQ1 on lifespan of rodents. sition pore induces reactive oxygen species production at the level of the respiratory
Aging 3: 1110 –1119, 2011. chain complex I. J Biol Chem 279: 17197–17204, 2004.

15. Antonenko YN, Avetisyan AV, Cherepanov DA, Knorre DA, Korshunova GA, 36. Baughman JM, Perocchi F, Girgis HS, Plovanich M, Belcher-Timme CA, Sancak Y, Bao
Markova OV, Ojovan SM, Perevoshchikova IV, Pustovidko AV, Rokitskaya TI, Sev- XR, Strittmatter L, Goldberger O, Bogorad RL, Koteliansky V, Mootha VK. Integrative
erina II, Simonyan RA, Smirnova EA, Sobko AA, Sumbatyan NV, Severin FF, Skulachev genomics identifies MCU as an essential component of the mitochondrial calcium
VP. Derivatives of rhodamine 19 as mild mitochondria-targeted cationic uncouplers. uniporter. Nature 476: 341–345, 2012.
J Biol Chem 286: 17831–17840, 2011.
37. Baysal BE, Willett-Brozick JE, Lawrence EC, Drovdlic CM, Savul SA, McLeod DR, Yee
16. Aon MA, Cortassa S, Akar FG, Brown DA, Zhou L, O’Rourke B. From mitochondrial HA, Brackmann DE, Slattery WH, 3rd Myers EN, Ferrell RE, Rubinstein WS. Preva-
dynamics to arrhythmias. Int J Biochem Cell Biol 41: 1940 –1948, 2009. lence of SDHB, SDHC, and SDHD germline mutations in clinic patients with head and
neck paragangliomas. J Med Genet 39: 178 –183, 2002.
17. Aon MA, Cortassa S, Akar FG, O’Rourke B. Mitochondrial criticality: a new concept at
the turning point of life or death. Biochim Biophys Acta 1762: 232–240, 2006. 38. Beavis AD. On the inhibition of the mitochondrial inner membrane anion uniporter by
cationic amphiphiles and other drugs. J Biol Chem 264: 1508 –1515, 1989.
18. Aon MA, Cortassa S, Marban E, O’Rourke B. Synchronized whole cell oscillations in
mitochondrial metabolism triggered by a local release of reactive oxygen species in 39. Beavis AD. Properties of the inner membrane anion channel in intact mitochondria. J
cardiac myocytes. J Biol Chem 278: 44735– 44744, 2003. Bioenerg Biomembr 24: 77–90, 1992.

19. Aon MA, Cortassa S, O’Rourke B. The fundamental organization of cardiac mitochon- 40. Beavis AD, Garlid KD. The mitochondrial inner membrane anion channel. Regulation
dria as a network of coupled oscillators. Biophys J 91: 4317– 4327, 2006. by divalent cations and protons. J Biol Chem 262: 15085–15093, 1987.

20. Aon MA, Cortassa S, O’Rourke B. Mitochondrial oscillations in physiology and patho- 41. Beckmann JD, Frerman FE. Reaction of electron-transfer flavoprotein with electron-
physiology. Adv Exp Med Biol 641: 98 –117, 2008. transfer flavoprotein-ubiquinone oxidoreductase. Biochemistry 24: 3922–3925, 1985.

21. Aon MA, Stanley BA, Sivakumaran V, Kembro JM, O’Rourke B, Paolocci N, Cortassa 42. Beinert H, Kennedy MC, Stout CD. Aconitase as iron ⫺ sulfur protein, enzyme, and
S. Glutathione/thioredoxin systems modulate mitochondrial H2O2 emission: an ex- iron-regulatory protein. Chem Rev 96: 2335–2374, 1996.
perimental-computational study. J Gen Physiol 139: 479 – 491, 2012.
43. Bellizzi D, Rose G, Cavalcante P, Covello G, Dato S, De Rango F, Greco V, Maggiolini
22. Appaix F, Kuznetsov AV, Usson Y, Kay L, Andrienko T, Olivares J, Kaambre T, Sikk P, M, Feraco E, Mari V, Franceschi C, Passarino G, De Benedictis G. A novel VNTR
Margreiter R, Saks V. Possible role of cytoskeleton in intracellular arrangement and enhancer within the SIRT3 gene, a human homologue of SIR2, is associated with
regulation of mitochondria. Exp Physiol 88: 175–190, 2003. survival at oldest ages. Genomics 85: 258 –263, 2005.

23. Azzi A, Azzone GF. Swelling and shrinkage phenomena in liver mitochondria. II. Low 44. Benzi G, Pastoris O, Dossena M. Relationships between gamma-aminobutyrate and
amplitude swelling-shrinkage cycles. Biochim Biophys Acta 105: 265–278, 1965. succinate cycles during and after cerebral ischemia. J Neurosci Res 7: 193–201, 1982.

24. Babior BM. NADPH oxidase: an update. Blood 93: 1464 –1476, 1999. 45. Berden JA, Slater EC. The allosteric binding of antimycin to cytochrome b in the
mitochondrial membrane. Biochim Biophys Acta 256: 199 –215, 1972.
25. Bae YS, Oh H, Rhee SG, Yoo YD. Regulation of reactive oxygen species generation in
cell signaling. Mol Cells 32: 491–509, 2011. 46. Bereiter-Hahn J, Voth M. Dynamics of mitochondria in living cells: shape changes,
dislocations, fusion, and fission of mitochondria. Microsc Res Tech 27: 198 –219, 1994.
26. Baines CP, Kaiser RA, Purcell NH, Blair NS, Osinska H, Hambleton MA, Brunskill EW,
Sayen MR, Gottlieb RA, Dorn GW, Robbins J, Molkentin JD. Loss of cyclophilin D 47. Bernardi P. Mitochondrial transport of cations: channels, exchangers, and permeabil-
reveals a critical role for mitochondrial permeability transition in cell death. Nature ity transition. Physiol Rev 79: 1127–1155, 1999.
434: 658 – 662, 2005.
48. Berniakovich I, Trinei M, Stendardo M, Migliaccio E, Minucci S, Bernardi P, Pelicci PG,
27. Bakeeva LE, Barskov IV, Egorov MV, Isaev NK, Kapelko VI, Kazachenko AV, Kirpa- Giorgio M. p66Shc-generated oxidative signal promotes fat accumulation. J Biol Chem
tovsky VI, Kozlovsky SV, Lakomkin VL, Levina SB, Pisarenko OI, Plotnikov EY, Sap- 283: 34283–34293, 2008.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 939


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

49. Berry C, La Vecchia C, Nicotera P. Paraquat and Parkinson’s disease. Cell Death Differ 72. Burgoyne JR, Mongue-Din H, Eaton P, Shah AM. Redox signaling in cardiac physiology
17: 1115–1125, 2010. and pathology. Circ Res 111: 1091–1106, 2012.

50. Bianchi P, Kunduzova O, Masini E, Cambon C, Bani D, Raimondi L, Seguelas MH, 73. Cadenas S, Buckingham JA, Samec S, Seydoux J, Din N, Dulloo AG, Brand MD. UCP2
Nistri S, Colucci W, Leducq N, Parini A. Oxidative stress by monoamine oxidase and UCP3 rise in starved rat skeletal muscle but mitochondrial proton conductance is
mediates receptor-independent cardiomyocyte apoptosis by serotonin and postisch- unchanged. FEBS Lett 462: 257–260, 1999.
emic myocardial injury. Circulation 112: 3297–3305, 2005.
74. Caldeira da Silva CC, Cerqueira FM, Barbosa LF, Medeiros MH, Kowaltowski AJ. Mild
51. Bianchi P, Pimentel DR, Murphy MP, Colucci WS, Parini A. A new hypertrophic mitochondrial uncoupling in mice affects energy metabolism, redox balance and lon-
mechanism of serotonin in cardiac myocytes: receptor-independent ROS generation. gevity. Aging Cell 7: 552–560, 2008.
FASEB J 19: 641– 643, 2005.
75. Carafoli E, Tiozzo R, Lugli G, Crovetti F, Kratzing C. The release of calcium from heart
52. Bianchi P, Seguelas MH, Parini A, Cambon C. Activation of pro-apoptotic cascade by mitochondria by sodium. J Mol Cell Cardiol 6: 361–371, 1974.
dopamine in renal epithelial cells is fully dependent on hydrogen peroxide generation
by monoamine oxidases. J Am Soc Nephrol 14: 855– 862, 2003. 76. Carreira RS, Lee Y, Ghochani M, Gustafsson AB, Gottlieb RA. Cyclophilin D is re-
quired for mitochondrial removal by autophagy in cardiac cells. Autophagy 6: 462–
53. Biary N, Xie C, Kauffman J, Akar FG. Biophysical properties and functional conse- 472, 2010.
quences of reactive oxygen species (ROS)-induced ROS release in intact myocardium.
J Physiol 589: 5167–5179, 2011. 77. Chance B. The energy-linked reaction of calcium with mitochondria. J Biol Chem 240:
2729 –2748, 1965.
54. Bindokas VP, Kuznetsov A, Sreenan S, Polonsky KS, Roe MW, Philipson LH. Visualiz-
ing superoxide production in normal and diabetic rat islets of Langerhans. J Biol Chem 78. Chance B, Cohen P, Jobsis F, Schoener B. Intracellular oxidation-reduction states in
278: 9796 –9801, 2003. vivo. Science 137: 499 –508, 1962.

55. Block K, Gorin Y, Abboud HE. Subcellular localization of Nox4 and regulation in 79. Chance B, Hollunger G. The interaction of energy and electron transfer reactions in
diabetes. Proc Natl Acad Sci USA 106: 14385–14390, 2009. mitochondria. I. General properties and nature of the products of succinate-linked
reduction of pyridine nucleotide. J Biol Chem 236: 1534 –1543, 1961.
56. Bonora M, Bononi A, De Marchi E, Giorgi C, Lebiedzinska M, Marchi S, Patergnani S,
Rimessi A, Suski JM, Wojtala A, Wieckowski MR, Kroemer G, Galluzzi L, Pinton P. 80. Chance B, Hollunger G. The interaction of energy and electron transfer reactions in
Role of the c subunit of the FO ATP synthase in mitochondrial permeability transition. mitochondria. IV. The pathway of electron transfer. J Biol Chem 236: 1562–1568,
Cell Cycle 12: 674 – 683, 2013. 1961.

57. Boveris A, Chance B. The mitochondrial generation of hydrogen peroxide. General 81. Chance B, Sies H, Boveris A. Hydroperoxide metabolism in mammalian organs.
properties and effect of hyperbaric oxygen. Biochem J 134: 707–716, 1973. Physiol Rev 59: 527– 605, 1979.

58. Boveris A, Oshino N, Chance B. The cellular production of hydrogen peroxide. 82. Chance B, Williams GR. Respiratory enzymes in oxidative phosphorylation. III. The
Biochem J 128: 617– 630, 1972. steady state. J Biol Chem 217: 409 – 427, 1955.

59. Braissant O. Current concepts in the pathogenesis of urea cycle disorders. Mol Genet 83. Chance B, Yoshioka T. Sustained oscillations of ionic constituents of mitochondria.
Metab 100 Suppl 1: S3–S12, 2010. Arch Biochem Biophys 117: 451– 465, 1966.

60. Brand MD. Electroneutral efflux of Ca2⫹ from liver mitochondria. Biochem J 225: 84. Chaudhari T, McGuire W. Allopurinol for preventing mortality and morbidity in new-
413– 419, 1985. born infants with suspected hypoxic-ischaemic encephalopathy. Cochrane Database
Syst Rev CD006817, 2008.
61. Brand MD. The stoichiometry of the exchange catalysed by the mitochondrial
calcium/sodium antiporter. Biochem J 229: 161–166, 1985. 85. Chen CH, Greenawalt JW, Lehninger AL. Biochemical and ultrastructural aspects of
Ca2⫹ transport by mitochondria of the hepatopancreas of the blue crab Callinectes
62. Brandt U. Energy converting NADH:quinone oxidoreductase (complex I). Annu Rev sapidus. J Cell Biol 61: 301–315, 1974.
Biochem 75: 69 –92, 2006.
86. Chiara F, Castellaro D, Marin O, Petronilli V, Brusilow WS, Juhaszova M, Sollott SJ,
63. Braunwald E, Kloner RA. Myocardial reperfusion: a double-edged sword? J Clin Invest Forte M, Bernardi P, Rasola A. Hexokinase II detachment from mitochondria triggers
76: 1713–1719, 1985. apoptosis through the permeability transition pore independent of voltage-dependent
anion channels. PLoS One 3: e1852, 2008.
64. Brdiczka DG, Zorov DB, Sheu SS. Mitochondrial contact sites: their role in energy
metabolism and apoptosis. Biochim Biophys Acta 1762: 148 –163, 2006. 87. Chinopoulos C, Adam-Vizi V. Modulation of the mitochondrial permeability transition
by cyclophilin D: moving closer to F(0)-F(1) ATP synthase? Mitochondrion 12: 41– 45,
65. Broekemeier KM, Dempsey ME, Pfeiffer DR. Cyclosporin A is a potent inhibitor of the 2013.
inner membrane permeability transition in liver mitochondria. J Biol Chem 264: 7826 –
7830, 1989. 88. Ciechanover A. The ubiquitin-proteasome proteolytic pathway. Cell 79: 13–21, 1994.

66. Broekemeier KM, Iben JR, LeVan EG, Crouser ED, Pfeiffer DR. Pore formation and 89. Clarke SJ, McStay GP, Halestrap AP. Sanglifehrin A acts as a potent inhibitor of the
uncoupling initiate a Ca2⫹-independent degradation of mitochondrial phospholipids. mitochondrial permeability transition and reperfusion injury of the heart by binding to
Biochemistry 41: 7771–7780, 2002. cyclophilin-D at a different site from cyclosporin A. J Biol Chem 277: 34793–34799,
2002.
67. Broekemeier KM, Klocek CK, Pfeiffer DR. Proton selective substate of the mitochon-
drial permeability transition pore: regulation by the redox state of the electron trans- 90. Colman E. Dinitrophenol and obesity: an early twentieth-century regulatory dilemma.
port chain. Biochemistry 37: 13059 –13065, 1998. Regul Toxicol Pharmacol 48: 115–117, 2007.

68. Brouillet E, Conde F, Beal MF, Hantraye P. Replicating Huntington’s disease pheno- 91. Cooper CE, Davies NA. Effects of nitric oxide and peroxynitrite on the cytochrome
type in experimental animals. Prog Neurobiol 59: 427– 468, 1999. oxidase K(m) for oxygen: implications for mitochondrial pathology. Biochim Biophys
Acta 1459: 390 –396, 2000.
69. Brown GC, Borutaite V. Nitric oxide and mitochondrial respiration in the heart.
Cardiovasc Res 75: 283–290, 2007. 92. Corkey BE, Tornheim K, Deeney JT, Glennon MC, Parker JC, Matschinsky FM,
Ruderman NB, Prentki M. Linked oscillations of free Ca2⫹ and the ATP/ADP ratio in
70. Brown GC, Cooper CE. Nanomolar concentrations of nitric oxide reversibly inhibit permeabilized RINm5F insulinoma cells supplemented with a glycolyzing cell-free
synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS muscle extract. J Biol Chem 263: 4254 – 4258, 1988.
Lett 356: 295–298, 1994.
93. Cortassa S, Aon MA, Marban E, Winslow RL, O’Rourke B. An integrated model of
71. Bunik VI, Sievers C. Inactivation of the 2-oxo acid dehydrogenase complexes upon cardiac mitochondrial energy metabolism and calcium dynamics. Biophys J 84: 2734 –
generation of intrinsic radical species. Eur J Biochem 269: 5004 –5015, 2002. 2755, 2003.

940 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

94. Cortopassi G, Wang E. Modelling the effects of age-related mtDNA mutation accu- 114. Dougherty TJ, Gomer CJ, Henderson BW, Jori G, Kessel D, Korbelik M, Moan J, Peng
mulation; complex I deficiency, superoxide and cell death. Biochim Biophys Acta 1271: Q. Photodynamic therapy. J Natl Cancer Inst 90: 889 –905, 1998.
171–176, 1995.
115. Drachev VA, Zorov DB. [Mitochondria as an electric cable. Experimental testing of a
95. Crofts AR, Barquera B, Gennis RB, Kuras R, Guergova-Kuras M, Berry EA. Mechanism hypothesis]. Dokl Akad Nauk SSSR 287: 1237–1238, 1986.
of ubiquinol oxidation by the bc(1) complex: different domains of the quinol binding
pocket and their role in the mechanism and binding of inhibitors. Biochemistry 38: 116. Drahota Z, Chowdhury SK, Floryk D, Mracek T, Wilhelm J, Rauchova H, Lenaz G,
15807–15826, 1999. Houstek J. Glycerophosphate-dependent hydrogen peroxide production by brown
adipose tissue mitochondria and its activation by ferricyanide. J Bioenerg Biomembr 34:
96. Crompton M, Andreeva L. On the involvement of a mitochondrial pore in reperfusion 105–113, 2002.
injury. Basic Res Cardiol 88: 513–523, 1993.
117. Droge W. Free radicals in the physiological control of cell function. Physiol Rev 82:
97. Crompton M, Costi A. A heart mitochondrial Ca2⫹-dependent pore of possible rel- 47–95, 2002.
evance to re-perfusion-induced injury. Evidence that ADP facilitates pore intercon-
version between the closed and open states. Biochem J 266: 33–39, 1990. 118. Duchen MR, McGuinness O, Brown LA, Crompton M. On the involvement of a
cyclosporin A sensitive mitochondrial pore in myocardial reperfusion injury. Cardio-
98. Crompton M, Kunzi M, Carafoli E. The calcium-induced and sodium-induced effluxes vasc Res 27: 1790 –1794, 1993.
of calcium from heart mitochondria. Evidence for a sodium-calcium carrier. Eur J
Biochem 79: 549 –558, 1977. 119. Duvert M, Verna A. Ultrastructure and architecture of the sarcoplasmic reticulum in
frog sino-atrial fibres: a comparative study with various preparatory procedures. J Mol
99. Cunha FM, Caldeira da Silva CC, Cerqueira FM, Kowaltowski AJ. Mild mitochondrial Cell Cardiol 17: 43–56, 1985.
uncoupling as a therapeutic strategy. Curr Drug Targets 12: 783–789, 2011.
120. Echtay KS, Roussel D, St-Pierre J, Jekabsons MB, Cadenas S, Stuart JA, Harper JA,
100. Dai DF, Rabinovitch PS, Ungvari Z. Mitochondria and cardiovascular aging. Circ Res Roebuck SJ, Morrison A, Pickering S, Clapham JC, Brand MD. Superoxide activates
110: 1109 –1124, 2012. mitochondrial uncoupling proteins. Nature 415: 96 –99, 2002.

101. Das DK, George A, Liu XK, Rao PS. Detection of hydroxyl radical in the mitochondria 121. Ekstedt B. Substrate specificity of the different forms of monoamine oxidase in rat
of ischemic-reperfused myocardium by trapping with salicylate. Biochem Biophys Res liver mitochondria. Biochem Pharmacol 25: 1133–1138, 1976.
Commun 165: 1004 –1009, 1989.
122. Eleff S, Kennaway NG, Buist NR, Darley-Usmar VM, Capaldi RA, Bank WJ, Chance B.
102. De Bie MK, Buiten MS, Rabelink TJ, Jukema JW. How to reduce sudden cardiac death 31
P NMR study of improvement in oxidative phosphorylation by vitamins K3 and C in
in patients with renal failure. Heart 98: 335–341, 2012. a patient with a defect in electron transport at complex III in skeletal muscle. Proc Natl
Acad Sci USA 81: 3529 –3533, 1984.
103. De Cavanagh EM, Fraga CG, Ferder L, Inserra F. Enalapril and captopril enhance
antioxidant defenses in mouse tissues. Am J Physiol Regul Integr Comp Physiol 272: 123. Elrod JW, Molkentin JD. Physiologic functions of cyclophilin D and the mitochondrial
R514 –R518, 1997. permeability transition pore. Circ J 77: 1111–1122, 2013.

104. De Jong AM, Kotlyar AB, Albracht SP. Energy-induced structural changes in NADH:Q 124. Elrod JW, Wong R, Mishra S, Vagnozzi RJ, Sakthievel B, Goonasekera SA, Karch J,
oxidoreductase of the mitochondrial respiratory chain. Biochim Biophys Acta 1186: Gabel S, Farber J, Force T, Brown JH, Murphy E, Molkentin JD. Cyclophilin D controls
163–171, 1994. mitochondrial pore-dependent Ca2⫹ exchange, metabolic flexibility, and propensity
for heart failure in mice. J Clin Invest 120: 3680 –3687, 2010.
105. De Stefani D, Raffaello A, Teardo E, Szabo I, Rizzuto R. A forty-kilodalton protein of
the inner membrane is the mitochondrial calcium uniporter. Nature 476: 336 –340, 125. Endo M. Calcium release from the sarcoplasmic reticulum. Physiol Rev 57: 71–108,
2012. 1977.

106. Deluca HF, Engstrom GW. Calcium uptake by rat kidney mitochondria. Proc Natl Acad 126. Evtodienko Yu V, Teplova V, Khawaja J, Saris NE. The Ca2⫹-induced permeability
Sci USA 47: 1744 –1750, 1961. transition pore is involved in Ca2⫹-induced mitochondrial oscillations. A study on
permeabilised Ehrlich ascites tumour cells. Cell Calcium 15: 143–152, 1994.
107. Denton RM, McCormack JG, Edgell NJ. Role of calcium ions in the regulation of
intramitochondrial metabolism. Effects of Na⫹, Mg2⫹ and ruthenium red on the 127. Fabiato A, Fabiato F. Calcium-induced release of calcium from the sarcoplasmic re-
Ca2⫹-stimulated oxidation of oxoglutarate and on pyruvate dehydrogenase activity in ticulum of skinned cells from adult human, dog, cat, rabbit, rat, and frog hearts and
intact rat heart mitochondria. Biochem J 190: 107–117, 1980. from fetal and new-born rat ventricles. Ann NY Acad Sci 307: 491–522, 1978.

108. Di Lisa F, Carpi A, Giorgio V, Bernardi P. The mitochondrial permeability transition 128. Fabiato A, Fabiato F. Calcium release from the sarcoplasmic reticulum. Circ Res 40:
pore and cyclophilin D in cardioprotection. Biochim Biophys Acta 1813: 1316 –1322, 119 –129, 1977.
2011.
129. Feng Y, Shi W, Huang M, LeBlanc MH. Oxypurinol administration fails to prevent
109. Di Lisa F, Kaludercic N, Carpi A, Menabo R, Giorgio M. Mitochondrial pathways for hypoxic-ischemic brain injury in neonatal rats. Brain Res Bull 59: 453– 457, 2003.
ROS formation and myocardial injury: the relevance of p66(Shc) and monoamine
oxidase. Basic Res Cardiol 104: 131–139, 2009. 130. Fenton HJH. The oxidation of tartaric acid in presence of iron. J Chem Soc Proc 9: 113,
1893.
110. Di Lisa F, Silverman HS, Hansford RG. Mitochondrial function and cell injury in single
cardiac myocytes exposed to anoxia and reoxygenation. Transplant Proc 27: 2829 – 131. Ferguson SJ. Fully delocalised chemiosmotic or Iocalised proton flow pathways in
2830, 1995. energy coupling? A scrutiny of experimental evidence. Biochim Biophys Acta 811:
47–95, 1985.
111. Dikalov S. Cross talk between mitochondria and NADPH oxidases. Free Radic Biol
Med 51: 1289 –1301, 2011. 132. Fernandez-Gomez FJ, Galindo MF, Gomez-Lazaro M, Gonzalez-Garcia C, Cena V,
Aguirre N, Jordan J. Involvement of mitochondrial potential and calcium buffering
112. Dileepan KN, Kennedy J. Complete inhibition of dihydro-orotate oxidation and su- capacity in minocycline cytoprotective actions. Neuroscience 133: 959 –967, 2005.
peroxide production by 1,1,1-trifluoro-3-thenoylacetone in rat liver mitochondria.
Biochem J 225: 189 –194, 1985. 133. Finetti F, Pellegrini M, Ulivieri C, Savino MT, Paccagnini E, Ginanneschi C, Lanfran-
cone L, Pelicci PG, Baldari CT. The proapoptotic and antimitogenic protein p66SHC
113. Dong LF, Jameson VJ, Tilly D, Cerny J, Mahdavian E, Marin-Hernandez A, Hernandez- acts as a negative regulator of lymphocyte activation and autoimmunity. Blood 111:
Esquivel L, Rodriguez-Enriquez S, Stursa J, Witting PK, Stantic B, Rohlena J, Truksa J, 5017–5027, 2008.
Kluckova K, Dyason JC, Ledvina M, Salvatore BA, Moreno-Sanchez R, Coster MJ,
Ralph SJ, Smith RA, Neuzil J. Mitochondrial targeting of vitamin E succinate enhances 134. Fischer G, Schmid FX. The mechanism of protein folding. Implications of in vitro
its pro-apoptotic and anti-cancer activity via mitochondrial complex II. J Biol Chem refolding models for de novo protein folding and translocation in the cell. Biochemistry
286: 3717–3728, 2011. 29: 2205–2212, 1990.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 941


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

135. Fischer G, Wittmann-Liebold B, Lang K, Kiefhaber T, Schmid FX. Cyclophilin and 156. Giorgio M, Migliaccio E, Orsini F, Paolucci D, Moroni M, Contursi C, Pelliccia G, Luzi
peptidyl-prolyl cis-trans isomerase are probably identical proteins. Nature 337: 476 – L, Minucci S, Marcaccio M, Pinton P, Rizzuto R, Bernardi P, Paolucci F, Pelicci PG.
478, 1989. Electron transfer between cytochrome c and p66Shc generates reactive oxygen spe-
cies that trigger mitochondrial apoptosis. Cell 122: 221–233, 2005.
136. Fiskum G, Cockrell RS. Ruthenium red sensitive and insensitive calcium transport in
rat liver and Ehrlich ascites tumor cell mitochondria. FEBS Lett 92: 125–128, 1978. 157. Giorgio V, von Stockum S, Antoniel M, Fabbro A, Fogolari F, Forte M, Glick GD,
Petronilli V, Zoratti M, Szabo I, Lippe G, Bernardi P. Dimers of mitochondrial ATP
137. Fiskum G, Lehninger AL. Regulated release of Ca2⫹ from respiring mitochondria by synthase form the permeability transition pore. Proc Natl Acad Sci USA 110: 5887–
Ca2⫹/2H⫹ antiport. J Biol Chem 254: 6236 – 6239, 1979. 5892, 2013.
138. Flint DH, Tuminello JF, Emptage MH. The inactivation of Fe-S cluster containing 158. Gooch VD, Packer L. Adenine nucleotide control of heart mitochondrial oscillations.
hydro-lyases by superoxide. J Biol Chem 268: 22369 –22376, 1993. Biochim Biophys Acta 245: 17–20, 1971.
139. Ford LE, Podolsky RJ. Regenerative calcium release within muscle cells. Science 167: 159. Graham NA, Tahmasian M, Kohli B, Komisopoulou E, Zhu M, Vivanco I, Teitell MA,
58 –59, 1970. Wu H, Ribas A, Lo RS, Mellinghoff IK, Mischel PS, Graeber TG. Glucose deprivation
activates a metabolic and signaling amplification loop leading to cell death. Mol Syst Biol
140. Forman HJ, Kennedy J. Superoxide production and electron transport in mitochon-
8: 589, 2012.
drial oxidation of dihydroorotic acid. J Biol Chem 250: 4322– 4326, 1975.
160. Green MF, Hirschey MD. SIRT3 weighs heavily in the metabolic balance: a new role
141. Frerman FE, Goodman SI. Deficiency of electron transfer flavoprotein or electron
for SIRT3 in metabolic syndrome. J Gerontol A Biol Sci Med Sci 68: 105–107, 2013.
transfer flavoprotein:ubiquinone oxidoreductase in glutaric acidemia type II fibro-
blasts. Proc Natl Acad Sci USA 82: 4517– 4520, 1985. 161. Griffiths EJ, Halestrap AP. Protection by Cyclosporin A of ischemia/reperfusion-in-
duced damage in isolated rat hearts. J Mol Cell Cardiol 25: 1461–1469, 1993.
142. Fridovich I. Superoxide dismutases. Adv Enzymol Relat Areas Mol Biol 41: 35–97, 1974.
162. Grimm S. Respiratory chain complex II as general sensor for apoptosis. Biochim Bio-
143. Fukai T, Ushio-Fukai M. Superoxide dismutases: role in redox signaling, vascular
phys Acta 1827: 565–572, 2013.
function, and diseases. Antioxid Redox Signal 15: 1583–1606, 2011.
163. Grivennikova VG, Cecchini G, Vinogradov AD. Ammonium-dependent hydrogen
144. Galloway JH, Cartwright IJ, Bennett MJ. Abnormal myocardial lipid composition in an
peroxide production by mitochondria. FEBS Lett 582: 2719 –2724, 2008.
infant with type II glutaric aciduria. J Lipid Res 28: 279 –284, 1987.
164. Grivennikova VG, Kareyeva AV, Vinogradov AD. What are the sources of hydrogen
145. Galluzzi L, Vitale I, Abrams JM, Alnemri ES, Baehrecke EH, Blagosklonny MV, Dawson
peroxide production by heart mitochondria? Biochim Biophys Acta 1797: 939 –944,
TM, Dawson VL, El-Deiry WS, Fulda S, Gottlieb E, Green DR, Hengartner MO, Kepp
2010.
O, Knight RA, Kumar S, Lipton SA, Lu X, Madeo F, Malorni W, Mehlen P, Nunez G,
Peter ME, Piacentini M, Rubinsztein DC, Shi Y, Simon HU, Vandenabeele P, White E, 165. Grivennikova VG, Vinogradov AD. Generation of superoxide by the mitochondrial
Yuan J, Zhivotovsky B, Melino G, Kroemer G. Molecular definitions of cell death
Complex I. Biochim Biophys Acta 1757: 553–561, 2006.
subroutines: recommendations of the Nomenclature Committee on Cell Death
2012. Cell Death Differ 19: 107–120, 2012. 166. Gunter TE, Pfeiffer DR. Mechanisms by which mitochondria transport calcium. Am J
Physiol Cell Physiol 258: C755–C786, 1990.
146. Gane EJ, Weilert F, Orr DW, Keogh GF, Gibson M, Lockhart MM, Frampton CM,
Taylor KM, Smith RA, Murphy MP. The mitochondria-targeted anti-oxidant mitoqui- 167. Guo J, Lemire BD. The ubiquinone-binding site of the Saccharomyces cerevisiae succi-
none decreases liver damage in a phase II study of hepatitis C patients. Liver Int 30: nate-ubiquinone oxidoreductase is a source of superoxide. J Biol Chem 278: 47629 –
1019 –1026, 2010. 47635, 2003.

147. Gardner PR, Fridovich I. Inactivation-reactivation of aconitase in Escherichia coli. A 168. Gylkhandanyan AV, Evtodienko YV, Zhabotinsky AM, Kondrashova MN. Continuous
sensitive measure of superoxide radical. J Biol Chem 267: 8757– 8763, 1992. Sr2⫹-induced oscillations of the ionic fluxes in mitochondria. FEBS Lett 66: 44 – 47,
1976.
148. Gardner PR, Nguyen DD, White CW. Aconitase is a sensitive and critical target of
oxygen poisoning in cultured mammalian cells and in rat lungs. Proc Natl Acad Sci USA 169. Haber F, Weiss J. über die Katalyse des Hydroperoxydes. Naturwissenshaften 20:
91: 12248 –12252, 1994. 948 –950, 1932.

149. Garlick PB, Davies MJ, Hearse DJ, Slater TF. Direct detection of free radicals in the 170. Halestrap AP. A pore way to die: the role of mitochondria in reperfusion injury and
reperfused rat heart using electron spin resonance spectroscopy. Circ Res 61: 757– cardioprotection. Biochem Soc Trans 38: 841– 860, 2010.
760, 1987.
171. Halestrap AP. The regulation of the oxidation of fatty acids and other substrates in rat
150. Gavrikova EV, Vinogradov AD. Active/de-active state transition of the mitochondrial heart mitochondria by changes in the matrix volume induced by osmotic strength,
complex I as revealed by specific sulfhydryl group labeling. FEBS Lett 455: 36 – 40, valinomycin and Ca2⫹. Biochem J 244: 159 –164, 1987.
1999.
172. Halestrap AP, Davidson AM. Inhibition of Ca2⫹-induced large-amplitude swelling of
151. Gazaryan IG, Krasnikov BF, Ashby GA, Thorneley RN, Kristal BS, Brown AM. Zinc is liver and heart mitochondria by cyclosporin is probably caused by the inhibitor binding
a potent inhibitor of thiol oxidoreductase activity and stimulates reactive oxygen to mitochondrial-matrix peptidyl-prolyl cis-trans isomerase and preventing it inter-
species production by lipoamide dehydrogenase. J Biol Chem 277: 10064 –10072, acting with the adenine nucleotide translocase. Biochem J 268: 153–160, 1990.
2002.
173. Hall AM, Crawford C, Unwin RJ, Duchen MR, Peppiatt-Wildman CM. Multiphoton
152. Genova ML, Ventura B, Giuliano G, Bovina C, Formiggini G, Parenti Castelli G, Lenaz imaging of the functioning kidney. J Am Soc Nephrol 22: 1297–1304, 2011.
G. The site of production of superoxide radical in mitochondrial Complex I is not a
bound ubisemiquinone but presumably iron-sulfur cluster N2. FEBS Lett 505: 364 – 174. Hansen JM, Go YM, Jones DP. Nuclear and mitochondrial compartmentation of
368, 2001. oxidative stress and redox signaling. Annu Rev Pharmacol Toxicol 46: 215–234, 2006.

153. Ghisla S, Thorpe C. Acyl-CoA dehydrogenases. A mechanistic overview. Eur J 175. Hansford RG, Hogue BA, Mildaziene V. Dependence of H2O2 formation by rat heart
Biochem 271: 494 –508, 2004. mitochondria on substrate availability and donor age. J Bioenerg Biomembr 29: 89 –95,
1997.
154. Gilbert HF. Molecular and cellular aspects of thiol-disulfide exchange. Adv Enzymol
Relat Areas Mol Biol 63: 69 –172, 1990. 176. Hansford RG, Zorov D. Role of mitochondrial calcium transport in the control of
substrate oxidation. Mol Cell Biochem 184: 359 –369, 1998.
155. Gimm O, Armanios M, Dziema H, Neumann HP, Eng C. Somatic and occult germ-line
mutations in SDHD, a mitochondrial complex II gene, in nonfamilial pheochromocy- 177. Harper ME, Ballantyne JS, Leach M, Brand MD. Effects of thyroid hormones on
toma. Cancer Res 60: 6822– 6825, 2000. oxidative phosphorylation. Biochem Soc Trans 21: 785–792, 1993.

942 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

178. Hassinen I, Chance B. Oxidation-reduction properties of the mitochondrial flavopro- 201. Inoue M, Sato EF, Nishikawa M, Park AM, Kira Y, Imada I, Utsumi K. Cross talk of
tein chain. Biochem Biophys Res Commun 31: 895–900, 1968. nitric oxide, oxygen radicals, and superoxide dismutase regulates the energy metab-
olism and cell death and determines the fates of aerobic life. Antioxid Redox Signal 5:
179. Hauptmann N, Grimsby J, Shih JC, Cadenas E. The metabolism of tyramine by mono- 475– 484, 2003.
amine oxidase A/B causes oxidative damage to mitochondrial DNA. Arch Biochem
Biophys 335: 295–304, 1996. 202. Inoue M, Sato EF, Nishikawa M, Park AM, Kira Y, Imada I, Utsumi K. Mitochondrial
generation of reactive oxygen species and its role in aerobic life. Curr Med Chem 10:
180. Hausenloy D, Wynne A, Duchen M, Yellon D. Transient mitochondrial permeability 2495–2505, 2003.
transition pore opening mediates preconditioning-induced protection. Circulation
109: 1714 –1717, 2004. 203. Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, Fearon ER, Sundaresan M, Finkel T,
Goldschmidt-Clermont PJ. Mitogenic signaling mediated by oxidants in Ras-trans-
181. Hausenloy DJ, Yellon DM. The therapeutic potential of ischemic conditioning: an formed fibroblasts. Science 275: 1649 –1652, 1997.
update. Nat Rev Cardiol 8: 619 – 629, 2011.
204. Ishii T, Yasuda K, Akatsuka A, Hino O, Hartman PS, Ishii N. A mutation in the SDHC
182. Haworth RA, Hunter DR. The Ca2⫹-induced membrane transition in mitochondria. gene of complex II increases oxidative stress, resulting in apoptosis and tumorigenesis.
II. Nature of the Ca2⫹ trigger site. Arch Biochem Biophys 195: 460 – 467, 1979. Cancer Res 65: 203–209, 2005.
183. Hearse DJ, Garlick PB, Humphrey SM. Ischemic contracture of the myocardium: 205. Ivanes F, Mewton N, Rioufol G, Piot C, Elbaz M, Revel D, Croisille P, Ovize M.
mechanisms and prevention. Am J Cardiol 39: 986 –993, 1977. Cardioprotection in the clinical setting. Cardiovasc Drugs Ther 24: 281–287, 2010.
184. Henderson PJ, Lardy HA. Bongkrekic acid. An inhibitor of the adenine nucleotide 206. Jankauskas SS, Plotnikov EY, Morosanova MA, Pevzner IB, Zorova LD, Skulachev VP,
translocase of mitochondria. J Biol Chem 245: 1319 –1326, 1970.
Zorov DB. Mitochondria-targeted antioxidant SkQR1 ameliorates gentamycin-in-
185. Hernandez-Garcia D, Wood CD, Castro-Obregon S, Covarrubias L. Reactive oxygen duced renal failure and hearing loss. Biochemistry 77: 666 – 670, 2012.
species: a radical role in development? Free Radic Biol Med 49: 130 –143, 2010.
207. Jennings RB. Commentary on selected aspects of cardioprotection. J Cardiovasc Phar-
186. Herrero A, Barja G. Localization of the site of oxygen radical generation inside the macol Ther 16: 340 –348, 2011.
complex I of heart and nonsynaptic brain mammalian mitochondria. J Bioenerg
208. Jennings RB, Sommers HM, Smyth GA, Flack HA, Linn H. Myocardial necrosis induced
Biomembr 32: 609 – 615, 2000.
by temporary occlusion of a coronary artery in the dog. Arch Pathol 70: 68 –78, 1960.
187. Hinkle PC, Butow RA, Racker E, Chance B. Partial resolution of the enzymes catalyz-
209. Jensen PK. Antimycin-insensitive oxidation of succinate and reduced nicotinamide-
ing oxidative phosphorylation. XV. Reverse electron transfer in the flavin-cytochrome
adenine dinucleotide in electron-transport particles. II. Steroid effects. Biochim Bio-
beta region of the respiratory chain of beef heart submitochondrial particles. J Biol
phys Acta 122: 167–174, 1966.
Chem 242: 5169 –5173, 1967.
210. Jin SM, Lazarou M, Wang C, Kane LA, Narendra DP, Youle RJ. Mitochondrial mem-
188. Hirsch T, Susin SA, Marzo I, Marchetti P, Zamzami N, Kroemer G. Mitochondrial
brane potential regulates PINK1 import and proteolytic destabilization by PARL. J Cell
permeability transition in apoptosis and necrosis. Cell Biol Toxicol 14: 141–145, 1998.
Biol 191: 933–942, 2010.
189. Hoffman DL, Salter JD, Brookes PS. Response of mitochondrial reactive oxygen
211. Johnson LV, Walsh ML, Chen LB. Localization of mitochondria in living cells with
species generation to steady-state oxygen tension: implications for hypoxic cell sig-
rhodamine 123. Proc Natl Acad Sci USA 77: 990 –994, 1980.
naling. Am J Physiol Heart Circ Physiol 292: H101–H108, 2007.
212. Joiner ML, Koval OM, Li J, He BJ, Allamargot C, Gao Z, Luczak ED, Hall DD, Fink BD,
190. Holmuhamedov EL, Teplova VV, Chukhlova EA, Evtodienko YV, Ulrich RG. Stron-
Chen B, Yang J, Moore SA, Scholz TD, Strack S, Mohler PJ, Sivitz WI, Song LS,
tium excitability of the inner mitochondrial membrane: regenerative strontium-in-
Anderson ME. CaMKII determines mitochondrial stress responses in heart. Nature
duced strontium release. Biochem Mol Biol Int 36: 39 – 49, 1995.
491: 269 –273, 2012.
191. Howitz KT, Bitterman KJ, Cohen HY, Lamming DW, Lavu S, Wood JG, Zipkin RE,
213. Jolly SR, Kane WJ, Bailie MB, Abrams GD, Lucchesi BR. Canine myocardial reperfusion
Chung P, Kisielewski A, Zhang LL, Scherer B, Sinclair DA. Small molecule activators of
injury. Its reduction by the combined administration of superoxide dismutase and
sirtuins extend Saccharomyces cerevisiae lifespan. Nature 425: 191–196, 2003.
catalase. Circ Res 54: 277–285, 1984.
192. Huang J, Lemire BD. Mutations in the C. elegans succinate dehydrogenase iron-sulfur
214. Jones DP. Intracellular diffusion gradients of O2 and ATP. Am J Physiol Cell Physiol 250:
subunit promote superoxide generation and premature aging. J Mol Biol 387: 559 –
C663–C675, 1986.
569, 2009.
215. Jouaville LS, Ichas F, Holmuhamedov EL, Camacho P, Lechleiter JD. Synchronization
193. Huang LS, Cobessi D, Tung EY, Berry EA. Binding of the respiratory chain inhibitor
of calcium waves by mitochondrial substrates in Xenopus laevis oocytes. Nature 377:
antimycin to the mitochondrial bc1 complex: a new crystal structure reveals an al-
438 – 441, 1995.
tered intramolecular hydrogen-bonding pattern. J Mol Biol 351: 573–597, 2005.
216. Juhaszova M, Rabuel C, Zorov DB, Lakatta EG, Sollott SJ. Protection in the aged heart:
194. Huang X, Zhai D, Huang Y. Study on the relationship between calcium-induced
calcium release from mitochondria and PTP opening. Mol Cell Biochem 213: 29 –35, preventing the heart-break of old age? Cardiovasc Res 66: 233–244, 2005.
2000.
217. Juhaszova M, Wang S, Zorov DB, Nuss HB, Gleichmann M, Mattson MP, Sollott SJ.
195. Hunter DR, Haworth RA. The Ca2⫹-induced membrane transition in mitochondria. I. The identity and regulation of the mitochondrial permeability transition pore: where
The protective mechanisms. Arch Biochem Biophys 195: 453– 459, 1979. the known meets the unknown. Ann NY Acad Sci 1123: 197–212, 2008.

196. Hunter DR, Haworth RA. The Ca2⫹-induced membrane transition in mitochondria. 218. Juhaszova M, Zorov DB, Kim SH, Pepe S, Fu Q, Fishbein KW, Ziman BD, Wang S,
III. Transitional Ca2⫹ release. Arch Biochem Biophys 195: 468 – 477, 1979. Ytrehus K, Antos CL, Olson EN, Sollott SJ. Glycogen synthase kinase-3beta mediates
convergence of protection signaling to inhibit the mitochondrial permeability transi-
197. Hunter DR, Haworth RA, Southard JH. Relationship between configuration, function, tion pore. J Clin Invest 113: 1535–1549, 2004.
and permeability in calcium-treated mitochondria. J Biol Chem 251: 5069 –5077, 1976.
219. Juhaszova M, Zorov DB, Yaniv Y, Nuss HB, Wang S, Sollott SJ. Role of glycogen
198. Huser J, Rechenmacher CE, Blatter LA. Imaging the permeability pore transition in synthase kinase-3beta in cardioprotection. Circ Res 104: 1240 –1252, 2009.
single mitochondria. Biophys J 74: 2129 –2137, 1998.
220. Kakar S, Hoffman FG, Storz JF, Fabian M, Hargrove MS. Structure and reactivity of
199. Ichas F, Jouaville LS, Mazat JP. Mitochondria are excitable organelles capable of gen- hexacoordinate hemoglobins. Biophys Chem 152: 1–14, 2010.
erating and conveying electrical and calcium signals. Cell 89: 1145–1153, 1997.
221. Kaludercic N, Carpi A, Menabo R, Di Lisa F, Paolocci N. Monoamine oxidases (MAO)
200. Imlay JA. A metabolic enzyme that rapidly produces superoxide, fumarate reductase in the pathogenesis of heart failure and ischemia/reperfusion injury. Biochim Biophys
of Escherichia coli. J Biol Chem 270: 19767–19777, 1995. Acta 1813: 1323–1332, 2011.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 943


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

222. Kaminski MM, Sauer SW, Kaminski M, Opp S, Ruppert T, Grigaravicius P, Grudnik P, 242. Kozlov AV, Bahrami S, Calzia E, Dungel P, Gille L, Kuznetsov AV, Troppmair J.
Grone HJ, Krammer PH, Gulow K. T cell activation is driven by an ADP-dependent Mitochondrial dysfunction and biogenesis: do ICU patients die from mitochondrial
glucokinase linking enhanced glycolysis with mitochondrial reactive oxygen species failure? Ann Intensive Care 1: 41, 2011.
generation. Cell Rep 2: 1300 –1315, 2012.
243. Krasnikov BF, Avad AS, Zorov DB, Yaguzhinsky LS. Effects of amyl ester of unsubsti-
223. Kapay NA, Popova OV, Isaev NK, Stelmashook EV, Kondratenko RV, Zorov DB, tuted rhodamine on respiration and Ca2⫹ transport in rat liver mitochondria. Biochem
Skrebitsky VG, Skulachev VP. Mitochondria-targeted plastoquinone antioxidant SkQ1 Biophys Res Commun 175: 1010 –1016, 1991.
prevents amyloid-beta-induced impairment of long-term potentiation in rat hip-
pocampal slices. J Alzheimers Dis 36: 377–383, 2013. 244. Krasnikov BF, Zorov DB, Antonenko YN, Zaspa AA, Kulikov IV, Kristal BS, Cooper AJ,
Brown AM. Comparative kinetic analysis reveals that inducer-specific ion release
224. Kareyeva AV, Grivennikova VG, Cecchini G, Vinogradov AD. Molecular identification precedes the mitochondrial permeability transition. Biochim Biophys Acta 1708: 375–
of the enzyme responsible for the mitochondrial NADH-supported ammonium-de- 392, 2005.
pendent hydrogen peroxide production. FEBS Lett 585: 385–389, 2011.
245. Krishnamoorthy G, Hinkle PC. Studies on the electron transfer pathway, topography
225. Kelso GF, Porteous CM, Coulter CV, Hughes G, Porteous WK, Ledgerwood EC, of iron-sulfur centers, and site of coupling in NADH-Q oxidoreductase. J Biol Chem
Smith RA, Murphy MP. Selective targeting of a redox-active ubiquinone to mitochon- 263: 17566 –17575, 1988.
dria within cells: antioxidant and antiapoptotic properties. J Biol Chem 276: 4588 –
246. Krogh A. The supply of oxygen to the tissues and the regulation of the capillary
4596, 2001.
circulation. J Physiol 52: 457– 474, 1919.
226. Khanday FA, Santhanam L, Kasuno K, Yamamori T, Naqvi A, Dericco J, Bugayenko A,
247. Kudin AP, Bimpong-Buta NY, Vielhaber S, Elger CE, Kunz WS. Characterization of
Mattagajasingh I, Disanza A, Scita G, Irani K. Sos-mediated activation of rac1 by
superoxide-producing sites in isolated brain mitochondria. J Biol Chem 279: 4127–
p66shc. J Cell Biol 172: 817– 822, 2006.
4135, 2004.
227. Kim JJ, Miura R. Acyl-CoA dehydrogenases and acyl-CoA oxidases. Structural basis for
248. Kummel L. Ca,Mg-ATPase activity of permeabilised rat heart cells and its functional
mechanistic similarities and differences. Eur J Biochem 271: 483– 493, 2004.
coupling to oxidative phosphorylation of the cells. Cardiovasc Res 22: 359 –367, 1988.
228. Kinnally KW, Campo ML, Tedeschi H. Mitochondrial channel activity studied by
249. Kunduzova OR, Bianchi P, Parini A, Cambon C. Hydrogen peroxide production by
patch-clamping mitoplasts. J Bioenerg Biomembr 21: 497–506, 1989.
monoamine oxidase during ischemia/reperfusion. Eur J Pharmacol 448: 225–230,
229. Kinnally KW, Zorov D, Antonenko Y, Perini S. Calcium modulation of mitochondrial 2002.
inner membrane channel activity. Biochem Biophys Res Commun 176: 1183–1188,
250. Kunz WS, Kunz W. Contribution of different enzymes to flavoprotein fluorescence of
1991.
isolated rat liver mitochondria. Biochim Biophys Acta 841: 237–246, 1985.
230. Kinnally KW, Zorov DB, Antonenko YN, Snyder SH, McEnery MW, Tedeschi H.
251. Kuroda J, Ago T, Matsushima S, Zhai P, Schneider MD, Sadoshima J. NADPH oxidase
Mitochondrial benzodiazepine receptor linked to inner membrane ion channels by
4 (Nox4) is a major source of oxidative stress in the failing heart. Proc Natl Acad Sci USA
nanomolar actions of ligands. Proc Natl Acad Sci USA 90: 1374 –1378, 1993.
107: 15565–15570, 2010.
231. Kirichok Y, Krapivinsky G, Clapham DE. The mitochondrial calcium uniporter is a 252. Kurz FT, Aon MA, O’Rourke B, Armoundas AA. Spatio-temporal oscillations of indi-
highly selective ion channel. Nature 427: 360 –364, 2004. vidual mitochondria in cardiac myocytes reveal modulation of synchronized mito-
chondrial clusters. Proc Natl Acad Sci USA 107: 14315–14320, 2010.
232. Klann E, Thiels E. Modulation of protein kinases and protein phosphatases by reactive
oxygen species: implications for hippocampal synaptic plasticity. Prog Neuropsychop- 253. Kurz FT, Aon MA, O’Rourke B, Armoundas AA. Wavelet analysis reveals heteroge-
harmacol Biol Psychiatry 23: 359 –376, 1999. neous time-dependent oscillations of individual mitochondria. Am J Physiol Heart Circ
Physiol 299: H1736 –H1740, 2010.
233. Klingenberg M, Grebe K, Scherer B. Opposite effects of bongkrekic acid and atrac-
tyloside on the adenine nucleotides induced mitochondrial volume changes and on the 254. Kushnareva Y, Murphy AN, Andreyev A. Complex I-mediated reactive oxygen spe-
efflux of adenine nucleotides. FEBS Lett 16: 253–256, 1971. cies generation: modulation by cytochrome c and NAD(P)⫹ oxidation-reduction
state. Biochem J 368: 545–553, 2002.
234. Klionsky DJ, Emr SD. Autophagy as a regulated pathway of cellular degradation.
Science 290: 1717–1721, 2000. 255. Kushnareva YE, Sokolove PM. Prooxidants open both the mitochondrial permeability
transition pore and a low-conductance channel in the inner mitochondrial membrane.
235. Kokoszka JE, Waymire KG, Levy SE, Sligh JE, Cai J, Jones DP, MacGregor GR, Wallace
Arch Biochem Biophys 376: 377–388, 2000.
DC. The ADP/ATP translocator is not essential for the mitochondrial permeability
transition pore. Nature 427: 461– 465, 2004. 256. Kussmaul L, Hirst J. The mechanism of superoxide production by NADH:ubiquinone
oxidoreductase (complex I) from bovine heart mitochondria. Proc Natl Acad Sci USA
236. Kolosova NG, Stefanova NA, Muraleva NA, Skulachev VP. The mitochondria-tar-
103: 7607–7612, 2006.
geted antioxidant SkQ1 but not N-acetylcysteine reverses aging-related biomarkers in
rats. Aging 4: 686 – 694, 2012. 257. Lakatta EG, Sollott SJ. The “heartbreak” of older age. Mol Interv 2: 431– 446, 2002.

237. Koppenol WH. The centennial of the Fenton reaction. Free Radic Biol Med 15: 645– 258. Lee J, Giordano S, Zhang J. Autophagy, mitochondria and oxidative stress: cross-talk
651, 1993. and redox signalling. Biochem J 441: 523–540, 2012.

238. Korde AS, Pettigrew LC, Craddock SD, Maragos WF. The mitochondrial uncoupler 259. Lehninger AL, Carafoli E, Rossi CS. Energy-linked ion movements in mitochondrial
2,4-dinitrophenol attenuates tissue damage and improves mitochondrial homeostasis systems. Adv Enzymol Relat Areas Mol Biol 29: 259 –320, 1967.
following transient focal cerebral ischemia. J Neurochem 94: 1676 –1684, 2005.
260. Lemarie A, Huc L, Pazarentzos E, Mahul-Mellier AL, Grimm S. Specific disintegration
239. Korge P, Yang L, Yang JH, Wang Y, Qu Z, Weiss JN. Protective role of transient pore of complex II succinate:ubiquinone oxidoreductase links pH changes to oxidative
openings in calcium handling by cardiac mitochondria. J Biol Chem 286: 34851–34857, stress for apoptosis induction. Cell Death Differ 18: 338 –349, 2011.
2011.
261. Lemasters JJ, Nieminen AL, Qian T, Trost LC, Elmore SP, Nishimura Y, Crowe RA,
240. Korshunov SS, Skulachev VP, Starkov AA. High protonic potential actuates a mecha- Cascio WE, Bradham CA, Brenner DA, Herman B. The mitochondrial permeability
nism of production of reactive oxygen species in mitochondria. FEBS Lett 416: 15–18, transition in cell death: a common mechanism in necrosis, apoptosis and autophagy.
1997. Biochim Biophys Acta 1366: 177–196, 1998.

241. Kotlyar AB, Sled VD, Burbaev DS, Moroz IA, Vinogradov AD. Coupling site I and the 262. Lenartowicz E, Bernardi P, Azzone GF. Phenylarsine oxide induces the cyclosporin
rotenone-sensitive ubisemiquinone in tightly coupled submitochondrial particles. A-sensitive membrane permeability transition in rat liver mitochondria. J Bioenerg
FEBS Lett 264: 17–20, 1990. Biomembr 23: 679 – 688, 1991.

944 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

263. Lenaz G. The mitochondrial production of reactive oxygen species: mechanisms and 284. Malkevitch NV, Dedukhova VI, Simonian RA, Skulachev VP, Starkov AA. Thyroxine
implications in human pathology. IUBMB Life 52: 159 –164, 2001. induces cyclosporin A-insensitive, Ca2⫹-dependent reversible permeability transition
pore in rat liver mitochondria. FEBS Lett 412: 173–178, 1997.
264. Lenaz G. Role of mitochondria in oxidative stress and ageing. Biochim Biophys Acta
1366: 53– 67, 1998. 285. Marklund SL. Extracellular superoxide dismutase and other superoxide dismutase
isoenzymes in tissues from nine mammalian species. Biochem J 222: 649 – 655, 1984.
265. Lenaz G, Genova ML. Supramolecular organisation of the mitochondrial respiratory
chain: a new challenge for the mechanism and control of oxidative phosphorylation. 286. Marshall C, Mamary AJ, Verhoeven AJ, Marshall BE. Pulmonary artery NADPH-
Adv Exp Med Biol 748: 107–144, 2012. oxidase is activated in hypoxic pulmonary vasoconstriction. Am J Respir Cell Mol Biol
15: 633– 644, 1996.
266. Levitt DG. Capillary-tissue exchange kinetics: an analysis of the Krogh cylinder model.
J Theor Biol 34: 103–124, 1972. 287. Massey V, Strickland S, Mayhew SG, Howell LG, Engel PC, Matthews RG, Schuman M,
Sullivan PA. The production of superoxide anion radicals in the reaction of reduced
267. Liang LP, Patel M. Iron-sulfur enzyme mediated mitochondrial superoxide toxicity in flavins and flavoproteins with molecular oxygen. Biochem Biophys Res Commun 36:
experimental Parkinson’s disease. J Neurochem 90: 1076 –1084, 2004. 891– 897, 1969.

268. Liberman EA, Topaly VP, Tsofina LM, Jasaitis AA, Skulachev VP. Mechanism of cou- 288. Maurel A, Hernandez C, Kunduzova O, Bompart G, Cambon C, Parini A, Frances B.
pling of oxidative phosphorylation and the membrane potential of mitochondria. Age-dependent increase in hydrogen peroxide production by cardiac monoamine
Nature 222: 1076 –1078, 1969. oxidase A in rats. Am J Physiol Heart Circ Physiol 284: H1460 –H1467, 2003.

269. Lim S, Rashid MA, Jang M, Kim Y, Won H, Lee J, Woo JT, Kim YS, Murphy MP, Ali L, 289. McGuire BJ, Secomb TW. A theoretical model for oxygen transport in skeletal muscle
Ha J, Kim SS. Mitochondria-targeted antioxidants protect pancreatic beta-cells against under conditions of high oxygen demand. J Appl Physiol 91: 2255–2265, 2001.
oxidative stress and improve insulin secretion in glucotoxicity and glucolipotoxicity.
Cell Physiol Biochem 28: 873– 886, 2011. 290. Meng Q, Wong YT, Chen J, Ruan R. Age-related changes in mitochondrial function and
antioxidative enzyme activity in fischer 344 rats. Mech Ageing Dev 128: 286 –292,
270. Liot G, Bossy B, Lubitz S, Kushnareva Y, Sejbuk N, Bossy-Wetzel E. Complex II 2007.
inhibition by 3-NP causes mitochondrial fragmentation and neuronal cell death via an
291. Menini S, Iacobini C, Ricci C, Oddi G, Pesce C, Pugliese F, Block K, Abboud HE,
NMDA- and ROS-dependent pathway. Cell Death Differ 16: 899 –909, 2009.
Giorgio M, Migliaccio E, Pelicci PG, Pugliese G. Ablation of the gene encoding p66Shc
271. Liu R, Liu W, Doctrow SR, Baudry M. Iron toxicity in organotypic cultures of hip- protects mice against AGE-induced glomerulopathy by preventing oxidant-depen-
pocampal slices: role of reactive oxygen species. J Neurochem 85: 492–502, 2003. dent tissue injury and further AGE accumulation. Diabetologia 50: 1997–2007, 2007.

272. Liu SS. Cooperation of a ”reactive oxygen cycle“ with the Q cycle and the proton cycle 292. Mercer JR, Yu E, Figg N, Cheng KK, Prime TA, Griffin JL, Masoodi M, Vidal-Puig A,
in the respiratory chain–superoxide generating and cycling mechanisms in mitochon- Murphy MP, Bennett MR. The mitochondria-targeted antioxidant MitoQ decreases
dria. J Bioenerg Biomembr 31: 367–376, 1999. features of the metabolic syndrome in ATM⫹/⫺/ApoE⫺/⫺ mice. Free Radic Biol Med
52: 841– 849, 2012.
273. Liu SS. Generating, partitioning, targeting and functioning of superoxide in mitochon-
dria. Biosci Rep 17: 259 –272, 1997. 293. Messner KR, Imlay JA. The identification of primary sites of superoxide and hydrogen
peroxide formation in the aerobic respiratory chain and sulfite reductase complex of
274. Liu SS. Mitochondrial Q cycle-derived superoxide and chemiosmotic bioenergetics. Escherichia coli. J Biol Chem 274: 10119 –10128, 1999.
Ann NY Acad Sci 1201: 84 –95, 2010.
294. Messner KR, Imlay JA. Mechanism of superoxide and hydrogen peroxide formation by
275. Liu SS, Huang JP. Co-existence of ”reactive oxygen species“ with Q cycle and proton fumarate reductase, succinate dehydrogenase, and aspartate oxidase. J Biol Chem 277:
cycle in respiratory chain fo mitochondria. In: Proceedings of the International Sympo- 42563– 42571, 2002.
sium on Natural Antioxidants: Molecular Mechanisms and Health Effects, edited by
295. Migliaccio E, Giorgio M, Mele S, Pelicci G, Reboldi P, Pandolfi PP, Lanfrancone L,
Parker L, Traber MG, Xin WJ. Champaign, IL: AOCS, 1996, p. 511–529.
Pelicci PG. The p66shc adaptor protein controls oxidative stress response and life
276. Loffler M, Becker C, Wegerle E, Schuster G. Catalytic enzyme histochemistry and span in mammals. Nature 402: 309 –313, 1999.
biochemical analysis of dihydroorotate dehydrogenase/oxidase and succinate dehy-
296. Minotti G, Recalcati S, Menna P, Salvatorelli E, Corna G, Cairo G. Doxorubicin car-
drogenase in mammalian tissues, cells and mitochondria. Histochem Cell Biol 105:
diotoxicity and the control of iron metabolism: quinone-dependent and independent
119 –128, 1996.
mechanisms. Methods Enzymol 378: 340 –361, 2004.
277. Lotscher HR, Winterhalter KH, Carafoli E, Richter C. Hydroperoxide-induced loss of
297. Mironov SL, Richter DW. Oscillations and hypoxic changes of mitochondrial variables
pyridine nucleotides and release of calcium from rat liver mitochondria. J Biol Chem
in neurons of the brainstem respiratory centre of mice. J Physiol 533: 227–236, 2001.
255: 9325–9330, 1980.
298. Mitchell P. Chemiosmotic coupling in oxidative and photosynthetic phosphorylation.
278. Lukacs GL, Kapus A, Fonyo A. Parallel measurement of oxoglutarate dehydrogenase
Biol Rev Camb Philos Soc 41: 445–502, 1966.
activity and matrix free Ca2⫹ in fura-2-loaded heart mitochondria. FEBS Lett 229:
219 –223, 1988. 299. Mitchell P. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-
osmotic type of mechanism. Nature 191: 144 –148, 1961.
279. Maalouf RM, Eid AA, Gorin YC, Block K, Escobar GP, Bailey S, Abboud HE. Nox4-
derived reactive oxygen species mediate cardiomyocyte injury in early type 1 diabe- 300. Miyata H, Silverman HS, Sollott SJ, Lakatta EG, Stern MD, Hansford RG. Measure-
tes. Am J Physiol Cell Physiol 302: C597–C604, 2012. ment of mitochondrial free Ca2⫹ concentration in living single rat cardiac myocytes.
Am J Physiol Heart Circ Physiol 261: H1123–H1134, 1991.
280. Maejima Y, Kuroda J, Matsushima S, Ago T, Sadoshima J. Regulation of myocardial
growth and death by NADPH oxidase. J Mol Cell Cardiol 50: 408 – 416, 2011. 301. Modica-Napolitano JS, Brunelli BT, Koya K, Chen LB. Photoactivation enhances the
mitochondrial toxicity of the cationic rhodacyanine MKT-077. Cancer Res 58: 71–75,
281. Magnus G, Keizer J. Model of beta-cell mitochondrial calcium handling and electrical 1998.
activity. II. Mitochondrial variables. Am J Physiol Cell Physiol 274: C1174 –C1184, 1998.
302. Moreno-Sanchez R, Hansford RG. Dependence of cardiac mitochondrial pyruvate
282. Maker HS, Weiss C, Silides DJ, Cohen G. Coupling of dopamine oxidation (mono- dehydrogenase activity on intramitochondrial free Ca2⫹ concentration. Biochem J
amine oxidase activity) to glutathione oxidation via the generation of hydrogen per- 256: 403– 412, 1988.
oxide in rat brain homogenates. J Neurochem 36: 589 –593, 1981.
303. Moreno G, Poussin K, Ricchelli F, Salet C. The effects of singlet oxygen produced by
283. Maklashina E, Sher Y, Zhou HZ, Gray MO, Karliner JS, Cecchini G. Effect of anoxia/ photodynamic action on the mitochondrial permeability transition differ in accor-
reperfusion on the reversible active/de-active transition of NADH-ubiquinone oxi- dance with the localization of the sensitizer. Arch Biochem Biophys 386: 243–250,
doreductase (complex I) in rat heart. Biochim Biophys Acta 1556: 6 –12, 2002. 2001.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 945


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

304. Mracek T, Pecinova A, Vrbacky M, Drahota Z, Houstek J. High efficiency of ROS 327. Orlowski S, Nowak W. Topology and thermodynamics of gaseous ligands diffusion
production by glycerophosphate dehydrogenase in mammalian mitochondria. Arch paths in human neuroglobin. Biosystems 94: 263–266, 2008.
Biochem Biophys 481: 30 –36, 2009.
328. Orsini F, Migliaccio E, Moroni M, Contursi C, Raker VA, Piccini D, Martin-Padura
305. Muller U. Pathological mechanisms and parent-of-origin effects in hereditary paragan- I, Pelliccia G, Trinei M, Bono M, Puri C, Tacchetti C, Ferrini M, Mannucci R,
glioma/pheochromocytoma (PGL/PCC). Neurogenetics 12: 175–181, 2011. Nicoletti I, Lanfrancone L, Giorgio M, Pelicci PG. The life span determinant p66Shc
localizes to mitochondria where it associates with mitochondrial heat shock protein
306. Murphy AN, Bredesen DE, Cortopassi G, Wang E, Fiskum G. Bcl-2 potentiates the 70 and regulates trans-membrane potential. J Biol Chem 279: 25689 –25695, 2004.
maximal calcium uptake capacity of neural cell mitochondria. Proc Natl Acad Sci USA
93: 9893–9898, 1996. 329. Otani H, Tanaka H, Inoue T, Umemoto M, Omoto K, Tanaka K, Sato T, Osako T,
Masuda A, Nonoyama A. In vitro study on contribution of oxidative metabolism of
307. Murphy MP. Development of lipophilic cations as therapies for disorders due to isolated rabbit heart mitochondria to myocardial reperfusion injury. Circ Res 55: 168 –
mitochondrial dysfunction. Expert Opin Biol Ther 1: 753–764, 2001. 175, 1984.

308. Murphy MP. How mitochondria produce reactive oxygen species. Biochem J 417: 330. Packer L, Utsumi R, Mustafa MG. Oscillatory states of mitochondria. 1. Electron and
1–13, 2009. energy transfer pathways. Arch Biochem Biophys 117: 381–393, 1966.

309. Murphy MP. Mitochondrial thiols in antioxidant protection and redox signaling: dis- 331. Palmer JW, Pfeiffer DR. The control of Ca2⫹ release from heart mitochondria. J Biol
tinct roles for glutathionylation and other thiol modifications. Antioxid Redox Signal 16: Chem 256: 6742– 6750, 1981.
476 – 495, 2012.
332. Palty R, Silverman WF, Hershfinkel M, Caporale T, Sensi SL, Parnis J, Nolte C, Fishman
310. Murphy MP, Siegel RM. Mitochondrial ROS fire up T cell activation. Immunity 38: D, Shoshan-Barmatz V, Herrmann S, Khananshvili D, Sekler I. NCLX is an essential
201–202, 2013. component of mitochondrial Na⫹/Ca2⫹ exchange. Proc Natl Acad Sci USA 107: 436 –
441, 2012.
311. Murry CE, Jennings RB, Reimer KA. Preconditioning with ischemia: a delay of lethal
cell injury in ischemic myocardium. Circulation 74: 1124 –1136, 1986. 333. Pastorino JG, Hoek JB. Regulation of hexokinase binding to VDAC. J Bioenerg
Biomembr 40: 171–182, 2008.
312. Mustafa MG, Utsumi K, Packer L. Damped oscillatory control of mitochondrial res-
piration and volume. Biochem Biophys Res Commun 24: 381–385, 1966. 334. Patole MS, Swaroop A, Ramasarma T. Generation of H2O2 in brain mitochondria. J
Neurochem 47: 1– 8, 1986.
313. Nakagawa T, Shimizu S, Watanabe T, Yamaguchi O, Otsu K, Yamagata H, Inohara H,
Kubo T, Tsujimoto Y. Cyclophilin D-dependent mitochondrial permeability transition 335. Penna C, Perrelli MG, Pagliaro P. Mitochondrial pathways, permeability transition
regulates some necrotic but not apoptotic cell death. Nature 434: 652– 658, 2005. pore, and redox signaling in cardioprotection: therapeutic implications. Antioxid Redox
Signal 18: 556 –599, 2013.
314. Napankangas JP, Liimatta EV, Joensuu P, Bergmann U, Ylitalo K, Hassinen IE. Super-
oxide production during ischemia-reperfusion in the perfused rat heart: a comparison 336. Perry TL, Godin DV, Hansen S. Parkinson’s disease: a disorder due to nigral glutathi-
of two methods of measurement. J Mol Cell Cardiol 53: 906 –915, 2012. one deficiency? Neurosci Lett 33: 305–310, 1982.

315. Naviaux RK. Oxidative shielding or oxidative stress? J Pharmacol Exp Ther 342: 608 – 337. Petlicki J, van de Ven TGM. The equilibrium between the oxidation of hydrogen
618, 2012. peroxide by oxygen and the dismutation of peroxyl or superoxide radicals in aqeous
solutions in contact with oxygen. J Chem Soc Faraday Trans 94: 2763–2767, 1998.
316. Negre-Salvayre A, Hirtz C, Carrera G, Cazenave R, Troly M, Salvayre R, Penicaud L,
Casteilla L. A role for uncoupling protein-2 as a regulator of mitochondrial hydrogen 338. Pfanner N, Neupert W. The mitochondrial protein import apparatus. Annu Rev
peroxide generation. FASEB J 11: 809 – 815, 1997. Biochem 59: 331–353, 1990.

317. Nelson JS, Liaw LH, Berns MW. Tumor destruction in photodynamic therapy. Pho- 339. Pfeiffer DR, Kauffman RF, Lardy HA. Effects of N-ethylmaleimide on the limited
tochem Photobiol 46: 829 – 835, 1987. uptake of Ca2⫹, Mn2⫹, and Sr2⫹ by rat liver mitochondria. J Biol Chem 253: 4165–
4171, 1978.
318. Neupert W. Protein import into mitochondria. Annu Rev Biochem 66: 863–917, 1997.
340. Picard M, Taivassalo T, Gouspillou G, Hepple RT. Mitochondria: isolation, structure
319. Nosoh Y, Kajioka J, Itoh M. Effect of menadione on the electron transport pathway of and function. J Physiol 589: 4413– 4421, 2012.
yeast mitochondria. Arch Biochem Biophys 127: 1– 6, 1968.
341. Pinton P, Rimessi A, Marchi S, Orsini F, Migliaccio E, Giorgio M, Contursi C, Minucci
320. Novgorodov SA, Gudz TI, Kushnareva YE, Zorov DB, Kudrjashov YB. Effect of ADP/ S, Mantovani F, Wieckowski MR, Del Sal G, Pelicci PG, Rizzuto R. Protein kinase C
ATP antiporter conformational state on the suppression of the nonspecific permea- beta and prolyl isomerase 1 regulate mitochondrial effects of the life-span determi-
bility of the inner mitochondrial membrane by cyclosporine A. FEBS Lett 277: 123– nant p66Shc. Science 315: 659 – 663, 2007.
126, 1990.
342. Plotnikov EY, Kazachenko AV, Vyssokikh MY, Vasileva AK, Tcvirkun DV, Isaev NK,
321. Novgorodov SA, Gudz TI, Kushnareva YE, Zorov DB, Kudrjashov YB. Effect of cy- Kirpatovsky VI, Zorov DB. The role of mitochondria in oxidative and nitrosative stress
closporine A and oligomycin on non-specific permeability of the inner mitochondrial during ischemia/reperfusion in the rat kidney. Kidney Int 72: 1493–1502, 2007.
membrane. FEBS Lett 270: 108 –110, 1990.
343. Plotnikov EY, Morosanova MA, Pevzner IB, Zorova LD, Manskikh VN, Pulkova NV,
322. Novgorodov SA, Gudz TI, Mohr Yu E, Goncharenko EN, Yaguzhinsky LS. ATP- Galkina SI, Skulachev VP, Zorov DB. Protective effect of mitochondria-targeted an-
synthase complex: the mechanism of control of ion fluxes induced by cumene hy- tioxidants in an acute bacterial infection. Proc Natl Acad Sci USA 110: E3100 –3108,
droperoxide in mitochondria. FEBS Lett 247: 255–258, 1989. 2013.

323. O’Rourke B. Pathophysiological and protective roles of mitochondrial ion channels. J 344. Popova EN, Pletjushkina OY, Dugina VB, Domnina LV, Ivanova OY, Izyumov DS,
Physiol 529: 23–36, 2000. Skulachev VP, Chernyak BV. Scavenging of reactive oxygen species in mitochondria
induces myofibroblast differentiation. Antioxid Redox Signal 13: 1297–1307, 2010.
324. O’Rourke B, Ramza BM, Marban E. Oscillations of membrane current and excitability
driven by metabolic oscillations in heart cells. Science 265: 962–966, 1994. 345. Pryde KR, Hirst J. Superoxide is produced by the reduced flavin in mitochondrial
complex I: a single, unified mechanism that applies during both forward and reverse
325. Ohnishi T. Iron-sulfur clusters/semiquinones in complex I. Biochim Biophys Acta 1364: electron transfer. J Biol Chem 286: 18056 –18065, 2011.
186 –206, 1998.
346. Pushpakiran G, Mahalakshmi K, Anuradha CV. Taurine restores ethanol-induced de-
326. Olsen RK, Olpin SE, Andresen BS, Miedzybrodzka ZH, Pourfarzam M, Merinero B, pletion of antioxidants and attenuates oxidative stress in rat tissues. Amino Acids 27:
Frerman FE, Beresford MW, Dean JC, Cornelius N, Andersen O, Oldfors A, Holme E, 91–96, 2004.
Gregersen N, Turnbull DM, Morris AA. ETFDH mutations as a major cause of ribo-
flavin-responsive multiple acyl-CoA dehydrogenation deficiency. Brain 130: 2045– 347. Raha S, Robinson BH. Mitochondria, oxygen free radicals, and apoptosis. Am J Med
2054, 2007. Genet 106: 62–70, 2001.

946 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

348. Raha S, Robinson BH. Mitochondria, oxygen free radicals, disease and ageing. Trends 370. Sandri G, Panfili E, Ernster L. Hydrogen peroxide production by monoamine oxidase
Biochem Sci 25: 502–508, 2000. in isolated rat-brain mitochondria: its effect on glutathione levels and Ca2⫹ efflux.
Biochim Biophys Acta 1035: 300 –305, 1990.
349. Rajasekaran NS, Connell P, Christians ES, Yan LJ, Taylor RP, Orosz A, Zhang XQ,
Stevenson TJ, Peshock RM, Leopold JA, Barry WH, Loscalzo J, Odelberg SJ, Benjamin 371. Sato K, Kashiwaya Y, Keon CA, Tsuchiya N, King MT, Radda GK, Chance B, Clarke K,
IJ. Human alpha B-crystallin mutation causes oxido-reductive stress and protein ag- Veech RL. Insulin, ketone bodies, and mitochondrial energy transduction. FASEB J 9:
gregation cardiomyopathy in mice. Cell 130: 427– 439, 2007. 651– 658, 1995.

350. Ramsay RR, Singer TP. Relation of superoxide generation and lipid peroxidation to the 372. Scarpa A, Azzone GF. The mechanism of ion translocation in mitochondria. 4. Cou-
inhibition of NADH-Q oxidoreductase by rotenone, piericidin A, and MPP⫹. Biochem pling of K⫹ efflux with Ca2⫹ uptake. Eur J Biochem 12: 328 –335, 1970.
Biophys Res Commun 189: 47–52, 1992.
373. Scholz W, Schutze K, Kunz W, Schwarz M. Phenobarbital enhances the formation of
351. Rasola A, Bernardi P. Mitochondrial permeability transition in Ca2⫹-dependent apo- reactive oxygen in neoplastic rat liver nodules. Cancer Res 50: 7015–7022, 1990.
ptosis and necrosis. Cell Calcium 50: 222–233, 2011.
374. Schumacker PT. Hypoxia, anoxia, and O2 sensing: the search continues. Am J Physiol
352. Ray PD, Huang BW, Tsuji Y. Reactive oxygen species (ROS) homeostasis and redox Lung Cell Mol Physiol 283: L918 –L921, 2002.
regulation in cellular signaling. Cell Signal 24: 981–990, 2012.
375. Schwartz Longacre L, Kloner RA, Arai AE, Baines CP, Bolli R, Braunwald E, Downey
353. Rebrin I, Kamzalov S, Sohal RS. Effects of age and caloric restriction on glutathione J, Gibbons RJ, Gottlieb RA, Heusch G, Jennings RB, Lefer DJ, Mentzer RM, Murphy E,
redox state in mice. Free Radic Biol Med 35: 626 – 635, 2003. Ovize M, Ping P, Przyklenk K, Sack MN, Vander Heide RS, Vinten-Johansen J, Yellon
DM. New horizons in cardioprotection: recommendations from the 2010 National
354. Reeder BJ. The redox activity of hemoglobins: from physiologic functions to patho- Heart, Lung, and Blood Institute Workshop. Circulation 124: 1172–1179, 2011.
logic mechanisms. Antioxid Redox Signal 13: 1087–1123, 2010.
376. Schwarzlander M, Logan DC, Fricker MD, Sweetlove LJ. The circularly permuted
355. Rieske JS, Baum H, Stoner CD, Lipton SH. On the antimycin-sensitive cleavage of yellow fluorescent protein cpYFP that has been used as a superoxide probe is highly
complex 3 of the mitochondrial respiratory chain. J Biol Chem 242: 4854 – 4866, 1967. responsive to pH but not superoxide in mitochondria: implications for the existence
of superoxide ”flashes.“ Biochem J 437: 381–387, 2011.
356. Rivera J, Sobey CG, Walduck AK, Drummond GR. Nox isoforms in vascular patho-
physiology: insights from transgenic and knockout mouse models. Redox Rep 15: 377. Schwarzlander M, Murphy MP, Duchen MR, Logan DC, Fricker MD, Halestrap AP,
50 – 63, 2010. Muller FL, Rizzuto R, Dick TP, Meyer AJ, Sweetlove LJ. Mitochondrial ”flashes“: a
radical concept repHined. Trends Cell Biol 22: 503–508, 2012.
357. Rizzuto R, Simpson AW, Brini M, Pozzan T. Rapid changes of mitochondrial Ca2⫹
revealed by specifically targeted recombinant aequorin. Nature 358: 325–327, 1992. 378. Schweizer M, Richter C. Nitric oxide potently and reversibly deenergizes mitochon-
dria at low oxygen tension. Biochem Biophys Res Commun 204: 169 –175, 1994.
358. Robinson BH. Human complex I deficiency: clinical spectrum and involvement of
oxygen free radicals in the pathogenicity of the defect. Biochim Biophys Acta 1364: 379. Segretain D, Rambourg A, Clermont Y. Three dimensional arrangement of mitochon-
271–286, 1998. dria and endoplasmic reticulum in the heart muscle fiber of the rat. Anat Rec 200:
139 –151, 1981.
359. Rodrigo GC, Lawrence CL, Standen NB. Dinitrophenol pretreatment of rat ventric-
ular myocytes protects against damage by metabolic inhibition and reperfusion. J Mol 380. Seifert EL, Estey C, Xuan JY, Harper ME. Electron transport chain-dependent and
Cell Cardiol 34: 555–569, 2002. -independent mechanisms of mitochondrial H2O2 emission during long-chain fatty
acid oxidation. J Biol Chem 285: 5748 –5758, 2010.
360. Romashko DN, Marban E, O’Rourke B. Subcellular metabolic transients and mito-
chondrial redox waves in heart cells. Proc Natl Acad Sci USA 95: 1618 –1623, 1998. 381. Selivanov VA, Zeak JA, Roca J, Cascante M, Trucco M, Votyakova TV. The role of
external and matrix pH in mitochondrial reactive oxygen species generation. J Biol
361. Rotig A, de Lonlay P, Chretien D, Foury F, Koenig M, Sidi D, Munnich A, Rustin P. Chem 283: 29292–29300, 2008.
Aconitase and mitochondrial iron-sulphur protein deficiency in Friedreich ataxia. Nat
Genet 17: 215–217, 1997. 382. Sena LA, Chandel NS. Physiological roles of mitochondrial reactive oxygen species.
Mol Cell 48: 158 –167, 2012.
362. Rugarli EI, Langer T. Mitochondrial quality control: a matter of life and death for
neurons. EMBO J 31: 1336 –1349, 2012. 383. Sena LA, Li S, Jairaman A, Prakriya M, Ezponda T, Hildeman DA, Wang CR, Schu-
macker PT, Licht JD, Perlman H, Bryce PJ, Chandel NS. Mitochondria are required for
363. Ruzicka FJ, Beinert H. A new iron-sulfur flavoprotein of the respiratory chain. A antigen-specific T cell activation through reactive oxygen species signaling. Immunity
component of the fatty acid beta oxidation pathway. J Biol Chem 252: 8440 – 8445, 38: 225–236, 2013.
1977.
384. Severs NJ, Slade AM, Powell T, Twist VW, Warren RL. Correlation of ultrastructure
364. Sakai C, Tomitsuka E, Esumi H, Harada S, Kita K. Mitochondrial fumarate reductase as and function in calcium-tolerant myocytes isolated from the adult rat heart. J Ultra-
a target of chemotherapy: From parasites to cancer cells. Biochim Biophys Acta 1820: struct Res 81: 222–239, 1982.
643– 651, 2012.
385. Shanmuganathan S, Hausenloy DJ, Duchen MR, Yellon DM. Mitochondrial permea-
365. Saks V, Dzeja PP, Guzun R, Aliev MK, Vndeli M, Terzic A, Wallimann T. System Analysis bility transition pore as a target for cardioprotection in the human heart. Am J Physiol
of Cardiac Energetics-Excitation-Contraction Coupling. Integration of Mitochondrial Res- Heart Circ Physiol 289: H237–H242, 2005.
piration. Darmstadt: Wiley-VCH Verlag, 2007, p. 367– 405.
386. Shih JC, Chen K, Ridd MJ. Monoamine oxidase: from genes to behavior. Annu Rev
366. Salet C, Moreno G. Photosensitization of mitochondria. Molecular and cellular as- Neurosci 22: 197–217, 1999.
pects. J Photochem Photobiol B 5: 133–150, 1990.
387. Shulga N, Wilson-Smith R, Pastorino JG. Sirtuin-3 deacetylation of cyclophilin D in-
367. Salet C, Moreno G, Ricchelli F, Bernardi P. Singlet oxygen produced by photodynamic duces dissociation of hexokinase II from the mitochondria. J Cell Sci 123: 894 –902,
action causes inactivation of the mitochondrial permeability transition pore. J Biol 2010.
Chem 272: 21938 –21943, 1997.
388. Silachev DN, Isaev NK, Pevzner IB, Zorova LD, Stelmashook EV, Novikova SV,
368. Samartsev VN, Smirnov AV, Zeldi IP, Markova OV, Mokhova EN, Skulachev VP. Plotnikov EY, Skulachev VP, Zorov DB. The mitochondria-targeted antioxidants and
Involvement of aspartate/glutamate antiporter in fatty acid-induced uncoupling of remote kidney preconditioning ameliorate brain damage through kidney-to-brain
liver mitochondria. Biochim Biophys Acta 1319: 251–257, 1997. cross-talk. PLoS One 7: e51553, 2012.

369. Samec S, Seydoux J, Dulloo AG. Role of UCP homologues in skeletal muscles and 389. Simonson SG, Zhang J, Canada AT Jr, Su YF, Benveniste H, Piantadosi CA. Hydrogen
brown adipose tissue: mediators of thermogenesis or regulators of lipids as fuel peroxide production by monoamine oxidase during ischemia-reperfusion in the rat
substrate? FASEB J 12: 715–724, 1998. brain. J Cereb Blood Flow Metab 13: 125–134, 1993.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 947


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

390. Singla M, Guzman G, Griffin AJ, Bharati S. Cardiomyopathy in multiple Acyl-CoA 411. Spencer JP, Jenner P, Daniel SE, Lees AJ, Marsden DC, Halliwell B. Conjugates of
dehydrogenase deficiency: a clinico-pathological correlation and review of literature. catecholamines with cysteine and GSH in Parkinson’s disease: possible mechanisms of
Pediatr Cardiol 29: 446 – 451, 2008. formation involving reactive oxygen species. J Neurochem 71: 2112–2122, 1998.

391. Skulachev MV, Antonenko YN, Anisimov VN, Chernyak BV, Cherepanov DA, 412. Srere PA. Protein crystals as a model for mitochondrial matrix proteins. Trends
Chistyakov VA, Egorov MV, Kolosova NG, Korshunova GA, Lyamzaev KG, Plotnikov Biochem Sci 6: 4 –7, 1981.
EY, Roginsky VA, Savchenko AY, Severina II, Severin FF, Shkurat TP, Tashlitsky VN,
Shidlovsky KM, Vyssokikh MY, Zamyatnin AA Jr, Zorov DB, Skulachev VP. Mitochon- 413. Srere PA. The structure of the mitochondrial inner membrane - matrix compartment.
drial-targeted plastoquinone derivatives Effect on senescence and acute age-related Trends Biochem Sci 7: 377–378, 1982.
pathologies. Curr Drug Targets 12: 800 – 826, 2011.
414. St-Pierre J, Buckingham JA, Roebuck SJ, Brand MD. Topology of superoxide produc-
392. Skulachev V. Cationic antioxidants as a powerful tool against mitochondrial oxidative tion from different sites in the mitochondrial electron transport chain. J Biol Chem 277:
stress. Biochem Biophys Res Commun 441: 275–279, 2013. 44784 – 44790, 2002.

393. Skulachev VP. A biochemical approach to the problem of aging: ”megaproject“ on 415. Starkov AA. “Mild” uncoupling of mitochondria. Biosci Rep 17: 273–279, 1997.
membrane-penetrating ions. The first results and prospects. Biochemistry 72: 1385–
416. Starkov AA. The role of mitochondria in reactive oxygen species metabolism and
1396, 2007.
signaling. Ann NY Acad Sci 1147: 37–52, 2008.
394. Skulachev VP. Energy transformation in the respiratory chain. Curr Top Bioenerg 4:
417. Starkov AA. An update on the role of mitochondrial alpha-ketoglutarate dehydroge-
127, 1971.
nase in oxidative stress. Mol Cell Neurosci 55: 13–16, 2013.
395. Skulachev VP. Mitochondrial filaments and clusters as intracellular power-transmitting
cables. Trends Biochem Sci 26: 23–29, 2001. 418. Starkov AA, Fiskum G. Myxothiazol induces H2O2 production from mitochondrial
respiratory chain. Biochem Biophys Res Commun 281: 645– 650, 2001.
396. Skulachev VP. Power transmission along biological membranes. J Membr Biol 114:
97–112, 1990. 419. Starkov AA, Fiskum G. Regulation of brain mitochondrial H2O2 production by mem-
brane potential and NAD(P)H redox state. J Neurochem 86: 1101–1107, 2003.
397. Skulachev VP. Role of uncoupled and non-coupled oxidations in maintenance of safely
low levels of oxygen and its one-electron reductants. Q Rev Biophys 29: 169 –202, 420. Starkov AA, Fiskum G, Chinopoulos C, Lorenzo BJ, Browne SE, Patel MS, Beal MF.
1996. Mitochondrial alpha-ketoglutarate dehydrogenase complex generates reactive oxy-
gen species. J Neurosci 24: 7779 –7788, 2004.
398. Skulachev VP. Uncoupling: new approaches to an old problem of bioenergetics.
Biochim Biophys Acta 1363: 100 –124, 1998. 421. Stelmashook EV, Isaev NK, Lozier ER, Goryacheva ES, Khaspekov LG. Role of glu-
tamine in neuronal survival and death during brain ischemia and hypoglycemia. Int J
399. Skulachev VP, Anisimov VN, Antonenko YN, Bakeeva LE, Chernyak BV, Erichev VP, Neurosci 121: 415– 422, 2011.
Filenko OF, Kalinina NI, Kapelko VI, Kolosova NG, Kopnin BP, Korshunova GA,
Lichinitser MR, Obukhova LA, Pasyukova EG, Pisarenko OI, Roginsky VA, Ruuge EK, 422. Stelmashook EV, Lozier ER, Goryacheva ES, Mergenthaler P, Novikova SV, Zorov
Senin II, Severina II, Skulachev MV, Spivak IM, Tashlitsky VN, Tkachuk VA, Vyssokikh DB, Isaev NK. Glutamine-mediated protection from neuronal cell death depends on
MY, Yaguzhinsky LS, Zorov DB. An attempt to prevent senescence: a mitochondrial mitochondrial activity. Neurosci Lett 482: 151–155, 2010.
approach. Biochim Biophys Acta 1787: 437– 461, 2009.
423. Stern MD, Cheng H. Putting out the fire: what terminates calcium-induced calcium
400. Skulachev VP, Bogachev AV, Kasparinsky FO. Principles of Bioenergetics. Berlin: release in cardiac muscle? Cell Calcium 35: 591– 601, 2004.
Springer-Verlag, 2013, p. 436.
424. Stock D, Gibbons C, Arechaga I, Leslie AG, Walker JE. The rotary mechanism of ATP
401. Slodzinski MK, Aon MA, O’Rourke B. Glutathione oxidation as a trigger of mitochon- synthase. Curr Opin Struct Biol 10: 672– 679, 2000.
drial depolarization and oscillation in intact hearts. J Mol Cell Cardiol 45: 650 – 660,
425. Stokes AH, Hastings TG, Vrana KE. Cytotoxic and genotoxic potential of dopamine. J
2008.
Neurosci Res 55: 659 – 665, 1999.
402. Smeitink J, van den Heuvel L, DiMauro S. The genetics and pathology of oxidative
426. Stoner CD, Sirak HD. Adenine nucleotide-induced contraction of the inner mitochon-
phosphorylation. Nat Rev Genet 2: 342–352, 2001.
drial membrane. I. General characterization. J Cell Biol 56: 51– 64, 1973.
403. Smiley ST, Reers M, Mottola-Hartshorn C, Lin M, Chen A, Smith TW, Steele GD Jr,
427. Stoner CD, Sirak HD. Adenine nucleotide-induced contraction on the inner mito-
Chen LB. Intracellular heterogeneity in mitochondrial membrane potentials revealed
chondrial membrane. II. Effect of bongkrekic acid. J Cell Biol 56: 65–73, 1973.
by a J-aggregate-forming lipophilic cation JC-1. Proc Natl Acad Sci USA 88: 3671–3675,
1991. 428. Sun Y, Jin K, Peel A, Mao XO, Xie L, Greenberg DA. Neuroglobin protects the brain
from experimental stroke in vivo. Proc Natl Acad Sci USA 100: 3497–3500, 2003.
404. Smith RA, Adlam VJ, Blaikie FH, Manas AR, Porteous CM, James AM, Ross MF, Logan
A, Cocheme HM, Trnka J, Prime TA, Abakumova I, Jones BA, Filipovska A, Murphy 429. Sundaresan NR, Gupta M, Kim G, Rajamohan SB, Isbatan A, Gupta MP. Sirt3 blocks
MP. Mitochondria-targeted antioxidants in the treatment of disease. Ann NY Acad Sci the cardiac hypertrophic response by augmenting Foxo3a-dependent antioxidant
1147: 105–111, 2008.
defense mechanisms in mice. J Clin Invest 119: 2758 –2771, 2009.
405. Smith RA, Hartley RC, Cocheme HM, Murphy MP. Mitochondrial pharmacology.
430. Szabo I, Zoratti M. The giant channel of the inner mitochondrial membrane is inhibited
Trends Pharmacol Sci 33: 341–352, 2012.
by cyclosporin A. J Biol Chem 266: 3376 –3379, 1991.
406. Smith RA, Murphy MP. Mitochondria-targeted antioxidants as therapies. Discov Med
431. Takahashi N, Hayano T, Suzuki M. Peptidyl-prolyl cis-trans isomerase is the cyclo-
11: 106 –114, 2011.
sporin A-binding protein cyclophilin. Nature 337: 473– 475, 1989.
407. Smith RA, Porteous CM, Coulter CV, Murphy MP. Selective targeting of an antioxi-
432. Takeshige K, Minakami S. NADH- and NADPH-dependent formation of superoxide
dant to mitochondria. Eur J Biochem 263: 709 –716, 1999.
anions by bovine heart submitochondrial particles and NADH-ubiquinone reductase
408. Smith RA, Porteous CM, Gane AM, Murphy MP. Delivery of bioactive molecules to preparation. Biochem J 180: 129 –135, 1979.
mitochondria in vivo. Proc Natl Acad Sci USA 100: 5407–5412, 2003.
433. Tannahill GM, Curtis AM, Adamik J, Palsson-McDermott EM, McGettrick AF, Goel G,
409. Sokolove PM, Kester MB, Haynes J. Interaction of adriamycin aglycones with isolated Frezza C, Bernard NJ, Kelly B, Foley NH, Zheng L, Gardet A, Tong Z, Jany SS, Corr
mitochondria. Effect of selenium deficiency. Biochem Pharmacol 46: 691– 697, 1993. SC, Haneklaus M, Caffrey BE, Pierce K, Walmsley S, Beasley FC, Cummins E, Nizet V,
Whyte M, Taylor CT, Lin H, Masters SL, Gottlieb E, Kelly VP, Clish C, Auron PE,
410. Sorgato MC, Keller BU, Stuhmer W. Patch-clamping of the inner mitochondrial mem- Xavier RJ, O’Neill LA. Succinate is an inflammatory signal that induces IL-1beta
brane reveals a voltage-dependent ion channel. Nature 330: 498 –500, 1987. through HIF-1alpha. Nature 496: 238 –242, 2013.

948 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
MITOCHONDRIAL ROS AND ROS-INDUCED ROS RELEASE

434. Tarasov AI, Griffiths EJ, Rutter GA. Regulation of ATP production by mitochondrial 456. Votyakova TV, Reynolds IJ. DeltaPsi(m)-dependent and -independent production of
Ca2⫹. Cell Calcium 52: 28 –35, 2012. reactive oxygen species by rat brain mitochondria. J Neurochem 79: 266 –277, 2001.

435. Territo PR, Mootha VK, French SA, Balaban RS. Ca2⫹ activation of heart mitochon- 457. Vyssokikh MY, Katz A, Rueck A, Wuensch C, Dorner A, Zorov DB, Brdiczka D.
drial oxidative phosphorylation: role of the F0/F1-ATPase. Am J Physiol Cell Physiol 278: Adenine nucleotide translocator isoforms 1 and 2 are differently distributed in the
C423–C435, 2000. mitochondrial inner membrane and have distinct affinities to cyclophilin D. Biochem J
358: 349 –358, 2001.
436. Tomitsuka E, Kita K, Esumi H. The NADH-fumarate reductase system, a novel mito-
chondrial energy metabolism, is a new target for anticancer therapy in tumor mi- 458. Waggoner AS. Dye indicators of membrane potential. Annu Rev Biophys Bioeng 8:
croenvironments. Ann NY Acad Sci 1201: 44 – 49, 2010. 47– 68, 1979.

437. Tormos KV, Anso E, Hamanaka RB, Eisenbart J, Joseph J, Kalyanaraman B, Chandel 459. Waggoner AS. The use of cyanine dyes for the determination of membrane potentials
NS. Mitochondrial complex III ROS regulate adipocyte differentiation. Cell Metab 14: in cells, organelles, and vesicles. Methods Enzymol 55: 689 – 695, 1979.
537–544, 2011.
460. Wahllander A, Soboll S, Sies H, Linke I, Muller M. Hepatic mitochondrial and cytosolic
438. Tretter L, Adam-Vizi V. Generation of reactive oxygen species in the reaction cata- glutathione content and the subcellular distribution of GSH-S-transferases. FEBS Lett
lyzed by alpha-ketoglutarate dehydrogenase. J Neurosci 24: 7771–7778, 2004. 97: 138 –140, 1979.

439. Trinei M, Giorgio M, Cicalese A, Barozzi S, Ventura A, Migliaccio E, Milia E, Padura IM, 461. Wang W, Fang H, Groom L, Cheng A, Zhang W, Liu J, Wang X, Li K, Han P, Zheng M,
Raker VA, Maccarana M, Petronilli V, Minucci S, Bernardi P, Lanfrancone L, Pelicci PG. Yin J, Wang W, Mattson MP, Kao JP, Lakatta EG, Sheu SS, Ouyang K, Chen J, Dirksen
A p53-p66Shc signalling pathway controls intracellular redox status, levels of oxida- RT, Cheng H. Superoxide flashes in single mitochondria. Cell 134: 279 –290, 2008.
tion-damaged DNA and oxidative stress-induced apoptosis. Oncogene 21: 3872–
3878, 2002. 462. Waypa GB, Schumacker PT. O2 sensing in hypoxic pulmonary vasoconstriction: the
mitochondrial door re-opens. Respir Physiol Neurobiol 132: 81–91, 2002.
440. Trotter EW, Grant CM. Thioredoxins are required for protection against a reductive
stress in the yeast Saccharomyces cerevisiae. Mol Microbiol 46: 869 – 878, 2002. 463. Weber NE, Blair PV. Ultrastructural studies of beef heart mitochondria. II. Adenine
nucleotide induced modifications of mitochondrial morphology. Biochem Biophys Res
441. Trouillard M, Meunier B, Rappaport F. Questioning the functional relevance of mito- Commun 41: 821– 829, 1970.
chondrial supercomplexes by time-resolved analysis of the respiratory chain. Proc Natl
Acad Sci USA 108: E1027–1034, 2011. 464. Weber NE, Blair PV. Ultrastuctural studies of beef heart mitochondria. I. Effects of
adenosine diphosphate on mitochondrial morphology. Biochem Biophys Res Commun
442. Trumpower BL. The protonmotive Q cycle. Energy transduction by coupling of
36: 987–993, 1969.
proton translocation to electron transfer by the cytochrome bc1 complex. J Biol Chem
265: 11409 –11412, 1990. 465. Wei AC, Liu T, Cortassa S, Winslow RL, O’Rourke B. Mitochondrial Ca2⫹ influx and
efflux rates in guinea pig cardiac mitochondria: low and high affinity effects of cyclo-
443. Turnbull DM, Bartlett K, Eyre JA, Gardner-Medwin D, Johnson MA, Fisher J, Wat-
sporine A. Biochim Biophys Acta 1813: 1373–1381, 2011.
mough NJ. Lipid storage myopathy due to glutaric aciduria type II: treatment of a
potentially fatal myopathy. Dev Med Child Neurol 30: 667– 672, 1988. 466. Wei L, Dirksen RT. Perspectives on: SGP symposium on mitochondrial physiology and
medicine: mitochondrial superoxide flashes: from discovery to new controversies. J
444. Turrens JF. Mitochondrial formation of reactive oxygen species. J Physiol 552: 335–
Gen Physiol 139: 425– 434, 2012.
344, 2003.
467. Weinberg F, Hamanaka R, Wheaton WW, Weinberg S, Joseph J, Lopez M, Kalyanara-
445. Twig G, Graf SA, Wikstrom JD, Mohamed H, Haigh SE, Elorza A, Deutsch M, Zurgil
man B, Mutlu GM, Budinger GR, Chandel NS. Mitochondrial metabolism and ROS
N, Reynolds N, Shirihai OS. Tagging and tracking individual networks within a com-
generation are essential for Kras-mediated tumorigenicity. Proc Natl Acad Sci USA 107:
plex mitochondrial web with photoactivatable GFP. Am J Physiol Cell Physiol 291:
8788 – 8793, 2010.
C176 –C184, 2006.
468. Weisiger RA, Fridovich I. Superoxide dismutase. Organelle specificity. J Biol Chem 248:
446. Vacek TP, Vacek JC, Tyagi SC. Mitochondrial mitophagic mechanisms of myocardial
3582–3592, 1973.
matrix metabolism and remodelling. Arch Physiol Biochem 118: 31– 42, 2012.

447. Valko M, Izakovic M, Mazur M, Rhodes CJ, Telser J. Role of oxygen radicals in DNA 469. Weisova P, Anilkumar U, Ryan C, Concannon CG, Prehn JH, Ward MW. “Mild mito-
damage and cancer incidence. Mol Cell Biochem 266: 37–56, 2004. chondrial uncoupling” induced protection against glutamate excitotoxicity in primary
neurons requires AMPK activity. Biochim Biophys Acta 1817: 744 –753, 2012.
448. Van den Berg CJ, Garfinkel D. A stimulation study of brain compartments. Metabolism
of glutamate and related substances in mouse brain. Biochem J 123: 211–218, 1971. 470. Whatley SA, Curti D, Das Gupta F, Ferrier IN, Jones S, Taylor C, Marchbanks RM.
Superoxide, neuroleptics and the ubiquinone and cytochrome b5 reductases in brain
449. Vanden Hoek TL, Becker LB, Shao Z, Li C, Schumacker PT. Reactive oxygen species and lymphocytes from normals and schizophrenic patients. Mol Psychiatry 3: 227–237,
released from mitochondria during brief hypoxia induce preconditioning in cardiomy- 1998.
ocytes. J Biol Chem 273: 18092–18098, 1998.
471. Wiesner RJ, Rosen P, Grieshaber MK. Pathways of succinate formation and their
450. Vasington FD, Gazzotti P, Tiozzo R, Carafoli E. The effect of ruthenium red on Ca2⫹ contribution to improvement of cardiac function in the hypoxic rat heart. Biochem
transport and respiration in rat liver mitochondria. Biochim Biophys Acta 256: 43–54, Med Metab Biol 40: 19 –34, 1988.
1972.
472. Williams RJ. Chemical advances in evolution by and changes in use of space during
451. Vasington FD, Murphy JV. Ca ion uptake by rat kidney mitochondria and its depen- time. J Theor Biol 268: 146 –159, 2011.
dence on respiration and phosphorylation. J Biol Chem 237: 2670 –2677, 1962.
473. Williams RJ. Possible functions of chains of catalysts. J Theor Biol 1: 1–17, 1961.
452. Vasquez-Vivar J, Kalyanaraman B, Kennedy MC. Mitochondrial aconitase is a source of
hydroxyl radical. An electron spin resonance investigation. J Biol Chem 275: 14064 – 474. Williamson JR, Corkey BE. Assay of citric acid cycle intermediates and related com-
14069, 2000. pounds– update with tissue metabolite levels and intracellular distribution. Methods
Enzymol 55: 200 –222, 1979.
453. Verbon EH, Post JA, Boonstra J. The influence of reactive oxygen species on cell cycle
progression in mammalian cells. Gene 511: 1– 6, 2012. 475. Wittenberg BA, Wittenberg JB. Oxygen pressure gradients in isolated cardiac myo-
cytes. J Biol Chem 260: 6548 – 6554, 1985.
454. Vinogradov AD. Catalytic properties of the mitochondrial NADH-ubiquinone oxi-
doreductase (complex I) and the pseudo-reversible active/inactive enzyme transition. 476. Wojtczak L, Schonfeld P. Effect of fatty acids on energy coupling processes in mito-
Biochim Biophys Acta 1364: 169 –185, 1998. chondria. Biochim Biophys Acta 1183: 41–57, 1993.

455. Vinogradov AD, Grivennikova VG. Generation of superoxide-radical by the NADH: 477. Wu CC, Liu YB, Lu LS, Lin CW. Imaging reactive oxygen species dynamics in living cells
ubiquinone oxidoreductase of heart mitochondria. Biochemistry 70: 120 –127, 2005. and tissues. Front Biosci 1: 39 – 44, 2009.

Physiol Rev • VOL 94 • JULY 2014 • www.prv.org 949


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.
ZOROV ET AL.

478. Wu YN, Munhall AC, Johnson SW. Mitochondrial uncoupling agents antagonize rote- 490. Zorov DB, Bannikova SY, Belousov VV, Vyssokikh MY, Zorova LD, Isaev NK, Kras-
none actions in rat substantia nigra dopamine neurons. Brain Res 1395: 86 –93, 2011. nikov BF, Plotnikov EY. Reactive oxygen and nitrogen species: friends or foes? Bio-
chemistry 70: 215–221, 2005.
479. Yaniv Y, Juhaszova M, Nuss HB, Wang S, Zorov DB, Lakatta EG, Sollott SJ. Matching
ATP supply and demand in mammalian heart: in vivo, in vitro, and in silico perspec- 491. Zorov DB, Filburn CR, Klotz LO, Zweier JL, Sollott SJ. Reactive oxygen species
tives. Ann NY Acad Sci 1188: 133–142, 2010. (ROS)-induced ROS release: a new phenomenon accompanying induction of the
mitochondrial permeability transition in cardiac myocytes. J Exp Med 192: 1001–1014,
480. Yaniv Y, Juhaszova M, Sollott SJ. Age-related changes in myocardical ATP supply and
2000.
demand mechanisms. Trends in Endocinol Metab 24: 495–505, 2013.
492. Zorov DB, Isaev NK, Plotnikov EY, Silachev DN, Zorova LD, Pevzner IB, Morosanova
481. Yankovskaya V, Horsefield R, Tornroth S, Luna-Chavez C, Miyoshi H, Leger C, Byrne
B, Cecchini G, Iwata S. Architecture of succinate dehydrogenase and reactive oxygen MA, Jankauskas SS, Zorov SD, Babenko VA. Perspectives of mitochondrial medicine.
species generation. Science 299: 700 –704, 2003. Biochemistry 78: 979 –990, 2013.

482. Zhang L, Yu L, Yu CA. Generation of superoxide anion by succinate-cytochrome c 493. Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial ROS-induced ROS release: an up-
reductase from bovine heart mitochondria. J Biol Chem 273: 33972–33976, 1998. date and review. Biochim Biophys Acta 1757: 509 –517, 2006.

483. Zhang M, Brewer AC, Schroder K, Santos CX, Grieve DJ, Wang M, Anilkumar N, Yu 494. Zorov DB, Juhaszova M, Yaniv Y, Nuss HB, Wang S, Sollott SJ. Regulation and phar-
B, Dong X, Walker SJ, Brandes RP, Shah AM. NADPH oxidase-4 mediates protection macology of the mitochondrial permeability transition pore. Cardiovasc Res 83: 213–
against chronic load-induced stress in mouse hearts by enhancing angiogenesis. Proc 225, 2009.
Natl Acad Sci USA 107: 18121–18126, 2010.
495. Zorov DB, Kinnally KW, Tedeschi H. Voltage activation of heart inner mitochondrial
484. Zhou L, Aon MA, Almas T, Cortassa S, Winslow RL, O’Rourke B. A reaction-diffusion membrane channels. J Bioenerg Biomembr 24: 119 –124, 1992.
model of ROS-induced ROS release in a mitochondrial network. PLoS Comput Biol 6:
e1000657, 2010. 496. Zorov DB, Kobrinsky E, Juhaszova M, Sollott SJ. Examining intracellular organelle
function using fluorescent probes: from animalcules to quantum dots. Circ Res 95:
2⫹
485. Zhou L, Aon MA, Liu T, O’Rourke B. Dynamic modulation of Ca sparks by mito- 239 –252, 2004.
chondrial oscillations in isolated guinea pig cardiomyocytes under oxidative stress. J
Mol Cell Cardiol 51: 632– 639, 2011. 497. Zorov DB, Krasnikov BF, Kuzminova AE, Vysokikh M, Zorova Mitochondria revisited
LD. Alternative functions of mitochondria. Biosci Rep 17: 507–520, 1997.
486. Zhou L, O’Rourke B. Cardiac mitochondrial network excitability: insights from com-
putational analysis. Am J Physiol Heart Circ Physiol 302: H2178 –H2189, 2012. 498. Zorov DB, Plotnikov EY, Jankauskas SS, Isaev NK, Silachev DN, Zorova LD, Pevzner
IB, Pulkova NV, Zorov SD, Morosanova MA. The phenoptosis problem: what is
487. Zhu J, Rebecchi MJ, Glass PS, Brink PR, Liu L. Interactions of GSK-3beta with mito-
causing the death of an organism? Biochemistry 77: 742–753, 2012.
chondrial permeability transition pore modulators during preconditioning: age-asso-
ciated differences. J Gerontol A Biol Sci Med Sci 68: 395– 403, 2013. 499. Zweier JL. Measurement of superoxide-derived free radicals in the reperfused heart.
488. Zoccarato F, Toscano P, Alexandre A. Dopamine-derived dopaminochrome pro- Evidence for a free radical mechanism of reperfusion injury. J Biol Chem 263: 1353–
motes H2O2 release at mitochondrial complex I: stimulation by rotenone, control by 1357, 1988.
Ca2⫹, and relevance to Parkinson disease. J Biol Chem 280: 15587–15594, 2005.
500. Zweier JL, Flaherty JT, Weisfeldt ML. Direct measurement of free radical generation
489. Zoratti M, Szabo I. The mitochondrial permeability transition. Biochim Biophys Acta following reperfusion of ischemic myocardium. Proc Natl Acad Sci USA 84: 1404 –1407,
1241: 139 –176, 1995. 1987.

950 Physiol Rev • VOL 94 • JULY 2014 • www.prv.org


Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (046.148.120.214) on January 17, 2019.

You might also like