ArticlePDF AvailableLiterature Review

Role of apoptosis in cardiovascular disease

Authors:

Abstract and Figures

Apoptosis plays a key role in the pathogenesis in a variety of cardiovascular diseases due to loss of terminally differentiated cardiac myocytes. Cardiac myocytes undergoing apoptosis have been identified in tissue samples from patients suffering from myocardial infarction, diabetic cardiomyopathy, and end-stage congestive heart failure. Apoptosis is a highly regulated program of cell death and can be mediated by death receptors in the plasma membrane, as well as the mitochondria and the endoplasmic reticulum. The cell death program is activated in cardiac myocytes by various stressors including cytokines, increased oxidative stress and DNA damage. Many studies have demonstrated that inhibition of apoptosis is cardioprotective and can prevent the development of heart failure. This review provides a current overview of the evidence of apoptosis in cardiovascular diseases and discusses the molecular pathways involved in cardiac myocyte apoptosis.
No caption available
… 
Content may be subject to copyright.
CELL DEATH AND DISEASE
Role of apoptosis in cardiovascular disease
Youngil Lee ÆA
˚sa B. Gustafsson
Published online: 14 January 2009
ÓSpringer Science+Business Media, LLC 2009
Abstract Apoptosis plays a key role in the pathogenesis
in a variety of cardiovascular diseases due to loss of ter-
minally differentiated cardiac myocytes. Cardiac myocytes
undergoing apoptosis have been identified in tissue samples
from patients suffering from myocardial infarction, dia-
betic cardiomyopathy, and end-stage congestive heart
failure. Apoptosis is a highly regulated program of cell
death and can be mediated by death receptors in the plasma
membrane, as well as the mitochondria and the endoplas-
mic reticulum. The cell death program is activated in
cardiac myocytes by various stressors including cytokines,
increased oxidative stress and DNA damage. Many studies
have demonstrated that inhibition of apoptosis is cardio-
protective and can prevent the development of heart
failure. This review provides a current overview of the
evidence of apoptosis in cardiovascular diseases and dis-
cusses the molecular pathways involved in cardiac
myocyte apoptosis.
Keywords Apoptosis Heart failure Death receptors
Mitochondria Bcl-2
Introduction
Cardiovascular disease is the leading cause of morbidity
and mortality in the developed world. A multitude of recent
studies suggest that loss of terminally differentiated cardiac
myocytes contributes to development of heart failure.
There are three morphologically and biochemically distinct
forms of cell death that occur in the heart; necrosis,
apoptosis, and possibly autophagy. Autophagy is a cellular
process that degrades long lived proteins and dysfunctional
organelles [1]. Autophagy is characterized by sequestration
of cytoplasm in double-membrane vesicles called auto-
phagosomes. Autophagosomes subsequently fuse with
lysosomes, leading to degradation of their content [1]. This
process is important for cellular homeostasis, and is a
survival response upregulated in response to stress or
starvation. However, excess autophagy can lead to cell
death due to excessive digestion of organelles and essential
proteins [2]. Necrosis is a passive form of cell death caused
by ATP depletion and rapid disruption of cell membrane
integrity resulting in spillage of intracellular contents into
interstitial and extracellular space, which initiates inflam-
mation and induces damage to neighboring cells [3,4]. In
contrast, apoptosis is an energy requiring form of pro-
grammed cell death whereby damaged cells are removed
without provoking inflammation. Apoptosis is character-
ized by chromatin condensation, DNA fragmentation,
plasma membrane blebbing (i.e., externalization of phos-
phatidylserine), and cell shrinkage due to reduction in
cytoplasm and organelles [57]. Finally, membrane-bound
apoptotic bodies containing cytosol and processed organ-
elles are formed and then removed by macrophages via
phagocytosis [8].
It is now evident that apoptosis plays a key role in the
pathogenesis of a variety of cardiovascular diseases.
Studies have reported that apoptosis occurs in myocardial
tissue samples from patients suffering from myocardial
infarction, dilated cardiomyopathy and end-stage heart
failure [912] as well as in animal models of ischemia–
reperfusion injury [1315]. Apoptosis is activated in car-
diac myocytes by multiple stressors that are commonly
seen in cardiovascular disease such as cytokine production
Y. Lee A
˚. B. Gustafsson (&)
BioScience Center, San Diego State University, 5500 Campanile
Drive, San Diego, CA 92182-4650, USA
e-mail: agustafs@sciences.sdsu.edu
123
Apoptosis (2009) 14:536–548
DOI 10.1007/s10495-008-0302-x
[16,17], increased oxidative stress [18], and DNA damage
[19]. Apoptosis is a highly regulated program of cell death
and inhibition of this process is cardioprotective under
many conditions. Therefore, this process represents a
potential target for therapeutic intervention to prevent heart
failure. This review provides a current overview of the
evidence and functional role of apoptosis in cardiovascular
diseases and discusses the molecular pathways involved in
cardiac myocyte apoptosis.
Evidence of apoptosis in cardiovascular diseases
Cardiac cells undergoing apoptosis have been detected in
many different diseases of the cardiovascular system
including atherosclerosis, myocardial ischemia and reper-
fusion injury, diabetic cardiomyopathy, and chronic heart
failure. Atherosclerotic vascular disease is a leading cause
of myocardial infarction and heart failure, and a major
factor in the development of acute coronary syndrome is
disruption of an atherosclerotic plaque. It has been reported
that macrophages and smooth muscle cells undergo apop-
tosis in unstable atherosclerotic plaques which can lead to
rupture of the plaque and thrombosis [20]. Recently, Clarke
et al. [21] demonstrated that vascular smooth muscle cell
apoptosis during either atherogenesis or within established
plaques of apolipoprotein (Apo) E(-/-) mice accelerated
plaque growth and was associated with features of plaque
vulnerability such as a thin fibrous cap and loss of collagen
and matrix. In addition, it has been proposed that poor
clearance of apoptotic macrophages may lead to accumu-
lation of cellular debris within the lipid-rich core of
atherosclerotic plaque, thus contributing to plaque pro-
gression and rupture. Rupture of atherosclerotic plaques
and thrombus formation cause occlusion of coronary
arteries which reduces the blood supply to the myocardium
and leads to myocardial infarction and possibly death.
Reperfusion is the treatment for acute myocardial infarc-
tion, but the reintroduction of oxygen can aggravate the
tissue damage, a process called reperfusion injury [22].
Loss of cardiac cells in response to ischemia/reperfusion
(I/R) injury was long considered to be due to necrotic cell
death, but studies over the past decade have identified
apoptosis as a significant component of cell loss during
reperfusion after a myocardial infarction [9,10,14].
Apoptosis seems to occur primarily after reperfusion fol-
lowing ischemia, whereas prolonged ischemia leads to
necrosis. There is also mounting evidence that apoptosis
plays an important role in both acute and chronic loss of
cardiac myocytes after a myocardial infarction. Studies
have reported the presence of apoptotic cells in the border
zone of the infarct and in remote myocardium in the early
phase [9], as well as months after myocardial infarction
[23], suggesting that apoptosis plays a role in remodeling
and development of heart failure after myocardial infarc-
tion. Since the regenerative capacity of the myocardium is
limited, there is intense interest in the prevention of car-
diomyocyte loss during ischemia and reperfusion.
Moreover, diabetes is a rapidly growing epidemic in the
Western world which is well known to increase the risk of
developing cardiovascular disease [24]. Increased levels of
apoptosis have been detected in the hearts of diabetic
patients and animal models of diabetes, and the loss of
myocytes has been implicated in the development of dia-
betic cardiomyopathy [25,26]. In addition, it has been
reported that the mortality of myocardial infarction in dia-
betic patients is more than double that of non-diabetic
patients [27]. It appears that cardiac myocytes in the dia-
betic myocardium are more susceptible to apoptosis [25],
suggesting that the higher sensitivity to MI is due to a higher
loss of cardiac myocytes via apoptosis in response to the
stress [28]. Also, Kuethe et al. [29] found that diabetic
patients with dilated cardiomyopathy had significantly more
apoptotic cells in the heart compared to patients without
diabetes. The pathogenesis of diabetic cardiomyopathy is a
complex process that is attributed to abnormal cellular
metabolism and defects in organelles such as mitochondria,
sarcolemma, and endoplasmic reticulum (ER) which leads
to activation of apoptosis. Hyperglycemia appears to play a
central role in the development of diabetic cardiomyopathy
by inducing apoptosis of cardiac myocytes via increased
levels of reactive oxygen and nitrogen species [30]. Using a
mouse model of diabetes, Cai et al. [25] found that
administration of insulin suppressed hyperglycemia and
inhibited diabetes-induced apoptosis in the heart. Other
studies have shown that suppression of myocardial cell
death by antioxidants or inhibition of apoptotic signaling
pathways result in a significant prevention of diabetic car-
diotoxicity, suggesting that oxidative-stress mediated
apoptosis plays an important role in the development of
diabetic cardiomyopathy [31,32].
There is also a connection between chronic heart failure
and apoptosis. It has been reported that patients with
advanced heart failure have higher rates of cardiac myocyte
apoptosis than normal subjects, (0.08–0.25 vs. 0.001–
0.002%) [9,10,33]. Genetic and pharmacological studies
from animal models suggest that apoptosis plays a crucial
role in the development and progression of heart failure.
Using transgenic mice that express a conditionally active
caspase-8 in the heart, Wencker et al. [34] found that
extremely low levels of chronic myocyte apoptosis were
sufficient to cause a lethal dilated cardiomyopathy. Inhi-
bition of cell death with a caspase inhibitor prevented left
ventricular dilation and improved ventricular function.
Thus, even if apoptosis occurs at a very low frequency after
an insult such as I/R, it is a constant process and results in
Apoptosis (2009) 14:536–548 537
123
progressive loss of myocytes, which gradually reduces the
ability of the heart to maintain contractile function. As a
result, the remaining myocytes are forced to work harder to
compensate for the loss in contractility, which, in turn, may
contribute to cardiac hypertrophy resulting in cardiomy-
opathy and heart failure.
Apoptotic signaling pathways in the heart
The major pathways involved in apoptotic signaling in the
heart involve the death receptor pathway, the mitochon-
drial and ER-stress death pathways. A schematic view of
apoptotic cascades is shown in Fig. 1.
Death receptor pathway in cardiovascular disease
Apoptosis can be initiated through activation of the death
receptor pathway (also called the extrinsic pathway) via a
complex signal transduction from the plasma membrane
leading to activation of the caspase cascade. The death
receptors belong to the tumor necrosis factor/nerve growth
factor receptor superfamily and are transmembrane pro-
teins with an extracellular ligand-interacting domain, a
transmembrane domain, and an intracellular death domain
[35]. Binding of the corresponding ligand to the death
receptor causes oligomerization and activation of the
receptor. For instance, after Fas ligand (FasL) binding, Fas
receptors undergo trimerization and recruit Fas-associated
death domain (FADD). Fas/FADD complex binds to pro-
caspase 8 leading to cleavage of the initiator pro-caspase 8
to active caspase 8. Activated caspase 8 propagates the
apoptotic signal through a direct activation of executioner
downstream caspases and via the release of cytochrome c
by mitochondria.
Several studies strongly suggest an important patho-
physiological role for the death receptor pathway in the
pathogenesis of heart failure. The best-characterized death
receptors are Fas (also called CD95 or Apo1) and tumor
necrosis factor receptor 1 (TNFR1). Both receptors are
present in cardiac myocytes and have been implicated to
contribute to cardiovascular disease. TRAIL is ligand for
Death Receptor 4 and 5 (DR4 and DR5) and has also been
reported to be released by cardiac myocytes [36], but not
much is known about this ligand and its receptors in the
heart. Patients with end-stage congestive heart failure have
elevated circulating levels of TNF-a, the ligand for TNFR1
[3739], and studies suggest that there is a relationship
between the serum levels of TNF- aand the severity of
heart failure [39,40]. In addition, several studies have
reported that cardiomyocytes can be an abundant source of
TNF-a[4143], and that failing human myocardium, but
not nonfailing hearts, express high levels of TNF-a[44,
45]. Following I/R, there is a sustained increase of TNF-a
both locally in the heart as well as in circulating levels in
blood [43,46]. The functional role of elevated levels of
TNF-ahas been investigated using transgenic mice over-
expressing TNF-aspecifically in the heart. For instance,
cardiac specific overexpression of TNF-ain transgenic
mice caused development of dilated cardiomyopathy and
heart failure [16,17], suggesting that increased levels of
Fig. 1 Scheme of apoptotic
signaling in cardiac myocytes.
Apoptosis can be mediated via
activation of death receptors,
permeabilization of the outer
mitochondrial membrane, or
ER-stress
538 Apoptosis (2009) 14:536–548
123
TNF-ais detrimental to the heart by activation of the death
receptor pathway.
In addition, both Fas and FasL expression are increased in
myocytes exposed to hypoxia [47] and activation of the Fas
pathway has been shown to induce apoptosis in cardiac
myocytes [48]. There is also evidence suggesting that Fas-
mediated apoptosis is associated with myocarditis [49],
myocardial reperfusion injury [50], and post-infarction
ventricular remodeling [51]. For instance, Gomez et al. [52]
recently reported that mice lacking functional Fas (lpr mice)
had no difference in infarct size and apoptosis after I/R,
suggesting that Fas plays a minor role in mediating acute I/R
injury. Similarly, Li et al. [51] found that lpr mice and mice
lacking Fas ligand (gld mice) had similar infarct sizes 2 days
after myocardial infarction. Instead, this study found
decreased apoptosis of granulation tissue cells which
reduced post-infarct remodeling. Granulation tissue cells are
removed via apoptosis to eventually make scar tissue [23].
Also, they found that infecting hearts with an adenovirus
encoding soluble Fas, a competitive inhibitor of FasL,
3 days after a myocardial infarction improved cardiac
function and survival. This suggests that the Fas pathway
contributes to cell death in the later stages of an infarction.
In contrast, two different studies found that the lpr mice
were resistant to acute I/R injury and hearts had reduced
infarct size and apoptosis compared to wild type [43,50]. In
support of this observation, vanadyl sulfate treatment caused
decreased expression of FasL which correlated with reduced
caspase-3 activation and cell death in an in vivo model of
I/R, suggesting that the Fas pathway plays a key role in
I/R-mediated apoptosis [53]. The reasons for the differences
in the findings in these studies using the same mice are not
clear and further studies are needed to clarify the role of Fas
in acute I/R injury. In addition, monocyte chemoattractant
protein-1 (MCP-1) plays a crucial role in initiating coronary
heart disease by recruiting monocytes/macrophages to the
vessel wall, and transgenic mice overexpressing MCP-1 in
the heart manifest cardiac inflammation and develop heart
failure. Interestingly, inhibition of FasL function through
cardiac-specific expression of soluble Fas rescued the MCP-1
transgenic mice from developing heart failure, suggesting
that the FasL derived from the infiltrating mononuclear cells
causes death of cardiac cells resulting in the development of
heart failure [54]. Clearly, Fas-mediated apoptosis plays a
role in contributing to heart failure but further studies are
needed to elucidate exactly how and under what conditions
(chronic or acute) Fas activates apoptosis.
Mitochondrial death pathway in heart disease
The primary function of mitochondria is to provide energy
for the cell in the form of ATP through oxidative
phosphorylation. In cardiac myocytes, mitochondria con-
stitute about 30% of cell volume and are located below the
sarcolemma as well as in intermyofibrillar spaces. This
ubiquitous presence and strategic location of mitochondria
ensures efficient ATP supply and delivery to the continu-
ously contracting myocyte. However, mitochondria can
also contribute to cell death in response to intracellular
stress such as increased oxidative stress, serum deprivation,
and DNA damage. In response to stress, mitochondria
release several pro-apoptotic factors such as cytochrome c,
apoptosis inducing factor (AIF), endonuclease G (Endo G),
second mitochondria-derived activator of caspases (Smac/
Diablo), and HtrA2/Omi, resulting in initiation of apopto-
sis. The mitochondrial death pathway appears to play a
significant role in contributing to cell death particularly in
I/R injury. Studies have reported that there is substantial
release of cytochrome cand activation of caspase-9 after
myocardial I/R injury [5557]. Also, it has been reported
that protecting mitochondrial integrity protects against I/R
injury [56,5860].
The mitochondrial cell death pathway is regulated by
the pro- and anti-apoptotic Bcl-2 proteins. These proteins
share up to four conserved Bcl-2 homology (BH) domains.
Anti-apoptotic Bcl-2 proteins, such as Bcl-2 and Bcl-X
L
,
contain all four subtypes of BH domains (BH1-4) and
promote cell survival. The pro-apoptotic Bcl-2 proteins
contain one or three BH domains and therefore are divided
into two structurally distinctive subfamilies: (1) multido-
main proteins such as Bax and Bak that share three BH
regions (BH1-3), and (2) BH3-only domain proteins such
as Bnip3, Nix/Bnip3L, Bad, Bid, Noxa, and Puma [61,62].
The anti-apoptotic members such as Bcl-2 and Bcl-X
L
are important in cell survival and protect cardiac myocytes
against various stressors. For instance, overexpression of
Bcl-X
L
in H9c2 cardiac cells protected against doxorubi-
cin- and hypoxia-mediated apoptosis by preserving
mitochondrial integrity [63], and Bcl-2 was shown to pre-
vent p53-mediated apoptosis in cardiac myocytes [64].
Moreover, transgenic mice showing accelerated accumu-
lation of mitochondrial DNA (mtDNA) mutations due to
expression of an error-prone mtDNA polymerase specifi-
cally in the heart activated a strong prosurvival response by
upregulation of Bcl-2 and Bcl-X
L
[65]. Transgenic mice
overexpressing Bcl-2 in the heart have reduced I/R injury
and fewer apoptotic cells compared to wild type mice,
suggesting that the cardioprotective effect of Bcl-2 is via
inhibition of the mitochondrial death pathway [60,66,67].
Bcl-2 has been reported to protect against cell death by
preventing permeabilization of the outer mitochondrial
membrane by inhibiting activation of pro-apoptotic Bax/
Bak [68]. For instance, mice null for desmin, a muscle-
specific member of the intermediate filament gene family,
develop cardiomyopathy characterized by extensive
Apoptosis (2009) 14:536–548 539
123
cardiomyocyte death, and subsequent heart failure [69].
Weisleder et al. [70] found that overexpression of Bcl-2 in
the desmin null heart attenuated the cardiomyopathy phe-
notype by preventing mitochondrial defects. However, new
protective functions for Bcl-2 have also been identified.
During ischemia, the sodium calcium exchanger (NCX)
operates in reverse to extrude Na
?
and as a result the cell
becomes loaded with Ca
2?
[71]. Interestingly, Bcl-2 was
reported to reduce NCX activity as well as increase resis-
tance to permeability transition in the mitochondria by
increasing the Ca
2?
threshold for mitochondrial perme-
ability transition pore opening in heart mitochondria [72].
Bcl-2 also provides protection by inhibiting consumption
of glycolytically generated ATP by the ATPase. During
myocardial ischemia, mitochondrial ATP generation is
inhibited and the F1F0-ATPase is running in reverse,
thereby consuming ATP generated by glycolysis [73].
Interestingly, Imahashi et al. [74] found that Bcl-2 over-
expression in the heart reduced the rate of ATP decline and
decreased acidosis during ischemia. These studies suggest
that the anti-apoptotic proteins can protect against cell
death via multiple mechanisms and have attractive thera-
peutic potential to treat or prevent heart disease.
The pro-apoptotic Bcl-2 proteins activate cell death by
permeabilizing the outer mitochondrial membrane, which
releases pro-apoptotic proteins such as cytochrome cand
AIF from the intermembrane space. The pro-apoptotic BH3-
only proteins function as cellular stress sensors and integrate
diverse cell death stimuli. For instance, Bnip3 is activated in
response to increased oxidative stress [75], whereas Bad is
activated in response to growth factor deprivation [76].
Activation of a BH3-only protein in response to a specific
stress induces activation of downstream effectors Bax and/
or Bak either by directly interacting with these proteins or
indirectly by sequestering anti-apoptotic Bcl-2/Bcl-X
L
[68].
Activation of Bax and/or Bak induce their oligomerization
in the mitochondrial membrane and formation of a pore in
the outer mitochondrial membrane which allows for release
of proteins from the intermembrane space [68]. Bax and Bak
are essential for cell death mediated via the mitochondrial
pathway, and mouse embryonic fibroblast (MEFs) isolated
from the Bax/Bak double knockout mouse are completely
resistant to cell death by staurosporine, growth factor
deprivation, and UV, as well as to overexpression of BH3-
only proteins such a tBid, Bad, and Bnip3 [75,77,79]. The
pro-apoptotic Bcl-2 proteins have been widely implicated in
cardiovascular disease. In the myocardial cells, Bax has
been reported to be activated in response to oxidative stress
[56] and simulated I/R [78,80,81]. Moreover, hearts of Bax
deficient mice have reduced mitochondrial damage and
infarcts size compared to wild type mice, suggesting that
Bax is an important contributor to I/R injury [82]. The most
studied BH3-only protein in the heart is Bnip3 which has
been reported to contribute to acute I/R injury [58] and post-
infarct remodeling [83]. Bnip3 is transcriptionally upregu-
lated by hypoxia in neonatal cardiac myocytes via HIF-1a
[84] and E2F-1 [85], and Frazier et al. [86] have reported
that the Bnip3 protein is stabilized by acidosis. Recently, we
reported that Bnip3 is activated by oxidative stress which
promotes homodimerization and activation of Bnip3 in the
myocardium [75]. Since apoptosis occurs mainly during
reperfusion, these studies suggest that oxygen deprivation
and acidosis promotes upregulation and stabilization of
Bnip3, respectively, and that increased oxidative stress
during reperfusion promotes homodimerization and acti-
vation of Bnip3 in the mitochondria. Bnip3 has been
reported to activate downstream Bax/Bak [75] as well as
induce opening of the mitochondrial permeability transition
pore [87,88]. In contrast, the Bnip3 homologue Nix/Bnip3L
has been implicated in cardiac hypertrophy and develop-
ment of cardiomyopathy [89]. Moreover, Bid is cleaved to
its truncated and active form in response to I/R [55,90], and
Bad is upregulated in response to I/R [91]. Since ischemia
and reperfusion induces multiple stress signals (oxygen
deprivation, oxidative stress, DNA damage, growth factor
deprivation etc.) in the cardiac myocyte, multiple BH3-only
proteins are activated in response to I/R ensuring activation
of the cell death program. Thus, the BH3-only proteins are
important activators of the mitochondrial cell death pathway
in response to myocardial infarction and pathological car-
diac hypertrophy.
Importantly, the relative level of the pro- and anti-
apoptotic Bcl-2 proteins determines whether a cell will
survive or die following an apoptotic stimulus. It has been
reported that there is a shift in the ratio of anti- to pro-
apoptotic Bcl-2 proteins to proteins that promote apoptosis
in the hearts of patients after a myocardial infarction, in
severe dilated cardiomyopathy and ischemic heart disease
[9294]. In addition, there was a significant shift in the
ratio of Bax to Bcl-X
L
during the transition from com-
pensated hypertrophy to heart failure in a model of pressure
overload which correlated with increased levels of cyto-
chrome cin the cytosol and caspase-3 activation [95].
Although the ratio is shifted towards anti-apoptotic pro-
teins, it may not activate apoptosis. However, the myocytes
will have a lower threshold for additional stress and will
induce apoptosis easier than a normal cell, making these
hearts more susceptible to develop heart failure.
Permeabilization of the mitochondrial membrane
Cytochrome c
In healthy cells, the role of cytochrome cis to shuttle
electrons from complex III to IV of the respiratory chain.
540 Apoptosis (2009) 14:536–548
123
However, during apoptosis, cytochrome cis released into
the cytosol where it binds to Apaf-1 and dATP which
induces a conformational change in Apaf-1 which recruits
and subsequently activates caspases-9. The activated cas-
pase-9 cleaves pro-caspases-3 to produce active effector
caspase-3 which leads to the cleavage of cellular target
proteins [9698]. In addition, it has been reported that
activated caspase-3 translocates to nucleus, then cleaves
the DNA repairing enzyme, poly (ADP-ribose) polymerase
(PARP) and activates endonucleases which cleave DNA.
These events culminate in apoptotic cell death [96,99,
100].
Apoptosis-inducing factor (AIF)
AIF is a flavoprotein localized in the mitochondrial inter-
membrane space and is required for oxidative
phosphorylation and for the assembly and/or stabilization
of respiratory complex I [101]. AIF is essential for the
maintenance of normal heart function, and inactivation of
AIF results in dilated cardiomyopathy [102]. In addition,
cardiac myocytes from Harlequin (Hq) mice, which has
*80% reduction in AIF protein levels due to a proviral
insertion in the first intron of the Aif gene, are sensitized to
oxidative stress-induced cell death, and Hq hearts display
more severe ischemic damage compared to wild-type
hearts after acute ischemia/reperfusion injury [103]. AIF is
anchored to the mitochondrial inner membrane via its
N-terminus. However, upon induction of apoptosis, AIF is
cleaved and released into the cytosol [104] where it then
translocates to the nucleus and mediates chromatin con-
densation and large-scale DNA fragmentation [105].
Interestingly, it was recently reported that increased levels
of mitochondrial Ca
2?
activated a mitochondrial calpain
which cleaved membrane bound AIF. Permeabilization of
the mitochondrial membrane resulted in release of the
soluble AIF [104]. It is well known that I/R causes Ca
2?
overload of mitochondria [106], and reduction of mito-
chondrial Ca
2?
uptake has beneficial effects on cardiac
function following I/R with improved functional recovery
and reduced infarct size [107,108]. Since it has been
reported that I/R induces release of AIF [109,110], it will
be interesting to investigate whether this is due to pro-
cessing and release of AIF by Ca
2?
-activated Calpain.
Moreover, considering the fact that AIF functions as a
NADH-oxidase involved in electron transport in complex I
of the electron transport chain, loss of AIF from mito-
chondria may cause further mitochondrial dysfunction due
to reduction in electron transport capacity. Thus, release of
AIF not only initiates apoptosis, but also subsequently
induces mitochondrial dysfunction which will ensure cell
death.
Omi/HtrA2, Smac/Diablo, and endonuclease G
(EndoG)
Other apoptotic effectors are released from mitochondria,
including the serine protease Omi/HtrA2, Smac/Diablo,
and Endo G. Similar to cytochrome c-mediated activation
of the caspase cascade, Smac/Diablo and Omi/HtrA2 are
indirectly involved in caspase activation. For example,
inhibitors of apoptosis proteins (IAPs) are endogenous
inhibitors of caspases and are bound to caspase-3 and -9
under normal conditions [111,112]. However, when
released to cytosol, Smac/Diablo and Omi/HtrA2 prevent
IAPs from inhibiting the caspases, resulting in activation of
caspase-9 and -3. A role for Omi/HtrA2 in the heart was
demonstrated by two studies where inhibition of Omi/
HtrA2 using ucf-101 reduced apoptosis and infarct size in
mice as well as rat after in vivo I/R [113,114]. Also,
transgenic mice overexpressing IAP2 had reduced infarct
size and fewer TUNEL-positive cells after I/R [115].
Endo G is a nuclear-encoded endonuclease and localized
in the inner membrane space of mitochondria. Under nor-
mal conditions, Endo G plays a role in maintenance of
mitochondria DNA by removing defective DNA [116].
However, similar to AIF once released, Endo G translo-
cates to the nucleus where it cleaves DNA. Although few
studies using Endo G-/-splenocytes and fibroblasts
report that Endo G is not essential for nuclear DNA frag-
mentation and apoptosis, studies using intact heart or
isolated cardiomyocytes demonstrate that Endo G plays a
role in I/R-mediated cell death. For instance, Javadov et al.
[117] reported that VAE-480, a specific Na
?
/H
?
exchan-
ger-1 (NHE-1) inhibitor, reduced release of Endo G and I/R
injury, suggesting that attenuation of Endo G release is
cardioprotective. In addition, Bahi et al. [118] demon-
strated that Endo G was released from mitochondria and
induced DNA damage, and importantly when Endo G was
downregulated using siRNA, DNA damage was signifi-
cantly reduced in adult cardiomyocytes during I/R.
Collectively, these studies suggest that Endo G is a critical
contributor to DNA degradation during I/R.
Activation of the ER stress death pathway in the heart
The ER is responsible for synthesis and folding of secreted
proteins as well as Ca
2?
storage. Under normal conditions,
there is a balance between import of newly unfolded pro-
teins into the ER and secretion of folded mature proteins.
When this balance is disturbed and there is an accumula-
tion of protein aggregates in the ER lumen, the unfolded
protein response (UPR) is activated in an attempt to restore
ER homeostasis [119]. Aberrant Ca
2?
regulation can also
activate the UPR. The UPR activates a transcriptional
Apoptosis (2009) 14:536–548 541
123
program to increase the protein folding capacity of the ER,
and the degradation of misfolded proteins, as well as sup-
press proteins synthesis in the cell [120]. Thus, the UPR is
primarily an adaptive response to restore ER homeostasis
and protect the cell from stress. However, if the stress is
prolonged and overwhelming, the UPR activates a pro-
apoptotic response instead which is independent of the
mitochondria.
Several recent studies have demonstrated a role for ER
stress in cardiovascular disease. For instance, Okada et al.
[121] found that pressure overload by aortic constriction
induced extensive ER stress during progression from car-
diac hypertrophy to heart failure. In addition, studies have
reported that ER stress occurs in response to myocardial
I/R [122,123], and that enhanced ER stress is associated
with the development of ischemic heart disease [124].
There is also evidence that ER stress-induced apoptosis is
involved in pathogenesis of diabetes and heart failure [26],
and ultrastructural analysis revealed swelling of the ER in
diabetic myocardium [125]. Consistent with these obser-
vations, defective ER quality control in transgenic mice
with mutant KDEL receptor (a receptor for ER chaperones)
caused dilated cardiomyopathy [126]. These studies sug-
gest that apoptosis mediated by ER stress may be a
significant contributor to cardiovascular disease.
Endoplasmic reticulum stress induces cell death via two
different mechanisms. Under ER stress, activated caspase-
12 activates caspase-3, leading to apoptosis [119]. The
second death-signaling pathway activated by ER stress is
activation of a transcriptional program via upregulation of
the transcription factor CHOP/GADD 153. CHOP activates
transcription of genes encoding pro-apoptotic proteins
including the BH3-only protein Puma [127]. Szegezdi et al.
[122] found that during prolonged ischemia, the UPR
switched to a pro-apoptotic response by activation of the
ER stress death pathway with upregulation of the tran-
scription factor CHOP, and activation of caspase-12 in
neonatal cardiac myocytes. Studies in cardiac myocytes
and the heart have demonstrated that increased CHOP
expression during I/R contributed towards apoptosis, and
inhibition of CHOP expression using a PKCdinhibitor
significantly attenuated ER-mediated apoptosis [128]. This
study suggests that translocation of PKCdto ER membrane
induced the UPR, which led to upregulation of CHOP and
subsequent apoptosis. In addition, Okada et al. [121]
demonstrated that pharmacological intervention inducing
ER stress increased CHOP expression and apoptosis in the
heart.
Interestingly, the BH3-only protein Puma was reported
to be upregulated in isolated cardiac myocytes exposed to
hypoxia/re-oxygensation, and genetic deletion of Puma
reduced I/R-mediated cell death and decreased infarct size
in Langendorff perfused hearts [129]. Recently, it was
demonstrated that myocytes deficient in Puma or down-
regulation via siRNA were resistant to ER stress-induced
apoptosis [130], suggesting that Puma is a critical com-
ponent of ER-stress induced apoptosis in cardiac myocytes.
The Bcl-2 proteins have been shown to localize to the ER
where they can regulate the levels of Ca
2?
stored in the ER
[131].
Cross talk between the apoptotic pathways
There is cross talk between the death receptor and mito-
chondrial cell death pathways. Activation of caspase 8 by
the death receptor pathway directly activates executioner
caspase-3, but also cleaves the BH3-only protein Bid at
aspartate 60 to generate a 15 kDa truncated form (tBid)
that facilitates release of cytochrome cfrom the mito-
chondria [132,133]. The release of cytochrome ccauses
activation of caspase-9 which in turn also activates cas-
pase-3, thus amplifying the death signal from the plasma
membrane. The cross talk between the pathways via death
receptor activation has been demonstrated in cardiac
myocytes and the heart. For instance, Date et al. [134]
found that overexpression of FasL activated both caspase-8
and -9 in neonatal cardiac myocytes. Moreover, cardiac
restricted overexpression of TNF-aresults in increased
apoptosis and development of heart failure, but when these
mice were crossed with transgenic mice overexpressing
Bcl-2 in the heart cardiac myocytes apoptosis and LV
remodeling was attenuated [135]. In addition, caspase-8
was activated, and Bid was cleaved to t-Bid, suggesting
concurrent activation of both the death receptor and
mitochondrial death pathways in the heart by TNF-a.In
support of this, Bcl-2 overexpression only partially atten-
uated cardiomyocyte apoptosis and had no effect on
extrinsic signaling.
There have also been reports of cross-talk between the
death receptor pathway and the ER-stress pathway. For
instance, treatment of L929 cells with FasL induced pro-
cessing of caspase-12 as well as activation of caspases-3,
-7, and -9 [136]. In addition, Bajaj and Sharma [137]
recently reported that TNF-atreatment caused activation of
both caspase-3 and -12 in the HL-1 myocyte cell line.
There is also evidence that there is communication between
the ER and the mitochondrial death pathways. For instance,
ER-targeted Bcl-2 was reported to inhibit mitochondrial
membrane depolarization and cytochrome crelease in
apoptotic myelodysplastic syndrome erythroid precursor
cells [138]. BAP31 is an ER-associated protein that is
cleaved by caspase-8. The cleaved fragment directs pro-
apoptotic signals between the ER and mitochondria by
inducing release of Ca
2?
from the ER which gets taken up
by the mitochondria [139]. In addition, the BH3-only
542 Apoptosis (2009) 14:536–548
123
protein Bik has been reported to localize to the mito-
chondria where it initiates release of Ca
2?
via activation of
Bax/Bak in the ER membrane [140]. These studies suggest
that the ER communicates with the mitochondria by
releasing Ca
2?
.
Autophagy as a mediator of cell death
Autophagy is an important cellular process involved in
recycling of long-lived proteins and organelles. In the
heart, autophagy is important to maintain homeostasis and
disruption or a defect in this pathway leads to ventricular
dysfunction and heat failure [141]. It is also an important
survival response which is upregulated during starvation
when the cells need to recycle amino acid and fatty acids.
In the heart, autophagy is upregulated in response to
ischemia/reperfusion [58,142], and pressure overload
[143]. However, the functional significance of increased
autophagy in the heart is not clear and autophagy has been
reported to protect against cell death as well be the cause of
cell death [81,142,144].
Interestingly, recent studies suggest that there is cross
talk between autophagy and apoptosis. Atg5, an essential
autophagy protein, has been reported to activate apoptosis.
Yousefi et al. [145] found that overexpression of Atg5
increased the cell’s susceptibility to apoptosis following
stimulation with several death triggers, including antican-
cer drugs. They found that the death stimulation resulted in
calpain-mediated cleavage of Atg5, and truncated Atg5
induced cytochrome crelease and apoptosis. Overexpression
of Bcl-2 protected against Atg5-mediated mitochondrial
dysfunction. This suggests that Atg5 can serve as a molecular
switch between autophagy and apoptosis, but it is not clear
how truncated Atg5 triggers mitochondrial permeabiliza-
tion. Another study reported that Atg5 associated with the
Fas-associated death domain (FADD) protein to mediate
IFN-c-induced cell death. In addition, a mutant of Atg5
(Atg5K130R), which was unable to activate autophagy,
was able to induce cell death [146]. Similarly, calpain-
mediated truncated Atg5 was unable to promote autophagy,
but induced apoptosis [145]. These two studies suggest that
Atg5-mediated apoptosis does not require the formation of
autophagosomes. Further investigations of this molecular
link between autophagy and apoptosis are needed.
Moreover, anti-apoptotic Bcl-2 family members and pro-
apoptotic BH3-only proteins may participate in the inhibi-
tion and induction of autophagy, respectively. Beclin 1, an
essential autophagy protein, is regulated by the Bcl-2 pro-
teins. Under normal conditions, Bcl-2 and Bcl-X
L
suppress
autophagy by associating with Beclin 1 through a BH3
domain in Beclin 1 and the BH3 binding groove of Bcl-2/
Bcl-X
L
[147]. The BH3-only proteins can disrupt the
interaction between Beclin 1 and Bcl-2/Bcl-X
L
to induce
autophagy [148]. The BH3-only protein Bik was found to
cause enhanced cell death with autophagic features in
Bcl-2 deficient cells [149], and downregulation of Bcl-2
using siRNA induced autophagic cell death in MCF-7 cells
[150]. This suggests that with a lack of Bcl-2, there is no
break on Beclin-1 which can now be activated and induce
excess autophagy. The BH3-only protein Bnip3 has been
shown to be a potent inducer of autophagy in cardiac cells
[58]. Bnip3 has been shown to interact with Bcl-2 and Bcl-
X
L
[151], but it is not known whether Bnip3 disrupts the
interaction between Beclin 1 and Bcl-2/Bcl-X
L
. These
studies suggest that there is cross talk between autophagy
and apoptosis via Beclin 1 and Atg5. It is interesting that
Beclin 1 contains a BH3-only domain, but it is not known
whether it can act as a BH3-only protein and activate
apoptosis. However, Beclin 1 heterozygous mice (Beclin
1
?/-
), which have reduced autophagy, also have reduced
apoptosis and infarcts size after I/R injury [142], suggesting
that Beclin 1 might activate apoptosis.
Conclusion and perspective
It is clear that apoptosis plays a critical role in pathogenesis
of various cardiovascular diseases. Currently, three apop-
totic pathways (death receptor, mitochondrial, and ER-
stress) have been identified in the heart to contribute to
myocyte loss in various cardiovascular diseases. Since the
regenerative capacity of the myocardium is limited, there is
intense interest in the prevention of cardiomyocyte loss in
cardiovascular diseases to prevent development of heart
failure. Since apoptosis is a highly regulated process, it is a
good potential target for therapeutic intervention. Inhibi-
tion of cardiac myocyte apoptosis using novel research
technology including transgenic mice, gene deletion,
recombinant proteins, and pharmacological inhibitors
results in cardioprotection and prevention of heart failure.
However, many important questions regarding the effect of
each anti-apoptotic intervention remain to be answered. For
instance, loss of mitochondrial membrane integrity is
generally considered a point of no return in the cell death
process. Therefore, inhibiting downstream effectors such as
caspase-3 will delay, but not prevent cell death. Thus, it is
essential to preserve mitochondrial integrity for effective
therapy. Another complication is that multiple pro-apop-
totic proteins can be activated in response to injury. For
instance, multiple BH3-only proteins are activated during
I/R, each of which would have to be targeted for effective
therapy. In addition, it is not clear how much cross-talk
there is between necrosis, apoptosis, and autophagy in the
heart. Enhanced autophagy is often seen in cells where
apoptosis is inhibited, and it is unknown whether inhibition
Apoptosis (2009) 14:536–548 543
123
of apoptosis for cardioprotection against various heart
diseases will subsequently lead to non-apoptotic cell death
later. Although several potential therapeutic agents have
been tested in animal models of I/R injury with success,
nearly none of the specific anti-apoptotic agents have
reached the stage of clinical research. Thus, a better
understanding of the complex mechanisms associated with
cardiac myocyte apoptosis is necessary to identify potential
targets and to develop novel therapeutic strategies for
cardiovascular diseases.
Acknowledgments This manuscript was supported by a Scientist
Development Award from AHA, and NIH grant HL087023 to A
˚.B.G.
References
1. Reggiori F, Klionsky DJ (2005) Autophagosomes: biogenesis
from scratch? Curr Opin Cell Biol 17:415–422. doi:10.1016/
j.ceb.2005.06.007
2. Kitsis RN, Peng CF, Cuervo AM (2007) Eat your heart out. Nat
Med 13:539–541. doi:10.1038/nm0507-539
3. Schweichel JU, Merker HJ (1973) The morphology of various
types of cell death in prenatal tissues. Teratology 7:253–266.
doi:10.1002/tera.1420070306
4. Searle J, Kerr JF, Bishop CJ (1982) Necrosis and apoptosis:
distinct modes of cell death with fundamentally different sig-
nificance. Pathol Annu 17(Pt 2):229–259
5. Kerr JF (1971) Shrinkage necrosis: a distinct mode of cellular
death. J Pathol 105:13–20. doi:10.1002/path.1711050103
6. Majno G, Joris I (1995) Apoptosis, oncosis, and necrosis. An
overview of cell death. Am J Pathol 146:3–15
7. Rich T, Watson CJ, Wyllie A (1999) Apoptosis: the germs of
death. Nat Cell Biol 1:E69–E71. doi:10.1038/11038
8. Savill J, Fadok V (2000) Corpse clearance defines the meaning
of cell death. Nature 407:784–788. doi:10.1038/35037722
9. Olivetti G, Quaini F, Sala R et al (1996) Acute myocardial
infarction in humans is associated with activation of programmed
myocyte cell death in the surviving portion of the heart. J Mol
Cell Cardiol 28:2005–2016. doi:10.1006/jmcc.1996.0193
10. Saraste A, Pulkki K, Kallajoki M, Henriksen K, Parvinen M,
Voipio-Pulkki LM (1997) Apoptosis in human acute myocardial
infarction. Circulation 95:320–323
11. Narula J, Haider N, Virmani R et al (1996) Apoptosis in myo-
cytes in end-stage heart failure. N Engl J Med 335:1182–1189.
doi:10.1056/NEJM199610173351603
12. Aharinejad S, Andrukhova O, Lucas T et al (2008) Programmed
cell death in idiopathic dilated cardiomyopathy is mediated by
suppression of the apoptosis inhibitor Apollon. Ann Thorac Surg
86:109–114. doi:10.1016/j.athoracsur.2008.03.057 discussion
114
13. Cheng W, Kajstura J, Nitahara JA et al (1996) Programmed
myocyte cell death affects the viable myocardium after infarc-
tion in rats. Exp Cell Res 226:316–327. doi:10.1006/excr.1996.
0232
14. Gottlieb RA, Burleson KO, Kloner RA, Babior BM, Engler RL
(1994) Reperfusion injury induces apoptosis in rabbit cardio-
myocytes. J Clin Invest 94:1621–1628. doi:10.1172/JCI117504
15. Gao HK, Yin Z, Zhou N, Feng XY, Gao F, Wang HC (2008)
Glycogen synthase kinase 3 inhibition protects the heart from
acute ischemia–reperfusion injury via inhibition of inflammation
and apoptosis. J Cardiovasc Pharmacol 52:286–292. doi:
10.1097/FJC.0b013e318186a84d
16. Kubota T, McTiernan CF, Frye CS et al (1997) Dilated cardio-
myopathy in transgenic mice with cardiac-specific overexpression
of tumor necrosis factor-alpha. Circ Res 81:627–635
17. Bryant D, Becker L, Richardson J et al (1998) Cardiac failure in
transgenic mice with myocardial expression of tumor necrosis
factor-alpha. Circulation 97:1375–1381
18. Sayen MR, Gustafsson AB, Sussman MA, Molkentin JD,
Gottlieb RA (2003) Calcineurin transgenic mice have mito-
chondrial dysfunction and elevated superoxide production. Am J
Physiol Cell Physiol 284:C562–C570
19. Wang J, Silva JP, Gustafsson CM, Rustin P, Larsson NG (2001)
Increased in vivo apoptosis in cells lacking mitochondrial DNA
gene expression. Proc Natl Acad Sci USA 98:4038–4043. doi:
10.1073/pnas.061038798
20. Kockx MM, De Meyer GR, Muhring J, Jacob W, Bult H, Her-
man AG (1998) Apoptosis and related proteins in different
stages of human atherosclerotic plaques. Circulation 97:2307–
2315
21. Clarke MC, Littlewood TD, Figg N et al (2008) Chronic
apoptosis of vascular smooth muscle cells accelerates athero-
sclerosis and promotes calcification and medial degeneration.
Circ Res 102:1529–1538. doi:10.1161/CIRCRESAHA.108.
175976
22. Matsumura K, Jeremy RW, Schaper J, Becker LC (1998) Pro-
gression of myocardial necrosis during reperfusion of ischemic
myocardium. Circulation 97:795–804
23. Takemura G, Ohno M, Hayakawa Y et al (1998) Role of
apoptosis in the disappearance of infiltrated and proliferated
interstitial cells after myocardial infarction. Circ Res 82:1130–
1138
24. Kannel WB, Hjortland M, Castelli WP (1974) Role of diabetes
in congestive heart failure: the Framingham study. Am J Cardiol
34:29–34. doi:10.1016/0002-9149(74)90089-7
25. Cai L, Li W, Wang G, Guo L, Jiang Y, Kang YJ (2002)
Hyperglycemia-induced apoptosis in mouse myocardium: mito-
chondrial cytochrome C-mediated caspase-3 activation pathway.
Diabetes 51:1938–1948. doi:10.2337/diabetes.51.6.1938
26. Li Z, Zhang T, Dai H et al (2007) Involvement of endoplasmic
reticulum stress in myocardial apoptosis of streptozocin-induced
diabetic rats. J Clin Biochem Nutr 41:58–67. doi:10.3164/jcbn.
2007008
27. Mukamal KJ, Nesto RW, Cohen MC et al (2001) Impact of
diabetes on long-term survival after acute myocardial infarction:
comparability of risk with prior myocardial infarction. Diabetes
Care 24:1422–1427. doi:10.2337/diacare.24.8.1422
28. Backlund T, Palojoki E, Saraste A et al (2004) Sustained car-
diomyocyte apoptosis and left ventricular remodelling after
myocardial infarction in experimental diabetes. Diabetologia
47:325–330. doi:10.1007/s00125-003-1311-5
29. Kuethe F, Sigusch HH, Bornstein SR, Hilbig K, Kamvissi V,
Figulla HR (2007) Apoptosis in patients with dilated cardio-
myopathy and diabetes: a feature of diabetic cardiomyopathy?
Horm Metab Res 39:672–676. doi:10.1055/s-2007-985823
30. Cai L, Kang YJ (2001) Oxidative stress and diabetic cardio-
myopathy: a brief review. Cardiovasc Toxicol 1:181–193. doi:
10.1385/CT:1:3:181
31. Song Y, Wang J, Li Y et al (2005) Cardiac metallothionein
synthesis in streptozotocin-induced diabetic mice, and its pro-
tection against diabetes-induced cardiac injury. Am J Pathol
167:17–26
32. Ye G, Metreveli NS, Ren J, Epstein PN (2003) Metallothionein
prevents diabetes-induced deficits in cardiomyocytes by inhib-
iting reactive oxygen species production. Diabetes 52:777–783.
doi:10.2337/diabetes.52.3.777
33. Guerra S, Leri A, Wang X et al (1999) Myocyte death in the
failing human heart is gender dependent. Circ Res 85:856–866
544 Apoptosis (2009) 14:536–548
123
34. Wencker D, Chandra M, Nguyen K et al (2003) A mechanistic
role for cardiac myocyte apoptosis in heart failure. J Clin Invest
111:1497–1504
35. Thorburn A (2004) Death receptor-induced cell killing. Cell
Signal 16:139–144. doi:10.1016/j.cellsig.2003.08.007
36. Liao X, Wang X, Gu Y, Chen Q, Chen LY (2005) Involvement
of death receptor signaling in mechanical stretch-induced car-
diomyocyte apoptosis. Life Sci 77:160–174. doi:10.1016/j.lfs.
2004.11.029
37. Levine B, Kalman J, Mayer L, Fillit HM, Packer M (1990)
Elevated circulating levels of tumor necrosis factor in severe
chronic heart failure. N Engl J Med 323:236–241
38. Dutka DP, Elborn JS, Delamere F, Shale DJ, Morris GK (1993)
Tumour necrosis factor alpha in severe congestive cardiac fail-
ure. Br Heart J 70:141–143. doi:10.1136/hrt.70.2.141
39. Ferrari R, Bachetti T, Confortini R et al (1995) Tumor necrosis
factor soluble receptors in patients with various degrees of
congestive heart failure. Circulation 92:1479–1486
40. Testa M, Yeh M, Lee P et al (1996) Circulating levels of
cytokines and their endogenous modulators in patients with mild
to severe congestive heart failure due to coronary artery disease
or hypertension. J Am Coll Cardiol 28:964–971. doi:10.1016/
S0735-1097(96)00268-9
41. Giroir BP, Johnson JH, Brown T, Allen GL, Beutler B (1992)
The tissue distribution of tumor necrosis factor biosynthesis
during endotoxemia. J Clin Invest 90:693–698. doi:10.1172/
JCI115939
42. Kapadia S, Lee J, Torre-Amione G, Birdsall HH, Ma TS, Mann
DL (1995) Tumor necrosis factor-alpha gene and protein
expression in adult feline myocardium after endotoxin admin-
istration. J Clin Invest 96:1042–1052. doi:10.1172/JCI118090
43. Jeremias I, Kupatt C, Martin-Villalba A et al (2000) Involve-
ment of CD95/Apo1/Fas in cell death after myocardial ischemia.
Circulation 102:915–920
44. Torre-Amione G, Kapadia S, Lee J et al (1996) Tumor necrosis
factor-alpha and tumor necrosis factor receptors in the failing
human heart. Circulation 93:704–711
45. Doyama K, Fujiwara H, Fukumoto M et al (1996) Tumour
necrosis factor is expressed in cardiac tissues of patients with
heart failure. Int J Cardiol 54:217–225. doi:10.1016/0167-5273
(96)02607-1
46. Gilles S, Zahler S, Welsch U, Sommerhoff CP, Becker BF
(2003) Release of TNF-alpha during myocardial reperfusion
depends on oxidative stress and is prevented by mast cell sta-
bilizers. Cardiovasc Res 60:608–616. doi:10.1016/j.cardiores.
2003.08.016
47. Tanaka M, Ito H, Adachi S et al (1994) Hypoxia induces apop-
tosis with enhanced expression of Fas antigen messenger RNA in
cultured neonatal rat cardiomyocytes. Circ Res 75:426–433
48. Meldrum DR (1998) Tumor necrosis factor in the heart. Am J
Physiol 274:R577–R595
49. Seko Y, Kayagaki N, Seino K, Yagita H, Okumura K, Nagai R
(2002) Role of Fas/FasL pathway in the activation of infiltrating
cells in murine acute myocarditis caused by Coxsackievirus B3.
J Am Coll Cardiol 39:1399–1403. doi:10.1016/S0735-1097(02)
01776-X
50. Lee P, Sata M, Lefer DJ, Factor SM, Walsh K, Kitsis RN (2003)
Fas pathway is a critical mediator of cardiac myocyte death and
MI during ischemia–reperfusion in vivo. Am J Physiol Heart
Circ Physiol 284:H456–H463
51. Li Y, Takemura G, Kosai K et al (2004) Critical roles for the
Fas/Fas ligand system in postinfarction ventricular remodeling
and heart failure. Circ Res 95:627–636. doi:10.1161/01.RES.000
0141528.54850.bd
52. Gomez L, Chavanis N, Argaud L et al (2005) Fas-independent
mitochondrial damage triggers cardiomyocyte death after
ischemia–reperfusion. Am J Physiol Heart Circ Physiol
289:H2153–H2158. doi:10.1152/ajpheart.00165.2005
53. Bhuiyan MS, Takada Y, Shioda N, Moriguchi S, Kasahara J,
Fukunaga K (2008) Cardioprotective effect of vanadyl sulfate on
ischemia/reperfusion-induced injury in rat heart in vivo is
mediated by activation of protein kinase B and induction of
FLICE-inhibitory protein. Cardiovasc Ther 26:10–23
54. Niu J, Azfer A, Deucher MF, Goldschmidt-Clermont PJ, Kol-
attukudy PE (2006) Targeted cardiac expression of soluble FAS
prevents the development of heart failure in mice with cardiac-
specific expression of MCP-1. J Mol Cell Cardiol 40:810–820.
doi:10.1016/j.yjmcc.2006.03.010
55. Scarabelli TM, Stephanou A, Pasini E et al (2002) Different
signaling pathways induce apoptosis in endothelial cells and
cardiac myocytes during ischemia/reperfusion injury. Circ Res
90:745–748. doi:10.1161/01.RES.0000015224.07870.9A
56. Gustafsson AB, Tsai JG, Logue SE, Crow MT, Gottlieb RA
(2004) Apoptosis repressor with caspase recruitment domain
protects against cell death by interfering with Bax activation.
J Biol Chem 279:21233–21238. doi:10.1074/jbc.M400695200
57. Chen M, Won DJ, Krajewski S, Gottlieb RA (2002) Calpain and
mitochondria in ischemia/reperfusion injury. J Biol Chem
277:29181–29186. doi:10.1074/jbc.M204951200
58. Hamacher-Brady A, Brady NR, Logue SE et al (2007) Response
to myocardial ischemia/reperfusion injury involves Bnip3 and
autophagy. Cell Death Differ 14:146–157. doi:10.1038/sj.cdd.
4401936
59. Iliodromitis EK, Lazou A, Kremastinos DT (2007) Ischemic
preconditioning: protection against myocardial necrosis and
apoptosis. Vasc Health Risk Manag 3:629–637
60. Imahashi K, Schneider MD, Steenbergen C, Murphy E (2004)
Transgenic expression of Bcl-2 modulates energy metabolism,
prevents cytosolic acidification during ischemia, and reduces
ischemia/reperfusion injury. Circ Res 95:734–741. doi:10.1161/
01.RES.0000143898.67182.4c
61. Danial NN, Korsmeyer SJ (2004) Cell death: critical control
points. Cell 116:205–219. doi:10.1016/S0092-8674(04)00046-7
62. Huang DC, Strasser A (2000) BH3-Only proteins-essential initi-
ators of apoptotic cell death. Cell 103:839–842. doi:10.1016/
S0092-8674(00)00187-2
63. Reeve JL, Szegezdi E, Logue SE et al (2007) Distinct mechanisms
of cardiomyocyte apoptosis induced by doxorubicin and hypoxia
converge on mitochondria and are inhibited by Bcl-xL. J Cell Mol
Med 11:509–520. doi:10.1111/j.1582-4934.2007.00042.x
64. Kirshenbaum LA, de Moissac D (1997) The bcl-2 gene product
prevents programmed cell death of ventricular myocytes. Cir-
culation 96:1580–1585
65. Zhang D, Mott JL, Chang SW, Stevens M, Mikolajczak P,
Zassenhaus HP (2005) Mitochondrial DNA mutations activate
programmed cell survival in the mouse heart. Am J Physiol
Heart Circ Physiol 288:H2476–H2483. doi:10.1152/ajpheart.
00670.2004
66. Brocheriou V, Hagege AA, Oubenaissa A et al (2000) Cardiac
functional improvement by a human Bcl-2 transgene in a mouse
model of ischemia/reperfusion injury. J Gene Med 2:326–333.
doi:10.1002/1521-2254(200009/10)2:5\326::AID-JGM133[3.0.
CO;2-1
67. Chen Z, Chua CC, Ho YS, Hamdy RC, Chua BH (2001)
Overexpression of Bcl-2 attenuates apoptosis and protects
against myocardial I/R injury in transgenic mice. Am J Physiol
Heart Circ Physiol 280:H2313–H2320
68. Gustafsson AB, Gottlieb RA (2007) Bcl-2 family members and
apoptosis, taken to heart. Am J Physiol Cell Physiol 292:C45–
C51
69. Milner DJ, Taffet GE, Wang X et al (1999) The absence of
desmin leads to cardiomyocyte hypertrophy and cardiac dilation
Apoptosis (2009) 14:536–548 545
123
with compromised systolic function. J Mol Cell Cardiol
31:2063–2076. doi:10.1006/jmcc.1999.1037
70. Weisleder N, Taffet GE, Capetanaki Y (2004) Bcl-2 overex-
pression corrects mitochondrial defects and ameliorates
inherited desmin null cardiomyopathy. Proc Natl Acad Sci USA
101:769–774. doi:10.1073/pnas.0303202101
71. Karmazyn M, Moffat MP (1993) Role of Na?/H?exchange in
cardiac physiology and pathophysiology: mediation of myo-
cardial reperfusion injury by the pH paradox. Cardiovasc Res
27:915–924. doi:10.1093/cvr/27.6.915
72. Zhu L, Yu Y, Chua BH, Ho YS, Kuo TH (2001) Regulation of
sodium–calcium exchange and mitochondrial energetics by Bcl-
2 in the heart of transgenic mice. J Mol Cell Cardiol 33:2135–
2144. doi:10.1006/jmcc.2001.1476
73. Rouslin W, Erickson JL, Solaro RJ (1986) Effects of oligomycin
and acidosis on rates of ATP depletion in ischemic heart muscle.
Am J Physiol 250:H503–H508
74. Imahashi K, Schneider M, Steenbergen C, Murphy E (2004)
BCL-2 reduces acidification and the fall in ATP during
Ischaemia. Cardiovasc J S Afr 15:S3
75. Kubli DA, Quinsay MN, Huang C, Lee Y, Gustafsson AB
(2008) Bnip3 functions as a mitochondrial sensor of oxidative
stress during myocardial ischemia and reperfusion. Am J
Physiol Heart Circ Physiol 295(5):H2025–H2031
76. Zha J, Harada H, Yang E, Jockel J, Korsmeyer SJ (1996) Serine
phosphorylation of death agonist BAD in response to survival
factor results in binding to 14–3-3 not BCL-X(L). Cell 87:619–
628. doi:10.1016/S0092-8674(00)81382-3
77. Wei MC, Zong WX, Cheng EH et al (2001) Proapoptotic BAX
and BAK: a requisite gateway to mitochondrial dysfunction and
death. Science 292:727–730. doi:10.1126/science.1059108
78. Kubli DA, Ycaza JE, Gustafsson AB (2007) Bnip3 mediates
mitochondrial dysfunction and cell death through Bax and Bak.
Biochem J 405:407–415. doi:10.1042/BJ20070319
79. Zong WX, Lindsten T, Ross AJ, MacGregor GR, Thompson CB
(2001) BH3-only proteins that bind pro-survival Bcl-2 family
members fail to induce apoptosis in the absence of Bax and Bak.
Genes Dev 15:1481–1486. doi:10.1101/gad.897601
80. Capano M, Crompton M (2006) Bax translocates to mitochon-
dria of heart cells during simulated ischaemia: involvement of
AMP-activated and p38 mitogen-activated protein kinases.
Biochem J 395:57–64. doi:10.1042/BJ20051654
81. Hamacher-Brady A, Brady NR, Gottlieb RA (2006) Enhancing
macroautophagy protects against ischemia/reperfusion injury in
cardiac myocytes. J Biol Chem 281:29776–29787. doi:10.1074/
jbc.M603783200
82. Hochhauser E, Kivity S, Offen D et al (2003) Bax ablation
protects against myocardial ischemia–reperfusion injury in
transgenic mice. Am J Physiol Heart Circ Physiol 284:H2351–
H2359
83. Diwan A, Krenz M, Syed FM et al (2007) Inhibition of ischemic
cardiomyocyte apoptosis through targeted ablation of Bnip3
restrains postinfarction remodeling in mice. J Clin Invest
117:2825–2833. doi:10.1172/JCI32490
84. Bruick RK (2000) Expression of the gene encoding the proa-
poptotic Nip3 protein is induced by hypoxia. Proc Natl Acad Sci
USA 97:9082–9087. doi:10.1073/pnas.97.16.9082
85. Yurkova N, Shaw J, Blackie K et al (2008) The cell cycle factor
E2F-1 activates Bnip3 and the intrinsic death pathway in
ventricular myocytes. Circ Res 102:472–479. doi:10.1161/
CIRCRESAHA.107.164731
86. Frazier DP, Wilson A, Graham RM, Thompson JW, Bishopric
NH, Webster KA (2006) Acidosis regulates the stability,
hydrophobicity, and activity of the BH3-only protein Bnip3.
Antioxid Redox Signal 8:1625–1634. doi:10.1089/ars.2006.8.
1625
87. Regula KM, Ens K, Kirshenbaum LA (2002) Inducible
expression of BNIP3 provokes mitochondrial defects and
hypoxia-mediated cell death of ventricular myocytes. Circ Res
91:226–231. doi:10.1161/01.RES.0000029232.42227.16
88. Vande VC, Cizeau J, Dubik D et al (2000) BNIP3 and genetic
control of necrosis-like cell death through the mitochondrial
permeability transition pore. Mol Cell Biol 20:5454–5468. doi:
10.1128/MCB.20.15.5454-5468.2000
89. Diwan A, Wansapura J, Syed FM, Matkovich SJ, Lorenz JN,
Dorn GW II (2008) Nix-mediated apoptosis links myocardial
fibrosis, cardiac remodeling, and hypertrophy decompensation.
Circulation 117:396–404. doi:10.1161/CIRCULATIONAHA.
107.727073
90. Chen M, He H, Zhan S, Krajewski S, Reed JC, Gottlieb RA
(2001) Bid is cleaved by calpain to an active fragment in vitro
and during myocardial ischemia/reperfusion. J Biol Chem
276:30724–30728. doi:10.1074/jbc.M103701200
91. Murriel CL, Churchill E, Inagaki K, Szweda LI, Mochly-Rosen
D (2004) Protein kinase Cdelta activation induces apoptosis in
response to cardiac ischemia and reperfusion damage: a mech-
anism involving BAD and the mitochondria. J Biol Chem
279:47985–47991. doi:10.1074/jbc.M405071200
92. Latif N, Khan MA, Birks E et al (2000) Upregulation of the Bcl-
2 family of proteins in end stage heart failure. J Am Coll Cardiol
35:1769–1777. doi:10.1016/S0735-1097(00)00647-1
93. Di Napoli P, Taccardi AA, Grilli A et al (2003) Left ventricular
wall stress as a direct correlate of cardiomyocyte apoptosis in
patients with severe dilated cardiomyopathy. Am Heart J
146:1105–1111. doi:10.1016/S0002-8703(03)00445-9
94. Baldi A, Abbate A, Bussani R et al (2002) Apoptosis and post-
infarction left ventricular remodeling. J Mol Cell Cardiol
34:165–174. doi:10.1006/jmcc.2001.1498
95. Sharma AK, Dhingra S, Khaper N, Singal PK (2007) Activation
of apoptotic processes during transition from hypertrophy to
heart failure in guinea pigs. Am J Physiol Heart Circ Physiol
293:H1384–H1390. doi:10.1152/ajpheart.00553.2007
96. Li P, Nijhawan D, Budihardjo I et al (1997) Cytochrome c and
dATP-dependent formation of Apaf-1/caspase-9 complex initi-
ates an apoptotic protease cascade. Cell 91:479–489. doi:
10.1016/S0092-8674(00)80434-1
97. Acehan D, Jiang X, Morgan DG, Heuser JE, Wang X, Akey CW
(2002) Three-dimensional structure of the apoptosome: impli-
cations for assembly, procaspase-9 binding, and activation. Mol
Cell 9:423–432. doi:10.1016/S1097-2765(02)00442-2
98. Jiang X, Wang X (2000) Cytochrome cpromotes caspase-9
activation by inducing nucleotide binding to Apaf-1. J Biol
Chem 275:31199–31203. doi:10.1074/jbc.C000405200
99. Pacher P, Szabo C (2007) Role of poly(ADP-ribose) polymerase
1 (PARP-1) in cardiovascular diseases: the therapeutic potential
of PARP inhibitors. Cardiovasc Drug Rev 25:235–260. doi:
10.1111/j.1527-3466.2007.00018.x
100. de Boer RA, van Veldhuisen DJ, van der Wijk J et al (2000)
Additional use of immunostaining for active caspase 3 and
cleaved actin and PARP fragments to detect apoptosis in
patients with chronic heart failure. J Card Fail 6:330–337. doi:
10.1054/jcaf.2000.20457
101. Vahsen N, Cande C, Briere JJ et al (2004) AIF deficiency
compromises oxidative phosphorylation. EMBO J 23:4679–
4689. doi:10.1038/sj.emboj.7600461
102. Joza N, Oudit GY, Brown D et al (2005) Muscle-specific loss of
apoptosis-inducing factor leads to mitochondrial dysfunction,
skeletal muscle atrophy, and dilated cardiomyopathy. Mol Cell
Biol 25:10261–10272. doi:10.1128/MCB.25.23.10261-10272.
2005
103. van Empel VP, Bertrand AT, van der Nagel R et al (2005) Down-
regulation of apoptosis-inducing factor in harlequin mutant mice
546 Apoptosis (2009) 14:536–548
123
sensitizesthe myocardium to oxidative stress-related cell death and
pressureoverload-induceddecompensation. Circ Res 96:e92–e101.
doi:10.1161/01.RES.0000172081.30327.28
104. Norberg E, Gogvadze V, Ott M et al (2008) An increase in
intracellular Ca(2?) is required for the activation of mitochon-
drial calpain to release AIF during cell death. Cell Death Differ
15(12):1857–1864
105. Daugas E, Susin SA, Zamzami N et al (2000) Mitochondrio-
nuclear translocation of AIF in apoptosis and necrosis. FASEB J
14:729–739
106. Halestrap AP (2006) Calcium, mitochondria and reperfusion
injury: a pore way to die. Biochem Soc Trans 34:232–237. doi:
10.1042/BST20060232
107. Garcia-Rivas Gde J, Carvajal K, Correa F, Zazueta C (2006)
Ru360, a specific mitochondrial calcium uptake inhibitor,
improves cardiac post-ischaemic functional recovery in rats in
vivo. Br J Pharmacol 149:829–837. doi:10.1038/sj.bjp.0706932
108. Wang J, Zhang Z, Hu Y et al (2007) SEA0400, a novel Na(?)/
Ca(2 ?) exchanger inhibitor, reduces calcium overload induced
by ischemia and reperfusion in mouse ventricular myocytes.
Physiol Res 56:17–23
109. Kim GT, Chun YS, Park JW, Kim MS (2003) Role of apoptosis-
inducing factor in myocardial cell death by ischemia–reperfu-
sion. Biochem Biophys Res Commun 309:619–624. doi:
10.1016/j.bbrc.2003.08.045
110. Song ZF, Ji XP, Li XX, Wang SJ, Wang SH, Zhang Y (2008)
Inhibition of the activity of poly (ADP-ribose) polymerase
reduces heart ischaemia/reperfusion injury via suppressing JNK-
mediated AIF translocation. J Cell Mol Med 12:1220–1228. doi:
10.1111/j.1582-4934.2008.00183.x
111. Roy N, Deveraux QL, Takahashi R, Salvesen GS, Reed JC
(1997) The c-IAP-1 and c-IAP-2 proteins are direct inhibitors of
specific caspases. EMBO J 16:6914–6925. doi:10.1093/emboj/
16.23.6914
112. Deveraux QL, Takahashi R, Salvesen GS, Reed JC (1997) X-
linked IAP is a direct inhibitor of cell-death proteases. Nature
388:300–304. doi:10.1038/40901
113. Liu HR, Gao E, Hu A et al (2005) Role of Omi/HtrA2 in apoptotic
cell death after myocardial ischemia and reperfusion. Circulation
111:90–96. doi:10.1161/01.CIR.0000151613.90994.17
114. Bhuiyan MS, Fukunaga K (2007) Inhibition of HtrA2/Omi
ameliorates heart dysfunction following ischemia/reperfusion
injury in rat heart in vivo. Eur J Pharmacol 557:168–177. doi:
10.1016/j.ejphar.2006.10.067
115. Chua CC, Gao J, Ho YS et al (2007) Overexpression of IAP-2
attenuates apoptosis and protects against myocardial ischemia/
reperfusion injury in transgenic mice. Biochim Biophys Acta
1773:577–583
116. Ikeda S, Ozaki K (1997) Action of mitochondrial endonuclease
G on DNA damaged by L-ascorbic acid, peplomycin, and cis-
diamminedichloroplatinum (II). Biochem Biophys Res Commun
235:291–294. doi:10.1006/bbrc.1997.6786
117. Javadov S, Choi A, Rajapurohitam V, Zeidan A, Basnakian AG,
Karmazyn M (2008) NHE-1 inhibition-induced cardioprotection
against ischaemia/reperfusion is associated with attenuation of
the mitochondrial permeability transition. Cardiovasc Res
77:416–424. doi:10.1093/cvr/cvm039
118. Bahi N, Zhang J, Llovera M, Ballester M, Comella JX, Sanchis
D (2006) Switch from caspase-dependent to caspase-indepen-
dent death during heart development: essential role of
endonuclease G in ischemia-induced DNA processing of dif-
ferentiated cardiomyocytes. J Biol Chem 281:22943–22952. doi:
10.1074/jbc.M601025200
119. Szegezdi E, Logue SE, Gorman AM, Samali A (2006) Media-
tors of endoplasmic reticulum stress-induced apoptosis. EMBO
Rep 7:880–885. doi:10.1038/sj.embor.7400779
120. Xu C, Bailly-Maitre B, Reed JC (2005) Endoplasmic reticulum
stress: cell life and death decisions. J Clin Invest 115:2656–
2664. doi:10.1172/JCI26373
121. Okada K, Minamino T, Tsukamoto Y et al (2004) Prolonged
endoplasmic reticulum stress in hypertrophic and failing heart
after aortic constriction: possible contribution of endoplasmic
reticulum stress to cardiac myocyte apoptosis. Circulation
110:705–712. doi:10.1161/01.CIR.0000137836.95625.D4
122. Szegezdi E, Duffy A, O’Mahoney ME et al (2006) ER stress
contributes to ischemia-induced cardiomyocyte apoptosis. Bio-
chem Biophys Res Commun 349:1406–1411. doi:10.1016/
j.bbrc.2006.09.009
123. Thuerauf DJ, Marcinko M, Gude N, Rubio M, Sussman MA,
Glembotski CC (2006) Activation of the unfolded protein
response in infarcted mouse heart and hypoxic cultured cardiac
myocytes. Circ Res 99:275–282. doi:10.1161/01.RES.0000233
317.70421.03
124. Azfer A, Niu J, Rogers LM, Adamski FM, Kolattukudy PE
(2006) Activation of endoplasmic reticulum stress response
during the development of ischemic heart disease. Am J Physiol
Heart Circ Physiol 291:H1411–H1420. doi:10.1152/ajpheart.
01378.2005
125. Jackson CV, McGrath GM, Tahiliani AG, Vadlamudi RV,
McNeill JH (1985) A functional and ultrastructural analysis of
experimental diabetic rat myocardium.Manifestation of a car-
diomyopathy. Diabetes 34:876–883. doi:10.2337/diabetes.34.
9.876
126. Hamada H, Suzuki M, Yuasa S et al (2004) Dilated cardiomy-
opathy caused by aberrant endoplasmic reticulum quality
control in mutant KDEL receptor transgenic mice. Mol Cell Biol
24:8007–8017. doi:10.1128/MCB.24.18.8007-8017.2004
127. Breckenridge DG, Germain M, Mathai JP, Nguyen M, Shore GC
(2003) Regulation of apoptosis by endoplasmic reticulum path-
ways. Oncogene 22:8608–8618. doi:10.1038/sj.onc.1207108
128. Qi X, Vallentin A, Churchill E, Mochly-Rosen D (2007)
deltaPKC participates in the endoplasmic reticulum stress-
induced response in cultured cardiac myocytes and ischemic
heart. J Mol Cell Cardiol 43:420–428. doi:10.1016/j.yjmcc.
2007.07.061
129. Toth A, Jeffers JR, Nickson P et al (2006) Targeted deletion of
puma attenuates cardiomyocyte death and improves cardiac
function during ischemia-reperfusion. Am J Physiol Heart Circ
Physiol 291(1):H52–H60
130. Nickson P, Toth A, Erhardt P (2007) PUMA is critical for neonatal
cardiomyocyte apoptosis induced by endoplasmic reticulum
stress. Cardiovasc Res 73:48–56. doi:10.1016/j.cardiores.2006.
10.001
131. Scorrano L, Oakes SA, Opferman JT et al (2003) BAX and BAK
regulation of endoplasmic reticulum Ca2?: a control point for
apoptosis. Science 300:135–139. doi:10.1126/science.1081208
132. Luo X, Budihardjo I, Zou H, Slaughter C, Wang X (1998) Bid, a
Bcl2 interacting protein, mediates cytochrome c release
from mitochondria in response to activation of cell surface death
receptors. Cell 94:481–490. doi:10.1016/S0092-8674(00)81589-5
133. Li H, Zhu H, Xu CJ, Yuan J (1998) Cleavage of BID by
caspase 8 mediates the mitochondrial damage in the Fas
pathway of apoptosis. Cell 94:491–501. doi:10.1016/S0092-
8674(00)81590-1
134. Date T, Mochizuki S, Belanger AJ et al (2003) Differential
effects of membrane and soluble Fas ligand on cardiomyocytes:
role in ischemia/reperfusion injury. J Mol Cell Cardiol 35:811–
821. doi:10.1016/S0022-2828(03)00139-1
135. Haudek SB, Taffet GE, Schneider MD, Mann DL (2007) TNF
provokes cardiomyocyte apoptosis and cardiac remodeling
through activation of multiple cell death pathways. J Clin Invest
117:2692–2701. doi:10.1172/JCI29134
Apoptosis (2009) 14:536–548 547
123
136. Kalai M, Lamkanfi M, Denecker G et al (2003) Regulation of
the expression and processing of caspase-12. J Cell Biol
162:457–467. doi:10.1083/jcb.200303157
137. Bajaj G, Sharma RK (2006) TNF-alpha-mediated cardiomyo-
cyte apoptosis involves caspase-12 and calpain. Biochem
Biophys Res Commun 345:1558–1564. doi:10.1016/j.bbrc.2006.
05.059
138. Gyan E, Frisan E, Beyne-Rauzy O et al (2008) Spontaneous and
Fas-induced apoptosis of low-grade MDS erythroid precursors
involves the endoplasmic reticulum. Leukemia 22:1864–1873.
doi:10.1038/leu.2008.172
139. Breckenridge DG, Stojanovic M, Marcellus RC, Shore GC
(2003) Caspase cleavage product of BAP31 induces mitochon-
drial fission through endoplasmic reticulum calcium signals,
enhancing cytochrome c release to the cytosol. J Cell Biol
160:1115–1127. doi:10.1083/jcb.200212059
140. Mathai JP, Germain M, Shore GC (2005) BH3-only BIK regu-
lates BAX, BAK-dependent release of Ca2?from endoplasmic
reticulum stores and mitochondrial apoptosis during stress-
induced cell death. J Biol Chem 280:23829–23836. doi:10.1074/
jbc.M500800200
141. Nakai A, Yamaguchi O, Takeda T et al (2007) The role of
autophagy in cardiomyocytes in the basal state and in response
to hemodynamic stress. Nat Med 13:619–624. doi:10.1038/
nm1574
142. Matsui Y, Takagi H, Qu X et al (2007) Distinct roles of
autophagy in the heart during ischemia and reperfusion: roles of
AMP-activated protein kinase and Beclin 1 in mediating
autophagy. Circ Res 100:914–922. doi:10.1161/01.RES.0000
261924.76669.36
143. Zhu H, Tannous P, Johnstone JL et al (2007) Cardiac autophagy
is a maladaptive response to hemodynamic stress. J Clin Invest
117:1782–1793. doi:10.1172/JCI27523
144. Yan L, Vatner DE, Kim SJ et al (2005) Autophagy in chroni-
cally ischemic myocardium. Proc Natl Acad Sci USA
102:13807–13812. doi:10.1073/pnas.0506843102
145. Yousefi S, Perozzo R, Schmid I et al (2006) Calpain-mediated
cleavage of Atg5 switches autophagy to apoptosis. Nat Cell Biol
8:1124–1132. doi:10.1038/ncb1482
146. Pyo JO, Jang MH, Kwon YK et al (2005) Essential roles of Atg5
and FADD in autophagic cell death: dissection of autophagic
cell death into vacuole formation and cell death. J Biol Chem
280:20722–20729. doi:10.1074/jbc.M413934200
147. Pattingre S, Tassa A, Qu X et al (2005) Bcl-2 antiapoptotic
proteins inhibit Beclin 1-dependent autophagy. Cell 122:927–
939. doi:10.1016/j.cell.2005.07.002
148. Maiuri MC, Criollo A, Tasdemir E et al (2007) BH3-only pro-
teins and BH3 mimetics induce autophagy by competitively
disrupting the interaction between Beclin 1 and Bcl-2/Bcl-X(L).
Autophagy 3:374–376
149. Rashmi R, Pillai SG, Vijayalingam S, Ryerse J, Chinnadurai G
(2008) BH3-only protein BIK induces caspase-independent cell
death with autophagic features in Bcl-2 null cells. Oncogene
27:1366–1375. doi:10.1038/sj.onc.1210783
150. Akar U, Chaves-Reyez A, Barria M et al (2008) Silencing of
Bcl-2 expression by small interfering RNA induces autophagic
cell death in MCF-7 breast cancer cells. Autophagy 4:669–679
151. Ray R, Chen G, Vande VC et al (2000) BNIP3 heterodimerizes
with Bcl-2/Bcl-X(L) and induces cell death independent of a
Bcl-2 homology 3 (BH3) domain at both mitochondrial and
nonmitochondrial sites. J Biol Chem 275:1439–1448. doi:
10.1074/jbc.275.2.1439
548 Apoptosis (2009) 14:536–548
123
... In apoptosis and cell survival pathways, HDACs affect the expression of pro-apoptotic and anti-apoptotic genes, thus maintaining the equilibrium between cell death and survival signals [198]. Apoptosis has been linked to several cardiovascular diseases, and a positive correlation between an increased EC apoptosis and the development and progression of cardiovascular diseases has been established [199][200][201][202]. In a HUVEC model of disturbed flow (atherosclerosis), β-catenin positively regulates endothelial nitric oxide synthase (eNOS) activity and anti-apoptotic gene expression [203]. ...
Chapter
Full-text available
A monolayer of endothelial cells (ECs) lines the lumen of blood vessels and, as such, provides a semi-selective barrier between the blood and the interstitial space. Compromise of the lung EC barrier due to inflammatory or toxic events may result in pulmonary edema, which is a cardinal feature of acute lung injury (ALI) and its more severe form, acute respiratory distress syndrome (ARDS). The EC functions are controlled, at least in part, via epigenetic mechanisms mediated by histone deacetylases (HDACs). Zinc-dependent HDACs represent the largest group of HDACs and are activated by Zn2+. Members of this HDAC group are involved in epigenetic regulation primarily by modifying the structure of chromatin upon removal of acetyl groups from histones. In addition, they can deacetylate many non-histone histone proteins, including those located in extranuclear compartments. Recently, the therapeutic potential of inhibiting zinc-dependent HDACs for EC barrier preservation has gained momentum. However, the role of specific HDAC subtypes in EC barrier regulation remains largely unknown. This review aims to provide an update on the role of zinc-dependent HDACs in endothelial dysfunction and its related diseases. We will broadly focus on biological contributions, signaling pathways and transcriptional roles of HDACs in endothelial pathobiology associated mainly with lung diseases, and we will discuss the potential of their inhibitors for lung injury prevention.
... 12 In apoptosis and cell survival pathways, HDACs affect the expression of pro-apoptotic and antiapoptotic genes, thus maintaining the equilibrium between cell death and survival signals [198]. Apoptosis has been linked to several cardiovascular diseases, and a positive correlation between an increased EC apoptosis and the development and progression of cardiovascular diseases has been established [199][200][201][202]. In a HUVEC model of disturbed flow (atherosclerosis), β-catenin positively regulates endothelial nitric oxide synthase (eNOS) activity and anti-apoptotic gene expression [203]. ...
Preprint
Full-text available
A monolayer of endothelial cells (ECs) lines the lumen of blood vessels and as such provides a semi-selective barrier between the blood and the interstitial space. Compromise of the lung EC barrier due to inflammatory or toxic events may results in pulmonary edema, which is a cardinal feature of acute lung injury (ALI) and its more severe form, the acute respiratory distress syndrome (ARDS). The EC functions are controlled, at least in part, via epigenetic mechanisms mediated by histone deacetylases (HDACs). Zinc-dependent HDACs represent the largest group of HDACs and are activated by Zn2+. Members of this HDAC group are involved in the epigenetic regulation primarily via modifying the structure of chromatin upon removal of acetyl groups from histones. In addition, they can deacetylate many non-histone histone proteins, including those located in extra nuclear compartments. Recently, the therapeutic potential of inhibiting zinc-dependent HDACs for EC barrier preservation has gained momentum. However, the role of specific HDAC subtypes in EC barrier regulation remains largely unknown. This review aims to provide an update on the role of zinc-dependent HDACs in endothelial dysfunction and its related diseases. We will broadly focus on biological contributions, signaling pathways and transcriptional roles of HDACs in endothelial pathobiology associated mainly with lung diseases and we will discuss the potential of their inhibitors for lung injury prevention.
Article
Full-text available
Diabetes mellitus (DM) is a highly prevalent disease worldwide, estimated to affect 1 in every 11 adults; among them, 90–95% of cases are type 2 diabetes mellitus. This is partly attributed to the surge in the prevalence of obesity, which has reached epidemic proportions since 2008. In these patients, cardiovascular (CV) risk stands as the primary cause of morbidity and mortality, placing a substantial burden on healthcare systems due to the potential for macrovascular and microvascular complications. In this context, leptin, an adipocyte-derived hormone, plays a fundamental role. This hormone is essential for regulating the cellular metabolism and energy balance, controlling inflammatory responses, and maintaining CV system homeostasis. Thus, leptin resistance not only contributes to weight gain but may also lead to increased cardiac inflammation, greater fibrosis, hypertension, and impairment of the cardiac metabolism. Understanding the relationship between leptin resistance and CV risk in obese individuals with type 2 DM (T2DM) could improve the management and prevention of this complication. Therefore, in this narrative review, we will discuss the evidence linking leptin with the presence, severity, and/or prognosis of obesity and T2DM regarding CV disease, aiming to shed light on the potential implications for better management and preventive strategies.
Article
Full-text available
A monolayer of endothelial cells (ECs) lines the lumen of blood vessels and, as such, provides a semi-selective barrier between the blood and the interstitial space. Compromise of the lung EC barrier due to inflammatory or toxic events may result in pulmonary edema, which is a cardinal feature of acute lung injury (ALI) and its more severe form, acute respiratory distress syndrome (ARDS). The EC functions are controlled, at least in part, via epigenetic mechanisms mediated by histone deacetylases (HDACs). Zinc-dependent HDACs represent the largest group of HDACs and are activated by Zn 2+. Members of this HDAC group are involved in epigenetic regulation primarily by modifying the structure of chromatin upon removal of acetyl groups from histones. In addition, they can deacetylate many non-histone histone proteins, including those located in extranuclear compartments. Recently, the therapeutic potential of inhibiting zinc-dependent HDACs for EC barrier preservation has gained momentum. However, the role of specific HDAC subtypes in EC barrier regulation remains largely unknown. This review aims to provide an update on the role of zinc-dependent HDACs in endothelial dysfunction and its related diseases. We will broadly focus on biological contributions, signaling pathways and transcriptional roles of HDACs in endothelial pathobiology associated mainly with lung diseases, and we will discuss the potential of their inhibitors for lung injury prevention.
Article
Full-text available
A growing body of literature has attempted to characterize how traffic-related air pollution (TRAP) affects molecular and subclinical biological processes in ways that could lead to cardiorespiratory disease. To provide a streamlined synthesis of what is known about the multiple mechanisms through which TRAP could lead to cardiorespiratory pathology, we conducted a systematic review of the epidemiological literature relating TRAP exposure to methylomic, proteomic, and metabolomic biomarkers in adult populations. Using the 139 papers that met our inclusion criteria, we identified the omic biomarkers significantly associated with short- or long-term TRAP and used these biomarkers to conduct pathway and network analyses. We considered the evidence for TRAP-related associations with biological pathways involving lipid metabolism, cellular energy production, amino acid metabolism, inflammation and immunity, coagulation, endothelial function, and oxidative stress. Our analysis suggests that an integrated multi-omics approach may provide critical new insights into the ways TRAP could lead to adverse clinical outcomes. We advocate for efforts to build a more unified approach for characterizing the dynamic and complex biological processes linking TRAP exposure and subclinical and clinical disease and highlight contemporary challenges and opportunities associated with such efforts.
Article
Cardiomyocyte apoptosis is an important factor in cardiac function decline observed in various cardiovascular diseases. To understand the progress in the field of cardiomyocyte apoptosis research, this paper uses bibliometrics to statistically analyze publications in this field. A total of 5939 articles were retrieved from the core Web of Science database, and then VOSviewer and Citespace were used to conduct a scientometric analysis of the authors, countries, institutions, references and keywords included in the articles to determine the cooperative relationships between researchers that study cardiomyocyte apoptosis. At present, the research hotspots in this field mainly include experimental research, molecular mechanisms, pathophysiology and cardiac regeneration of cardiomyocyte apoptosis-related diseases. NOD-like receptor thermal protein domain associated protein 3 inflammasome, circular RNA, and sepsis are the research frontiers in this field and are emerging as new areas of research focus. This work provides insight into research directions and the clinical application value for the continued advancement of cardiomyocyte apoptosis research.
Article
Background Myocardial infarction poses major risks to human health because of their incredibly high rates of morbidity and mortality. Infarctions are more likely to develop as a result of dysregulation of cell death. Myrtenal can be considered for their bioactive beneficial activity in the context of cardiovascular pathologies and, particularly, in the protection toward oxidative stress followed by ischemic injury. Objective This study aimed to put limelight on the antioxidant, anti‐apoptotic, and antibacterial properties of Myrtenal. Methods An in vitro model of oxidative stress‐induced injury was entrenched in H9c2 cells using hydrogen peroxide, and the effects of Myrtenal were investigated. The MTT, cellular enzyme level, staining, and flow cytometry analysis were used to examine protective, antioxidant, and anti‐apoptotic effects. The gene expressions were detected by qPCR. Antibacterial effect and biofilm formation were also done. Result The findings revealed that Myrtenal alone had negligible cytotoxic effects and that Myrtenal protects H9c2 against H 2 O 2 ‐induced cell death at micromolar concentrations. Myrtenal pre‐treatment inhibited the generation of reactive oxygen species (ROS) as well as remarkably decreased the fluorescence intensity of ROS. Additionally, Myrtenal considerably increased the synthesis of antioxidant enzymes while dramatically decreasing the production of MDA and LDH. qPCR demonstrated the downregulation of Cas‐9, TNF‐α, NF‐κB, P53, BAX, iNOS, and IL‐6 expression while an upregulation of Bcl‐2 expression in Myrtenal pre‐treated groups. Myrtenal also holds the magnificent property of inhibiting bacterial growth. Conclusion Myrtenal ameliorates H 2 O 2 ‐induced cardiomyocyte injury and protects cardiomyocyte by inhibiting oxidative stress, inflammation, and apoptosis and may be a promise drug for the treatment of heart diseases.
Article
Full-text available
Cardiac hypertrophy is a major predictor of heart failure and a prevalent disorder with high mortality. Little is known, however, regarding mechanisms governing the transition from stable cardiac hypertrophy to decompensated heart failure. Here, we tested the role of autophagy, a conserved pathway mediating bulk degradation of long-lived proteins and cellular organelles that can lead to cell death. To quantify autophagic activity, we engineered a line of “autophagy reporter” mice and confirmed that cardiomyocyte autophagy can be induced by short-term nutrient deprivation in vivo. Pressure overload induced by aortic banding induced heart failure and greatly increased cardiac autophagy. Load-induced autophagic activity peaked at 48 hours and remained significantly elevated for at least 3 weeks. In addition, autophagic activity was not spatially homogeneous but rather was seen at particularly high levels in basal septum. Heterozygous disruption of the gene coding for Beclin 1, a protein required for early autophagosome formation, decreased cardiomyocyte autophagy and diminished pathological remodeling induced by severe pressure stress. Conversely, Beclin 1 overexpression heightened autophagic activity and accentuated pathological remodeling. Taken together, these findings implicate autophagy in the pathogenesis of load-induced heart failure and suggest it may be a target for novel therapeutic intervention.
Article
Extracellular survival factors alter a cell's susceptibility to apoptosis, often through posttranslational mechanisms. However, no consistent relationship has been established between such survival signals and the BCL-2 family, where the balance of death agonists versus antagonists determines susceptibility. One distant member, BAD, heterodimerizes with BCL-X(L) or BCL-2, neutralizing their protective effect and promoting cell death. In the presence of survival factor IL-3, cells phosphorylated BAD on two serine residues embedded in 14-3-3 consensus binding sites. Only the nonphosphorylated BAD heterodimerized with BCL-X(L) at membrane sites to promote cell death. Phosphorylated BAD was sequestered in the cytosol bound to 14-3-3. Substitution of serine phosphorylation sites further enhanced BAD's death-promoting activity. The rapid phosphorylation of BAD following IL-3 connects a proximal survival signal with the BCL-2 family, modulating this checkpoint for apoptosis.
Article
The historical development of the cell death concept is reviewed, with special attention to the origin of the terms necrosis, coagulation necrosis, autolysis, physiological cell death, programmed cell death, chromatolysis (the first name of apoptosis in 1914), karyorhexis, karyolysis, and cell suicide, of which there are three forms: by lysosomes, by free radicals, and by a genetic mechanism (apoptosis). Some of the typical features of apoptosis are discussed, such as budding (as opposed to blebbing and zeiosis) and the inflammatory response. For cell death not by apoptosis the most satisfactory term is accidental cell death. Necrosis is commonly used but it is not appropriate, because it does not indicate a form of cell death but refers to changes secondary to cell death by any mechanism, including apoptosis. Abundant data are available on one form of accidental cell death, namely ischemic cell death, which can be considered an entity of its own, caused by failure of the ionic pumps of the plasma membrane. Because ischemic cell death (in known models) is accompanied by swelling, the name oncosis is proposed for this condition. The term oncosis (derived from onkos, meaning swelling) was proposed in 1910 by von Reckling-hausen precisely to mean cell death with swelling. Oncosis leads to necrosis with karyolysis and stands in contrast to apoptosis, which leads to necrosis with karyorhexis and cell shrinkage.
Article
Apoptosis contributes, with necrosis, to the cardiac cell loss after ischemia/reperfusion injury. The apoptotic cascade is initiated either by mitochondrial damage and activation of caspase-9 or by death receptor ligation and activation of caspase-8. In the present study, performed in the isolated rat heart exposed either to ischemia alone or ischemia followed by reperfusion, cleavage of caspase-9 was observed primarily in endothelial cells. Conversely, caspase-8 cleavage was only found in cardiomyocytes, where it progressively increased throughout reperfusion. Addition of a specific caspase-9 inhibitor to the perfusate before ischemia prevented endothelial apoptosis, whereas preischemic infusion of a specific caspase-8 inhibitor affected only myocyte apoptosis. Additionally, caspase-8 – mediated BID processing was observed only during reperfusion. Production of tBID then sustains mitochondrial injury and perpetuates caspase-9 activation. (Circ Res. 2002;90:745-748.) Key Words: apoptosis endothelium myocytes ischemia reperfusion A poptosis is an active form of cell suicide affecting both endothelial cells and cardiac myocytes during ischemia/ reperfusion injury. Mitochondrial damage leads to activation of the initiator protease, caspase-9, which then propagates the cascade of downstream caspases. In contrast, ligation of death receptors, such as Fas, activates caspase-8, which also then processes downstream effector enzymes. 1 Enzymatically active caspase-8 cleaves BID, and the truncated protein, tBID, relocates to the mitochondria where it induces caspase-9 processing. 2 In this study, we have addressed three questions in the isolated rat heart: the level of enzymatic activity of the two initiator caspases during ischemia versus reperfusion; the selective contribution of these caspases to apoptosis of endothelial cells and myocytes; and the role played by BID in linking the two pathways.
Article
Conditions of diastolic overload associated with increases in filling pressure trigger apoptosis. Moreover, ischemia alone and ischemia followed by reperfusion induce programmed cell death in myocytesin vitro. On this basis, the possibility was raised that apoptotic myocyte cell death may occur in the surviving myocardium acutely after infarction. Myocardial samples were obtained from the region adjacent to and remote from infarction in patients who died within 10 days from the initial clinical symptoms. Apoptosis was measured quantitatively by the terminal deoxynucleotidyl transferase assay and confirmed biochemically by DNA extraction and agarose gel electrophoresis. This analysis included 20 infarcted and ten control hearts. DNA strand breaks in myocyte nuclei were observed in all 20 infarcted hearts in both the regions bordering on and distant from the necrotic myocardium. However, the number of apoptotic nuclei was greater in the peri-infarcted region than in that away from infarction. Quantitatively, 12% of myocytes in the border zone showed DNA strand breaks, whereas 1% of cells were undergoing apoptosis in the remote myocardium. Moreover, DNA laddering was detected biochemically in these two regions of the heart. Thus, apoptosis appears to be a significant complicating factor of acute myocardial infarction increasing the magnitude of myocyte cell death associated with coronary artery occlusion.
Article
Objectives: This study sought to determine the circulating levels of cytokines and their respective endogenous modulators in patients with congestive heart failure of variable severity. Background: Activation of immune elements localized in the heart or periphery, or both, may promote release of cytokines in patients with congestive heart failure. Although an increased circulating level of tumor necrosis factor-alpha (TNF-alpha) and its soluble receptor type II (sTNF-RII) is well documented, less is known about other cytokines (i.e., interleukin-1-beta [IL-1-beta], interleukin-6 [IL-6] and interleukin-2 [IL-2] and their soluble receptor/receptor antagonists). Methods: Circulating levels of TNF-alpha and sTNF-RII, IL-1-beta, IL-1 receptor antagonist (IL-1-Ra), IL-6, IL-6 soluble receptor (IL-6-sR), IL-2 and IL-2 soluble receptor-alpha were measured using enzyme-linked immunosorbent assay kits (Quantikine, R&D Systems) in 80 patients with congestive heart failure due to coronary artery disease or hypertension. The severity of their symptoms, which ranged from New York Heart Association functional class I to IV, was confirmed by measurement of peak oxygen consumption. Results: The percentage of patients with elevated levels of cytokines and their corresponding soluble receptor/receptor antagonists significantly increased with functional class. For TNF-alpha and IL-1-beta, the percentage of patients with elevated levels of soluble receptor/receptor antagonists was higher than that of patients with elevated levels of the cytokine itself. For IL-6, the percentage of patients with elevated levels of IL-6-sR tended to be lower than that of patients with elevated levels of IL-6. All but two patients had undetectable levels of IL-2, and all but seven had levels of IL-2-sR within a normal range. Conclusions: In patients with congestive heart failure, circulating levels of cytokines increased with the severity of symptoms. In these patients, circulating levels of sTNF-RII and IL-1-Ra are more sensitive markers of immune activation than are circulating levels of TNF-alpha and IL-1-beta, respectively. Levels of IL-2 and IL-2-sR are not elevated when congestive heart failure is due to coronary artery disease or hypertension.
Article
The heart is a tumor necrosis factor (TNF)-producing organ. Both myocardial macrophages and cardiac myocytes themselves synthesize TNF. Accumulating evidence indicates that myocardial TNF is an autocrine contributor to myocardial dysfunction and cardiomyocyte death in ischemia-reperfusion injury, sepsis, chronic heart failure, viral myocarditis, and cardiac allograft rejection. Indeed, locally (vs. systemically) produced TNF contributes to postischemic myocardial dysfunction via direct depression of contractility and induction of myocyte apoptosis. Lipopolysaccharide or ischemia-reperfusion activates myocardial P38 mitogen-activated protein (MAP) kinase and nuclear factor kappa B, which lead to TNF production. TNF depresses myocardial function by nitric oxide (NO)-dependent and NO-independent (sphingosine dependent) mechanisms. TNF activation of TNF receptor 1 or Fas may induce cardiac myocyte apoptosis. MAP kinases and TNF transcription factors are feasible targets for anti-TNF (i.e., cardioprotective) strategies. Endogenous anti-inflammatory ligands, which trigger the gp130 signaling cascade, heat shock proteins, and TNF-binding proteins, also control TNF production and activity. Thus modulation of TNF in cardiovascular disease represents a realistic goal for clinical medicine.
Article
The occurrence of myocyte necrosis during reperfusion of ischemic myocardium is controversial. This study measured myocardial 2-deoxyglucose uptake, correlated with histology, to determine whether loss of viability occurred during reperfusion of ischemic myocardium. In 12 anesthetized dogs, the left anterior descending coronary artery was occluded for 90 minutes before 4 hours reperfusion. Myocardial blood flow was measured by microspheres and the tracers 14C-2-deoxyglucose and 18F-2-deoxyglucose were injected intravenously after 5 and 180 minutes of reperfusion, respectively. After 240 minutes, the heart was stained with thioflavin-S (size of no-reflow zone) and triphenyl-tetrazolium chloride (TTC, extent of necrosis). Samples from normal, salvaged, and necrotic myocardium were counted for 14C- and 18F-deoxyglucose and microspheres. With the use of a three-compartment model of 2-deoxyglucose uptake, the rate constant k3 for phosphorylation of 14C- and 18F-2-deoxyglucose was calculated for each sample. Viability was defined as k3> or = 0.125 min(-1) (predictive accuracy 88% versus electron microscopy and 97% versus TTC). Among 58 samples from no-reflow regions, 97% were nonviable after 5 minutes of reperfusion (k3=0.096 +/- 0.027 min[-1]). Among 164 samples from salvaged myocardium, 95% were viable after both 5 and 180 minutes of reperfusion (k3=0.170 +/- 0.056 min[-1] P<.01 versus no-reflow). Among 179 samples from infarcted myocardium, mean k3 after 5 minutes of reperfusion was 0.184 +/- 0.070 min(-1) and 65% of samples were viable, but after 180 minutes of reperfusion mean k3 had decreased to 0.077 +/- 0.032 min(-1) (P<.0001) and 98% of samples were nonviable. A large proportion of samples from infarcted myocardium are viable at the end of the ischemic period but lose viability during the first hours of reperfusion.