SlideShare a Scribd company logo
1 of 13
Download to read offline
Review
eNOS phosphorylation in health and disease
Gopi Krishna Kolluru, Jamila H. Siamwala, Suvro Chatterjee*
Vascular Biology Lab, AU-KBC Research Centre, MIT Campus, Anna University, Chennai 600 044, TN, India
a r t i c l e i n f o
Article history:
Received 13 February 2010
Accepted 29 March 2010
Available online 2 April 2010
Keywords:
eNOS
Serine phosphorylation
Nitric oxide
a b s t r a c t
Endothelium plays a fundamental role in maintaining the vascular tone by releasing various biochemical
factors that modulate the contractile and relaxatory behavior of the underlying vascular smooth muscle,
regulation of inflammation, immunomodulation, platelet aggregation, and thrombosis. Endothelium
regulates these cellular processes by activating endothelial nitric oxide synthase (eNOS) responsible for
nitric oxide (NO) production. eNOS is constitutively expressed in ECs in response to humoral, mechanical
or pharmacological stimulus. eNOS activity is regulated mainly by protein-protein interactions and
multisite phosphorylations. The phosphorylation state of specific serine, threonine and tyrosine residues
of the enzyme plays a pivotal role in regulation of eNOS activity. Perturbations of eNOS phosphorylation
have been reported in a number of diseases thereby emphasizing the importance of regulation of eNOS
activity. This review summarizes the mechanism of eNOS regulation through multi-site phosphorylation
in different pathologies. Attempts have been made to highlight phosphorylation of eNOS at various
residues, regulation of the enzyme activity via posttranslational modifications and its implications on
health and disease.
Ó 2010 Elsevier Masson SAS. All rights reserved.
1. Introduction
There has been a drastic transformation from times past when
eNOS regulation was thought to be a very simple process. Many
studies and research groups working on eNOS regulation have
come a long way in discovering the increasing number of inter-
acting protein kinases and the role of eNOS phosphorylation on
specific serine (Ser) and threonine (Thr) residues. The plethora of
studies on eNOS phosphorylation and control of NO synthesis has
furthered the understanding of pathophysiological conditions of
endothelial dysfunctions and rectification. Inspite of several studies
on eNOS phosphorylation over a decade, the complexity of eNOS
regulation still exists and poses to be a topic of interest for
researchers. The changes in eNOS activity in relation to eNOS
phosphorylation pattern will provide information regarding
attenuated NO availability and the pathogenic mechanisms
involved in multiple disease conditions and their potential treat-
ment conditions related to NO deficiency. This review attempts to
summarize the present knowledge of the multi-site phosphoryla-
tion of eNOS under different physiological and pathological
conditions.
1.1. NO in physiology and pathology
Nitric oxide (NO) is one of the simplest gaseous free radicals and
mediates a significant and diverse number of signaling functions in
nearly every organ system in the body. NO produced in the endo-
thelium via the enzyme eNOS is an important vasoactive compound
[1,2]. NO is responsible to regulate a diverse range of physiological
and cellular processes including endothelial cell migration, prolif-
eration, extracellular matrix degradation, and angiogenesis [3]. NO
is a key component in endothelium-dependent regulation of
vascular tone, platelet function, angiogenesis and mitogenesis
that are crucial in cardiovascular physiology [4]. A potent anti-
inflammatory agent, NO inhibits leukocyte interactions with the
vessel wall, thereby reducing pathological inflammation and
thrombosis [5]. eNOS knockout mice studies showed that leukocyte
adherence to the vessel wall is elevated 10-fold [6]. Also NO inhibits
inflammation in various vascular disease models like myocardial
infarction, glomerulonephritis, lung injury, and stroke [7,8].
Thereby NO controls inflammation and thrombosis in part by
regulating vesicle trafficking. NO decreases granule trafficking from
the Golgi to the plasma membrane by targeting a key component of
exocytic machinery [9]. Loss of endothelial function due to phos-
phorylation and sub-cellular localization of eNOS [11,12], has been
implicated in a number of cardiovascular diseases, making regula-
tion of eNOS through posttranslational modifications, a promising
therapeutic target [2,10].
* Corresponding author. Tel.: þ91 44 2223 4885/2711x48; fax: þ91 44 2223 1034.
E-mail address: soovro@yahoo.ca (S. Chatterjee).
Contents lists available at ScienceDirect
Biochimie
journal homepage: www.elsevier.com/locate/biochi
0300-9084/$ e see front matter Ó 2010 Elsevier Masson SAS. All rights reserved.
doi:10.1016/j.biochi.2010.03.020
Biochimie 92 (2010) 1186e1198
1.2. Expression and regulation of eNOS
NO is an important signaling molecule but is also highly
reactive and highly diffusible. It is therefore important that there
is strict control and regulation of NO production. NO is produced
by a group of enzymes called neuronal NOS (nNOS), inducible NOS
(iNOS) and endothelial NOS (eNOS) [13,14]. These enzymes
convert arginine into citrulline, producing NO in the process. The
activity of NOS enzymes is subject to a discreet and multiple
interconnected mechanisms of regulation. There are many ways
by which regulation of NOS occurs, such as the intracellular
distribution, gene expression, protein binding, enzymatic activa-
tion by phosphorylation and cellular inhibitors of NOS activity.
iNOS expression is controlled by inflammatory mediators and
cytokines and produces large, unregulated quantities of NO,
whereas nNOS and eNOS produce low amounts of NO. nNOS and
especially eNOS activity are induced by physiological and patho-
physiological stimuli [15e17]. Various chemical stimuli like VEGF,
estrogen receptor modulator and sphingosine-1-phosphate or
mechanical factors like shear stress activate eNOS [18e20]. eNOS
is regulated by multiple interdependent control mechanisms and
signaling pathways which can be calcium-dependent and/or
-independent [13,21e24].
1.3. Mechansim of eNOS activation
It is well-established that eNOS is a calmodulin dependent
enzyme [6,25]. eNOS is activated by elevation of intracellular
calcium and subsequent activation of CaM-binding domain by
calmodulin [26,27]. eNOS comprises of a C-terminal reductase
domain, which transfers the electron flux via NADPH to the
N-terminal oxygenase domain, which contains a heme, and binding
sites for arginine, tetrahydrobiopterin and CaM. Further the CaM-
binding domain facilitates electron flux from the reductase domain
to the oxygenase domain [28]. The phosphorylation of eNOS allows
active flux of electrons from the reductase domain to NO generation
in the oxygenase domain by guarding the steric hindrance due to
some non-catalytic inserts [29]. eNOS activity can be induced by
several agonists like acetylcholine, bradykinin B2 receptor,
thrombin and ATP, through intracellular elevation of calcium levels
[30]. Studies [1,31] have shown that the contribution of calcium and
substrate/co-factor have been crucial to the activity of eNOS in
intracellular organelles like Golgi, mitochondria, nuclei and plasma
membrane (Figs. 1 and 2).
1.4. Calcium independent eNOS activation
Apart from canonical calmodulin/Ca-dependent eNOS activa-
tion, there is a Ca-independent regulation of eNOS. Substantial
increase in NO release induced by hemodynamic shear stress
appears to be largely independent of intracellular calcium [26].
Shear stress induces eNOS mRNA and protein levels consistent with
eNOS-derived NO, which is important for flow-dependent vasodi-
lation and remodeling of vessels [32,33]. Increase in blood flow
through exercise upregulates eNOS, which is part of the cardio- and
athero-protective phenomenon [34]. It was also reported that
ceramide promotes eNOS activation and enzyme translocation in
endothelial cells [24].
Various studies have reinforced that substantial amount of NO is
produced in unstimulated endothelial cells by a Ca-independent
mechanism [35e37]. Mechanisms like phosphorylation and asso-
ciated regulatory proteins like Hsp90; gp60, Caveolin-1, a major
caveolae-localized protein through caveolae internalization, play
a crucial role in Ca2þ
-independent activation of eNOS and conse-
quent NO production (discussed below).
1.5. eNOS phosphorylation
Protein phosphorylation, a posttranslational modification and
a key regulator of eNOS activity [6,38e41] is modulated by kinases,
phosphatases [42e45] and proteineprotein interactions [21]. The
requisite substrates and cofactors for eNOS function are L-arginine,
tetrahydrobiopterin (BH4), iron (Fe), FMN, FAD and NADPH [45].
Though eNOS activity is coupled to changes in Ca2þ
levels, it is not
the sole factor required for the regulation of the enzyme activity.
The binding of calmodulin (CaM) and the flow of electrons from the
reductase to the oxygenase domain of the enzyme is also depen-
dent on protein phosphorylation and dephosphorylation [1]. The
primary sites where eNOS gets phosphorylated are serine residues
and, to a lesser extent, on tyrosine (Tyr) and threonine residues.
While phosphorylation of Ser617, 635, and 1179 results in the
activation of eNOS, the phosphorylation of Ser116 and Thr497
reduces the eNOS function [44,46e48]. Recently few reports have
shown that tyrosine residues present within oxygenase domain of
eNOS regulate the production of NO, as the modulation of eNOS
activity by tyrosine phosphorylation is supposedly prominent in
primary endothelial cells [49,50].
Tyr81 and Tyr657 are the residues, which gets phosphorylated
due to oxidative-stress and overexpression of v-Src leading to the
regulation of eNOS activity [46]. It appears that Tyr81 mediates
phosphorylation of eNOS through Src mediated mechanism under
different agonists [47]. While the role played by the other Tyr
residue Tyr657 is on the contrary. This particular tyrosine residue,
which is prominent during shear stress works through PYK2/c-Src
in phosphorylating eNOS [48,49,51e53]. It seems to provide
a negative feed back to eNOS phosphorylation for NO production.
However the proper mechanistic role through which the two Tyro-
sine residues target the phosphorylation is an open area to address.
1.6. Kinases involved in eNOS phosphorylation
Calmodulin is one of the very first protein involved in the
regulation of eNOS [54,55]. Increase in intracellular calcium levels
activates CaM that in turn activates CaM kinase II, which phos-
phorylates Ser1179 on eNOS. Ser1177 (human eNOS) and Ser1179
(bovine eNOS) are phosphorylated by the serine/threonine kinase
Akt (protein kinase B) [6,22,23]. Under physiological conditions,
esatcudeResanegyxO Calmodulin
COOH
P T495 /
497
S116 S1177/1179
P
PS615/
617
S633/
635
P P
HDL
Shear stress
Shear stress, VEGF, IGF-1
bradykinin, insulin, estrogen
sphingosine 1-phosphate,
adiponectin, leptin, statins
8-Br-cAMP
Shear stress, VEGF,
bradykinin, statins
8-Br-cAMP
PMA
NH2
Fig. 1. Regulators of eNOS acting on various phosphorylation sites. The eNOS phos-
phorylation sites are numbered according to human and bovine eNOS sequence.
Abbreviations: VEGF, vascular endothelial cell growth factor; 8-Br-cAMP, 8-bromoa-
denosine-30,50-cyclic monophosphate; S-1-P, sphingosine 1-phosphate.
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1187
shear stress, hormones and autacoids activate eNOS by phosphor-
ylation/dephosphorylation mechanisms [22,39,40,46,56e59].
Humoral factors like Bradykinin stimulate NO synthesis by
increasing phosphorylation of multi-sites at eNOS-Ser1177
[45,47,60], eNOS-Ser633 and eNOS-Ser615 [61], and stimulating
dephosphorylation of eNOS-Thr495 [45,47] as well. The phos-
phorylation of Ser1179 is common for multiple diverse signaling
systems, such as insulin and adipokines [56,61,62]. Apart from
these, kinases like Akt [47], and PKA, phosphatases like PP1 and
calcineurin (PP2B) [45] have been implicated in the regulation of
phosphorylation.
2. eNOS e posttranslational modifications
2.1. Myristoylation and palmitoylation
eNOS comprises two types of fatty acid modifications: irre-
versible acylation i.e., myristoylation and reversible palmitoylation.
Myristoylation provides general membrane association, while
palmitoylation directs proteins specifically to the plasma
membrane (PM). The presence of eNOS at the PM may bring eNOS
to interact with the factors that are required for its proper function,
such as arginine, calcium, and cofactor BH4. Proteins during acyl-
ation get depalmitoylated at the PM and are redistributed to other
intracellular membranes [40]. Repalmitoylation occurs specifically
at the Golgi, which further directs the proteins to the PM [63,64].
Studies have demonstrated that the myristoyl moiety is an absolute
requirement for the membrane localization and activity of eNOS
[65,66]. Without this modification, eNOS is almost completely
cytosolic and lacks palmitoyl moieties [67,68]. While myr-
istoylation is an absolute necessity for eNOS membrane activity,
palmitoylation determines the sub-cellular localization of the
eNOS. Palmitoylation is in turn regulated by eNOS activators like
bradykinin [69]. The cycle of depalmitoylation and repalmitoyla-
tion, results in a constant shuttling of the enzyme between Golgi
and PM, redistributing eNOS to distinct membrane microdomains.
2.2. Other interacting proteins of eNOS
eNOS interacts with the proteins like caveolae coat protein
caveolin-1 (Cav-1) and heat shock protein 90 (Hsp90) during its
active and inactive states which regulates its activity. Caveolae are
specialized invaginations of the PM composed of cholesterol,
glycosphingolipids, and some structural proteins, such as caveolin
[70,71]. eNOS is bound to caveolin-1 in caveolae facilitated by
myristoylation and palmitoylation, due to which the enzyme
activity of eNOS is basally repressed [40,72e75]. Caveolin-1 medi-
ated regulation of eNOS through caveolae-mediated endocytosis
gets initiated once the glycoprotein gp60 is activated in endothelial
cells [76]. Maniatis et al., had shown that NO production in
pulmonary endothelial cells is significantly mediated by caveolae
internalization and is independent of the increase of intracellular
Ca2þ
. Compartmentalization of eNOS in caveolae is necessary for its
interaction with regulatory proteins, and calcium- and phosphor-
ylation-dependent signal transduction events that modify the
response of the enzyme to extracellular stimuli. Stimuli, such as
shear stress, induce calcium increase thereby displacing eNOS from
caveolin-1 and activating eNOS by redistributing it from plasma
membrane caveolae [77,78].
The chaperone protein Hsp90 can interact with eNOS and
positively influence its function. The C-terminal Hsp70-interacting
protein (CHIP) interacts with both Hsp70 and Hsp90, and nega-
tively regulates eNOS trafficking into the Golgi complex. By
contrast, the NOS interacting protein (NOSIP) and the NOS traffic
inducer (NOSTRIN) can negatively regulate eNOS localization in the
plasma membrane [28,79e81].
2.3. eNOS localization
eNOS, the key endothelial isoform has a unique sub-cellular
localization pattern. In addition to phosphorylation, the location
of eNOS is important for its activation [74,82,83]. Functional eNOS
was found to be primarily located in the plasma membrane
regions of endothelial cells [84]. The eNOS expression and the site
of its synthesis as well, have a major influence on the biological
activity of the molecule. Complex mechanisms that include the
cell specificity of eNOS expression, and co- and posttranslational
processing lead to trafficking of the enzyme to plasma membrane
caveolae [21].
In the plasma membrane, eNOS is mainly targeted to the cav-
eolae [72,83], where it is inhibited by binding to caveolin-1 (cav-1)
through a consensus site [11,74]. Calcium-calmodulin and intra-
cellular calcium can dissociate eNOS from cav-1, allowing activation
of the enzyme [74,85]. On external stimulus, eNOS shuttles
between caveolae and sub-cellular compartments such as cytosol,
Golgi, and/or perinuclear structures [86,87e91]. This sub-cellular
T495S116 S1177
Atherosclerosis
Coronary Artery Disease
Hyperhomocysteinaemia
Myocardial infarction
Cerebral Ischaemia
Severe subarachnoid haemorrhage
Liver Ischemia
Hypoxia
Diabetes
Portal hypertension
Erectile dysfunction
Hertwig's epithelial root sheath
Alzheimers Disease
Cancer
esatcudeResanegyxO Calmodulin
Hyperhomocysteinaemia
Cerebral Ischaemia
Hypoxia
Diabetes
Erectile dysfunction
Alzheimers Disease
P PP
Atherosclerosis
Coronary Artery Disease
Erectile dysfunction
Hertwig's epithelial root sheath
Alzheimers Disease
Fig. 2. Effect of various pathophysiological and pathological conditions on the phosphorylation sites of eNOS.
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981188
localization of eNOS plays an important role in the regulation of its
activity possibly along with phosphorylation [86e91].
Lipid rafts, which are rich cholesterol and sphingolipid regions
found in surface invaginations of cell, polymerize the caveolins to
form caveoli, which are implicated in endocytosis [92]. Many
studies implicate a coordinated role played by lipid rafts and
caveolins in eNOS sub-cellular signaling [71,78,93]. Studies provide
evidence that lipid rafts are implicated in oxidative-stress signaling
on endothelial cells [93,94]. This information hints a possible
regulatory mechanism to control selective movement of eNOS
signaling by modifying the signaling activity of lipid raft or caveolae,
to counteract and protect the cellular sensitivity to redox stress.
2.4. eNOS phosphorylation in human pathology
2.4.1. Artherosclerosis e endothelial dysfunction
Atherosclerosis is the major cause of chronic vascular diseases
such as coronary artery disease, cerebrovascular disease and
peripheral arterial occlusive disease [95]. Coronary artery disease is
characterized by accumulation of atheromatous plagues in lumen of
arteries which supply nutrients, oxygen to the myocardium. Causes
of the disease are obesity, diabetes, hypertension and smoking.
Upon progress of the disease, the lumen becomes obstructed
completely, leading to myocardial infarction, chronic coronary
ischemia, angina and flash edema. Coronary artery disease starts
with the formation of atherosclerotic lesions. It is characterized by
decreased eNOS activity and NO bioavailability and increased
expression of cell adhesion molecules such as VCAM-1 and ICAM-1
[96e98]. eNOS inhibition has also been shown to accelerate
atherosclerosis suggesting that NO may inhibit several key steps in
the atherosclerotic process. Thus, eNOS could be a candidate gene,
which is implicated in atherosclerosis. Modulation of eNOS activity
by dynamic changes in phosphorylation of eNOS has been of
considerable interest because of its pathophysiological role.
In failing cardiomyocytes Napp et al. showed beta (3) adrenergic
stimulation seems to deactivate rather than activate eNOS. During
stimulation with beta (3) adrenergic agonist BRL improvises
a further dephosphorylation of eNOS (Ser1177) and Akt, while the
treatment increases the Ser114 phosphorylation of eNOS. In
atherosclerosis caused by cytomegalovirus infection Ser1177
phosphorylation was shown to be decreased dramatically in
infected cells [99e102].
Black tea was found to increase NO bioavailability in patients
with artherosclerosis [103]. ERa (estrogen receptor alpha) plays
a key role in mediating the activation of eNOS in response to black
tea polyphenols. In cultured cells, the black tea polyphenolic frac-
tion promotes both eNOS catalytic activity and NO bioactivity. This
effect is because of activation of phosphoinositol 3-kinase (PI3-K)
and Akt via a p38 MAPK-dependent mechanism [104]. This was
supported by Anter et al. 2004 who showed that black tea poly-
phenols induced time-dependent phosphorylation of ERa on
Ser118, which was inhibited by ER antagonist ICI 182,780. Ser118
phosphorylation was observed in response to black tea polyphenols
and mutation of this residue abrogated polyphenol-induced eNOS
activation in COS cells. In addition to S118 phosphorylation, Ser1177
was also found to be phosphorylated by 300% in CAD after 4 weeks
of daily aerobic exercise training compared with sedentary controls
[105]. Therefore two key residues S118 and Ser1177 phosphoryla-
tion were required for regulation of NO production in patients with
coronary artery disease (CAD).
2.4.2. Hyperhomocysteinaemia
Homocysteine produces endothelial injury and stimulates
platelet aggregation. Hyperhomocysteinaemia has been associated
with increased risk of thrombosis and atherosclerosis.
Hyperhomocysteinemia (HHcy) impairs endothelium-dependent
vasodilation by increasing reactive oxygen species, thereby
reducing NO bioavailability [106]. HHcy impairs endothelial func-
tion and eNOS activity, primarily through PKC activation [107].
Robin et al. showed significantly less basal eNOS and phospho-
Ser1179-eNOS/eNOS in mesenteric arteries from HHcy mice but no
difference in phospho-thr495-eNOS/eNOS [108]. Signerello et al.
has shown that PKC stimulates the eNOS phosphorylation of the
negative regulatory residue thr495 and the dephosphorylation of
the positive regulatory site Ser1177 [109].
2.4.3. Myocardial infarction
Myocardial infarction (MI) or acute myocardial infarction (AMI)
occurs due to blockage in coronary artery following rupture of
a vulnerable atherosclerotic plague. In chronic myocardial infarc-
tion infarct size in eNOSÀ/À mice was unchanged but evident
remodeling with less capillary density and hypertrophy accompa-
nied with subsequent systolic and diastolic dysfunction and
increased mortality at 28 days was observed [110], emphasizing the
beneficial effects of eNOS-derived NO on ventricular remodeling
after myocardial infarction. Recent data showed that blocking
mineralocorticoid receptor improved endothelial dysfunction and
oxidative stress by increasing reduced eNOS phosphorylation at
Ser1177 thereby making NO available in experimental myocardial
infarction [111]. Furthermore, in rats with chronic heart failure after
large myocardial infarction [112] eplerenone and metformin [113]
in combination with an angiotensin-converting enzyme inhibitor
increased myocardial eNOS phosphorylation at Ser1177.
2.4.4. Ischemia
Ischemia occurs due to shortage of oxygen, glucose and other
blood born nutrients and is caused due to constriction or blockage
of blood vessels Myocardial infarction, stroke, organ trans-
plantation, and cardiopulmonary bypass causes reperfusion injury
which is leading cause of tissue damage leading to ischemia. In all
these conditions, the initial trigger of the damage is the transient
disruption of the normal blood supply to target organs followed by
reperfusion. From a clinical viewpoint, no therapy is currently
available to limit reperfusion injury, which emphasizes the
importance of a better understanding of its underlying pathological
mechanisms, to devise potential future therapeutic strategies.
2.4.5. Cerebral ischemia
Prominent roles of eNOS and vascular NO in maintaining cere-
bral blood flow and prevention of neuronal injury have been shown
in by pharmacological and genetical approaches in animal models
of cerebral ischemia [114e117]. Vascular NO production protects
against stroke regulates and cerebrovascular perfusion by
increasing collateral flow to the ischemic area. eNOS knockout mice
show decreased blood flow in the ischemic border zone and
develop larger cerebral infarctions [116].
The vascular endothelial growth factor (VEGF) contributes to
activation of eNOS by Ca2þ
/calmodulin and also stimulates the
protein kinase Akt, which directly phosphorylate eNOS on Ser1177
and increases enzyme activity. Increase in Ser1177 phospho-eNOS
occurs in endothelial cells of microvessels after ischemic episodes
with temporal expression of VEGF [118]. Osuka et al. (2004) showed
0.5e2 h transient increase in phospho-Akt at Ser 437 after reper-
fusion, whereas after 6 h there was an elevation of phospho-eNOS
at Ser1177. Hashiguchi et al. (2005) showed in the gerbil hippo-
campal microvasculature transient ischemia model that Ser1177
phosphorylation was unchanged by 24 h after reperfusion, despite
post-ischemic up-regulation of eNOS protein [119]. However,
Thr495 phosphorylation significantly and persistently decreased by
48 h [119].
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1189
Stroke is loss of brain functions due to ischemia caused by
thrombosis or embolism or haemorrhage. Stroke can lead to
permanent neurological damage, complications and death. Risk
factors include advance age, hypertension, transient ischemic
attack, diabetes, high-cholesterol, cigarette smoking and atrial
fibrillation. Stroke can be ischemic stroke or haemorrhagic stroke.
Dmitriy et al. determined the effects of modulation of the Ser1179
phosphorylation site on vascular reactivity, cerebral blood flow, and
outcome in a middle cerebral artery (MCA) occlusion model of
stroke [120]. Their results indicate that modulation of the Ser1179
phosphorylation site affects endothelium-dependent vasodilation
and cerebral blood flow and that these effects determine outcome
of stroke in vivo.
2.4.6. Liver ischemia
Ischemic hepatitis is caused by diffusion ischemia while hepatic
infarction results from hepatic artery disorders. Ischemic hepatitis
is characterized by diffuse liver damage due to an inadequate blood
or O2 supply. Causes are most often systemic e impaired hepatic
perfusion (due to heart failure or acute hypotension), Hypoxemia
(due to respiratory failure or carbon monoxide toxicity) or
increased metabolic demand (due to sepsis).
Recent findings demonstrate that the acute activation of the
serine-threonine kinase Akt is cardioprotective and PI3K/Akt
pathway activation has protective effects on hepatic I/R injury
[121]. Activation of PI3K leads to phosphorylation of membrane
phosphatidylinositol 3,4 biphosphate, which recruits Akt to the
cell membrane leading to phosphorylation and activation of Akt.
Activated Akt in turn promotes eNOS phosphorylation. In the
liver PI3K-Akt-eNOS thus appears to play a central role in pro-
tecting against ischemiaereperfusion (IeR) injury [122,123].
Roviezzo et al. showed activation of the Akt pathway in ischemic
regions of reperfused ileum through an increased S1179 phos-
phorylation in reperfused intestinal tissue coupled to Akt
activation [124].
2.4.7. Severe subarachnoid haemorrhage
Severe subarachnoid haemorrhage (SAH) is a form of stroke that
induces dysfunction of endothelial nitirc oxide synthase (eNOS),
resulting in severe vasospasm. Clinically, however, some portions of
cerebral arteries may show only mild vasospasm. Although severe
vasospastic arteries after SAH have been intensively studied,
activity of eNOS with the mild form of the disease has received less
attention. Osuka et al. 2004 showed that SAH induces a temporary
activation of AMPK alpha, which phosphorylates eNOS at Ser1177 in
endothelial cells of mild vasospastic basilar arteries. This signal
transduction might play an important role in controlling cerebral
blood flow after SAH [118].
2.4.8. Hypoxia
Hypoxia is a pathological condition in which body as a whole or
a region of the body is deprived of adequate oxygen supply. Chronic
hypoxia increases endothelial nitric oxide synthase (eNOS)
production of nitric oxide (NO) and cardioprotection in neonatal
rabbit hearts [125]. The effect of hypoxia on eNOS expression
remains controversial. The presence of inflammatory mediators
and cytokines under hypoxic conditions further decreases expres-
sion of eNOS and production of NO [126].
Ser1177 phosphorylation under hypoxia was observed by
a number of workers. Shi et al. 2002 showed that association of
Hsp90 with eNOS is important for increasing NO production by
Ser1177 phosphorylation and limiting eNOS-dependent superoxide
anion generation. Hsp90 is associated with eNOS and the extent to
which the enzyme is activated is based on phosphorylation of eNOS
at Ser1177 [126].
Additionally, hypoxic conditions attenuate Akt-mediated phos-
phorylation at Ser1177, alter calcium metabolism, and alter the
balance of eNOS proteineprotein interactions with caveolin and
calmodulin [127]. Coulet et al. 2003 showed that human eNOS is
a hypoxia-inducible gene, whose transcription is stimulated
through HIF-2 interaction with two contiguous sites located at
À5375 to À5366 of the human eNOS promoter [128]. Liu J et al.
2009 have shown that phosphorylation of Ser1177 in eNOS
decreased, whereas phosphorylation of Thr495 increased, in the
prenatally hypoxic pulmonary arteries [129]. These data demon-
strate that prenatal hypoxia results in persistent abnormalities in
endothelium-dependent relaxation responses of pulmonary
arteries in adult sheep due to decreased eNOS activity resulting
from altered posttranslational regulation.
2.4.9. Diabetes
In diabetes, the body either fails to properly respond to its own
insulin, does not make enough insulin, or both. This causes glucose
to accumulate in the blood, often leading to various complications
[130,131]. Endothelial dysfunction is a critical initial factor 2 in the
development of diabetic vascular disease [132]. Endothelial
dysfunction, as represented by impaired endothelium functions,
and NO-mediated relaxation, occurs in cellular and experimental
models of diabetes [133e136]. Similarly, many, but not all, clinical
studies have found that endothelium-dependent vasodilation is
abnormal in patients with type 1 or type 2 diabetes [137e140].
Thus, decreased levels of NO in diabetes may underlie its athero-
genic predisposition.
The defect of Akt/eNOS signaling may play a primary role in
endothelial dysfunction in type 2 diabetes mellitus. In aorta from
diabetic animals, as well as in type 2 diabetic patients, Akt/eNOS
phosphorylation has been shown to be decreased [141,142], at least
partly via hyperglycemia-induced GlcNAc modification of the
enzyme, which may explain the development and progression of
diabetes-associated atherosclerosis [141]. Insulin also mediates its
effects through binding to insulin receptors and triggering down-
stream signaling pathways, of which the most important is the
phosphatidylinositol 3-kinase (PI3K)-Akt/protein kinase B (Akt/
PKB) pathway. This pathway is involved in a variety of insulin
responses including transport of glucose through cell membranes,
myocardial survival, and anti-apoptotic effects in endothelial cells
[143,144]. As such, PI3K-Akt pathway activation by eNOS-derived
NO may result in improved endothelial function and rescue of
impaired myocardial cells [145,146]. Recent studies on eNOS gene
disruption studies in mice revealed that, this leads to insulin
resistance, resulting in hypertension and hyperlipidemia
[147e149]. Further biochemical studies in insulin-responsive cells
have uncovered phosphorylation-dependent signaling role in
insulin-stimulated activation of eNOS [150,151].
In diabetic rat model, Biljana Musicki et al. showed that eNOS
dysfunction in the penis in diabetes by O-GlcNAc modification
affecting phosphorylation at Ser1177 residue, and rendering it
incapable of activation by normal fluid shear stress stimuli and
VEGF signaling [152]. Such eNOS impairment could contribute to
erectile dysfunction and affect long-term penile health in diabetes.
In diabetic liver Elrod et al. showed that although there was no
difference in total hepatic eNOS protein between non-diabetic and
diabetic animals, they discovered decreased phosphorylation of
both Thr495 and Ser1177 residues in diabetic mice as compared
with non-diabetic control animals [153]. In another study of dia-
betic animals [141] it has also been shown that this regulatory site
is subject to posttranslational modification, thereby decreasing
eNOS activity.
NO-based therapies have been proved by numerous investiga-
tions in various animal models [154e156]. Evidence also suggest
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981190
that NO is involved in the diabetic pathology [157]. In Type II dia-
betes mellitus (DM) where neointimal hyperplasia conditions lead
to vascular interventions, external delivery of NO proved to recover
arterial injury from neointimal hyperplasia [158,159]. Uncoupling
of eNOS has been demonstrated in animal models of diabetes [160].
Together, these data indicate that diabetes and insulin resistance
are characterized, at least in part, by endothelial dysfunction and
potentially by altered eNOS expression and NO production.
Insulin treatments have stimulated phosphorylation of eNOS
and protein kinase B (Akt) in arteries of diabetic mice [161]. We can
interpret from these studies that activation or inhibition of eNOS
would rectify diabetes pathology. We presume that phosphoryla-
tion of eNOS play key role in rectifying the defects under this
diseased condition and further sites involved in eNOS phosphory-
lation under these conditions should be identified and extensively
studied.
2.4.10. Oxidative stress
Oxidative stress, a pathological imbalance due to anomalous
reactive oxygen species (ROS) production, is a common feature of
many of the pathological conditions like cardiovascular anomalies,
hypertension, diabetes, atherosclerosis, cancer, which lead to
endothelial dysfunction [162]. ROS production involves inactivation
of the signaling molecule nitric oxide (NO), leading to endothelial
dysfunction. Uncoupled endothelial nitric oxide synthase (eNOS)
could be another important source of ROS in the endothelium.
Uncoupling of eNOS is now recognized as important in the patho-
physiology of several cardiovascular disorders and a predictor of
future adverse vascular events such as hypertension, atheroscle-
rosis and diabetes [163,164]. Under deficiency of the NOS cofactor
tetrahydrobiopterin (BH4), the enzyme can become uncoupled and
generate superoxide instead of NO. Furthermore, BH4 itself is prone
to degradation by oxidation, which leads to further amplification of
ROS production due to NOS uncoupling [164,165]. (ONOOÀ
)
oxidizes BH4 and reduces the availability of BH4 resulting in eNOS
uncoupling, thereby increased superoxide and decreased NO
production [166,167].
Role of ROS in oxidative stress under pathological conditions
lead by decoupling of eNOS and consequent reduction of NO
production is very much evident from the literature [165,166]. It
has been studied that Thr495 dephosphorylation has been impli-
cated in the uncoupling of eNOS resulting to increased superoxide
(O2
À
) stress instead of NO, which has resulted in atherosclerosis
[168]. But the increased NO production even under Thr495
dephosphorylation under inducers like bradykinin is still a point of
study [44,169,170]. It has to be explored further as such why Thr495
dephosphorylation behaving in a different manner under various
conditions of pathology and physiology, and is there any feed back
mechanism from this event to the NO signaling system?
Tyrosine phosphorylation is a novel mechanism of eNOS regu-
lation, which can affect the activity of eNOS and also influence the
distribution of eNOS in the cellular compartments. Tyrosine may
interact with other intercellular regulatory proteins like caveolin-1
that may influence the activity of eNOS [40]. However, studies have
to be carried out to elucidate the proper mechanism of tyrosine
phosphorylation and its interaction with other proteins in regu-
lating eNOS activity and thereby NO production, which may
provide a valuable information in therapeutic studies related to
vascular anomalies.
2.4.11. Oxidative stress in type II diabetes
Superoxide production contributes to atherogenesis as evi-
denced by increased superoxide anion content in vessels from
animals fed with high-cholesterol diets [171]. Superoxide reduces
expression of eNOS protein, as well as decrease the number of
caveolae in endothelial cells [172]. The most significant source of
superoxide in the vascular wall is NADPH oxidase [173]. The
enzyme superoxide dismutase acts to detoxify superoxide and
releases hydrogen peroxide (H2O2). H2O2 induces phosphorylation
of eNOS-Ser1177 and dephosphorylation of eNOS-Thr495, by
tyrosine kinaseedependent PI3K/Akt mechanisms [174]. Oxidative
stress plays an important role in type 2 diabetes-related endothelial
dysfunction. Zhang H et al. 2009 showed that Resveratrol restored
endothelial function in type 2 diabetes by inhibiting TNF-alpha-
induced activation of NAD(P)H oxidase and preserving eNOS
phosphorylation (Ser1177) [175].
2.4.12. Hypertension
Hypertension is a risk factor for all clinical manifestations of
atherosclerosis since it is a risk factor for atherosclerosis itself
[176e180]. It is an independent predisposing factor for heart failure
[181], coronary artery disease [182e184] stroke, renal disease
[185e187], and peripheral arterial disease [188,189]. Cirrhosis is
one of the main causes of portal hypertension. Adrenal cortical
abnormalities, kidney diseases, neuroendocrin tumors are also
known to cause secondary hypertension. Defect in eNOS phos-
phorylation has been accounted for endothelial dysfunction further
leading to hypertension, hyperlipidemia, and other diseased
conditions. Atochin et al. has shown modulation of S1179 phos-
phorylation as an approach for treating cardiovascular diseases,
particularly influenced by diabetes, obesity, metabolic syndrome,
hyperlipidemia, and hypertension [120]. However the effects of
phosphorylation sites of eNOS apart from S1179 on vascular path-
ogenesis have to be further studied in detail.
2.4.13. Portal hypertension and liver cirrhosis
Normally, blood from the intestines and spleen is carried to the
liver through the portal vein. But cirrhosis slows the normal flow of
blood, which increases the pressure in the portal vein resulting in
the condition called portal hypertension. Overproduction of
vascular NO plays a central role in both systemic and splanchnic
vasodilatation, which is a hallmark of portal hypertension
[190e193]. In the chronic model of portal hypertension increased
eNOS expression and enzyme activity are well-established events
[191e194]. However, the mechanism of the early induction of
excessive NO production by eNOS remains to be elucidated. The
phosphorylation of eNOS by Akt activates the enzyme and may be
the first step in increasing NO production in portal hypertension.
Iwakiri et al. showed that the phosphorylation of eNOS at Ser1176
was significantly increased in the PVL group [193]. Furthermore,
PVL significantly increased Akt phosphorylation (an active form of
Akt). When vessels were treated with wortmannin (10 nM) to block
the PI3K/Akt pathway, NO-induced vasodilatation was significantly
reduced.
2.4.14. Erectile dysfunction
Erectile dysfunction (ED) is defined as consistent inability to
obtain or maintain an erection for satisfactory sexual intercourse.
ED is predominantly the disease of vascular origin. The incidence of
ED dramatically increases in men with diabetes mellitus, hyper-
cholesterolemia, and cardiovascular disease. Loss of the functional
integrity of the endothelium and subsequent endothelial dysfunc-
tion plays an integral role in the occurrence of ED in this cohort of
men [195].
Age related ED is associated with eNOS activation through
dysregulation of its phosphorylation. PI3-K/Aktedependent eNOS
activation has recently been shown in the penis. In rats and mice,
both neuro- and agonist-induced penile erection produced rapid
increase in Akt and eNOS phosphorylation at Ser1177 in the penis,
which remain elevated after the termination of the initial stimulus
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1191
[196]. In another study Akt dependent phosphorylation of a posi-
tive regulatory site Ser1177 on eNOS was shown to be decreased
whereas phosphorylation of a negative regulatory site thr495 on
eNOS increased in the aged rat penis. In diabetic rat penis decreased
eNOS phosphorylation on Ser1177 caused by O-linked N-acetyl-
glucoseamine (O-GlcNAc) modification of eNOS.
2.4.15. eNOS and human aging
Hidetaka Ota et al. 2008 showed the phenomenon of human
aging is known to be a critical cardiovascular risk factor [197].
Cellular senescence of endothelial cells has been proposed to be
involved in endothelial dysfunction and atherogenesis [198]. The
lesions of human atherosclerosis have been extensively studied
histologically, and these studies have demonstrated that there are
vascular cells that exhibit the morphological features of cellular
senescence [199]. Although there is substantial evidence demon-
strating an aging-associated development of cardiac and vascular
dysfunction, the mechanisms responsible for this phenomenon
have not yet been clearly established. Aging is associated with
erectile dysfunction (ED) attributed to reduced nitric oxide syn-
thase (NOS) activity and nitric oxide bioavailability. A lot of work
has been carried out on endothelial cellular senescence [200e202].
Endothelial NO can protect against a state of oxidative stress,
and activation of eNOS and subsequent production of NO delay
endothelial cellular senescence [200,201]. A PDE3 inhibitor, cil-
ostazol, used as a vasodilating antiplatelet drug for treating inter-
mittent claudication, and in preclinical studies was shown to have
a protective effect on endothelial cells by increasing eNOS activity
[202]. Cilostazol increases intracellular cAMP content accordingly
and activates protein kinase A (PKA) or PI3K/Akt signaling [203]. In
yeast, Sir2 (silent information regulator-2) has been identified as an
NADþ
-dependent histone deacetylase [204]. Mammalian sirtuin 1
(Sirt1), the closest homolog of Sir2, regulates the cell cycle, senes-
cence, apoptosis, and metabolism, by interacting with a number of
molecules, including p53, PML, and PPAR-&UnknownEntity;
[205e207]. A recent study showed that production of NO by caloric
restriction increases Sirt1 expression and suggested that eNOS may
be involved in regulating the expression of Sirt1 in murine white
adipocytes [208]. HUVECs treated with other cAMP-elevating
agents and DETA-NO showed a reduction of SA-bgal-positive cells
as well [209]. Cilostazol increased phosphorylation of Akt at Ser473
and of endothelial nitric oxide synthase (eNOS) at Ser1177, with
a dose-dependent increase in Sirt1 expression. Moreover, the effect
of cilostazol on premature senescence was abrogated through
inhibition of Sirt1 [210].
2.4.16. eNOS phosphorylation and Hertwig’s epithelial root sheath
Epithelial cell rests of Malassez (ERM) are quiescent epithelial
remnants of Hertwig’s epithelial root sheath (HERS) that are
involved in the formation of tooth roots. After completion of crown
formation, HERS are converted from cervical loop cells, which have
the potential to generate enamel for tooth crown formation.
Korkmaz 2005 has shown that low concentrations of NO produced
by nNOS, and eNOS activate intracellular soluble guanylate cyclase
(sGC) to produce intracellular cyclic guanosine 30:50-mono-
phosphate (cGMP), which triggers rapid cellular responses such as
cell proliferation, cell differentiation, and apoptosis under physio-
logical conditions [211]. He later showed the basal production of NO
by eNOS in the ERM is modulated by phosphorylation of eNOS at
Ser1177 and Ser116 residues, while the basal activity of the eNOS is
not influenced by phosphorylation of eNOS at Thr495 residue.
2.4.17. Alzheimers disease
Alzheimer’s disease (AD) is associated to a cerebral amyloid
angiopathy with dysregulation of cerebral blood flow (CBF).
Vascular risk factors (VRFs) are probably determinants of
incidence and course of AD. b-amyloid peptides (AbP), the main
component of senile plaques typical of AD, circulate in soluble
globular form in bloodstream. Interestingly, AbP is able to induce
endothelial dysfunction, and this effect may represent the
link between vascular and neuronal pathophysiological factors
involved in AD. Gentile et al. 2004 showed that AbP enhances
eNOS phosphorylation on thr495 and Ser116 and reduces
acetylcholine-induced phosphorylation on Ser1177 [212]. Such an
effect depends on a PKC signaling pathway, as suggested by its
phosphorylation on Ser 660.
2.4.18. RBC-eNOS
RBC-NOS-generated NO contributes to the overall NO pool and
also play prominent role in regulation of blood flow and platelet
Table 1
List of publications showing the effect of different diseases on the phosphorylation status of eNOS.
Phosphorylation site Origin Mode of induction/
inhibition
References
Ser1177; Thr495 Hypoxic pulmonary arteries of sheep Chronic hypoxia Liu J, et al./Am. J. Physiol. Lung Cell. Mol. Physiol. 2009;296(3):L547e54.
Ser1177 Diabetic Mice Aortic EC Diabetes Zhang H, et al./Arterioscler. Thromb. Vasc. Biol. 2009;29(8):1164e71.
Ser1177 Diabetic animal’s aorta Telmisartan Wenzel P, et al./Free Radic. Biol. Med. 2008;45(5):619e26.
Ser1177 Aortas from DOCA rats & unineph-
rectomized rats
PugNAc Lima VV, et al./Hypertension 2009;53(2):166e74.
Ser1177 Ischemic myocardium affected rats WY-14643 Bulhak AA, et al./Am. J. Physiol. Heart Circ. Physiol. 2009; 296(3):H719e27.
Ser1177; Thr495 Mouse aorta Tacrolimus Cook LG, et al./Kidney Int. 2009;75(7):719e26.
Ser1177; Ser114 eNOS in failng myocardium BRL37344 Napp A, et al./J Card Fail. 2009;15(1):57e67.
Ser1177 HAECs TNF-alpha-neutralizing
antibody
Yuen DY, et.al/Diabetes 2009 May;58(5):1086e95.
Ser1177; Thr495 Recombinant BH(4)-free eNOS VEGF/PKC Chen CA, et.al/J. Biol. Chem. 2008 Oct 3;283(40):27038e47.
Ser1177; Thr495 Renal vascular -eNOS Radicicol Ramírez V, et.al/Am. J. Physiol. Renal Physiol. 2008;295(4):F1044e51.
Ser1177 ECV-D8eNOSGFP and CVEC PAF Sánchez FA, et.al/Proc. Natl. Acad. Sci. U. S. A. 2009 21;106(16):6849e53.
Ser1179 Cerebral artery Iscemic stroke Dmitriy et al./Clin. Invest. 2007;117:1961e7.
Ser1177 Forebrain Cerebral ischemia Osuka et al./Stroke 2004; (35) 2582
Ser1177 Mice Myocardial infarction NOS3 deficiency Scherrer-Crosbie et al./Circulation 2001;104:1286e1291
Ser1177 Diabetic myocardial tissue Eplerenone and metformin Calvert et al./Diabetes 2008;57:696e705
Ser1177 Human dermal microvascular
endothelial cells
HIF-1, HIF-2 Coulet et al./J. Biol. Chem. 2003;278: 46230e46240.
Ser1177; Ser116;
Thr495
Vessels of cerebral amyloid aorta AbP Gentile et al./J. Biol. Chem. 2004; 279: 48135e48142
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981192
aggregation [213,214]. RBCs play a role in the pathogenesis of
hypertension and stroke [215,216]. Recently, functionally active
endothelial type NO synthase was discovered in mature murine
and human red blood cells (RBC-eNOS). Nikolaev et al. found that
the treatment of mouse erythrocytes with rHuEpo resulted in
a time- and dose-dependent up-regulation of NO production,
mediated via activation of the PI3K/Akt pathway and RBC-eNOS
phosphorylation at Ser1177 [217]. Red blood cell (RBC)-derived NOS
has common but also distinct regulatory mechanisms when
compared with eNOS [218], which depends on intracellular Ca2þ
level and phosphorylation at Ser1177, regulated by PI3K [215]. Suhr
et al. 2009 has investigated the influence of intensive exercise on
eNOS content and the phosphorylation states of the eNOS at Ser116,
Ser1177, and Thr495 in human erythrocytes [219].
Evidence from the literature suggests that in vivo flow dynamics
are more affected by RBC aggregation on to endothelial cells leading
to pathological conditions [220]. RBC adhesiveness and aggregation
have proved to be useful markers for detecting vascular inflam-
mation and atherosclerosis progression in patients with coronary
artery disease [221]. As RBC-derived NO significantly contributes to
the intravascular NO pool, a lack of NOS activity and eNOS phos-
phorylation seems to be likely in these diseases. Rosuvastatin has
shown an improved RBC-eNOS activity through phosphorylation of
eNOS at Ser1177 residue. This improvement resembled the in vivo
situation of RBCs passing through capillaries [222]. RBC-NOS will
serve as a key factor for fields such as atherosclerosis, microcircu-
latory diseases, RBC aging and storage and adaptation mechanisms
for high altitude [218].
2.4.19. eNOS in cancer
eNOS has been detected in tumour cells [223], and catalyzes the
synthesis of NO, which can facilitate S-nitrosylation of the thiol
group of cysteines (Cys) in proteins [224], such as that of Cys118 of
HRas, which enhances the dissociation of guanine nucleotides
thereby increasing GTP-bound HRas [225]. Wildtype Ras proteins
can be required for activation of the MAPK pathway by oncogenic
Ras [226], and membrane-targeting of RasGAP, which inhibits
wildtype but not oncogenic Ras, reverts oncogenic Ras trans-
formation of NIH3T3 cells [227], suggesting that wildtype Ras
proteins may facilitate oncogenic signaling. Kian-Huat Lim et al. has
shown that AKT activation of eNOS maintains tumour growth in the
absence of oncogenic Ras by activating wildtype Ras through
S-nitrosylation of Cys118.
Activated KRas and Ser1177 phosphorylated eNOS were also
elevated in the tumour specimens compared to matched and
unmatched normal tissue controls [228], with the caveat that
biopsies also contain stromal tissue that could contribute to
detected eNOS phosphorylation.
3. Therapeutic challenges
Phosphorylation of eNOS is a key mechanism responsible for
eNOS activity and subsequent NO production [22,23,27,39,41].
Anomalies like cardiovascular dysfunctions, erectile dysfunctions,
stroke and several vascular abnormalities have been implicated for
erratic NO production. In the recent years eNOS phosphorylation
has been a potential target as a new therapeutic area of interven-
tion in many pathological situations, to develop a way of promoting
NO production [120,229]. Dmitriy et al. has shown in vivo that
eNOS phosphorylation is an important determinant of vascular
function, blood flow, and cerebral ischemia. This work also provides
proof of concept for the modulation of Ser1179 phosphorylation as
an approach to prevent cardiovascular disease, particularly influ-
enced by risk factors of diabetes, metabolic syndrome, hyperlipid-
emia, and hypertension.
Recent advances have shown that pharmacological agents like
Raloxifene, Fasudil and Y27632 improve cardiovascular function
[231e234]. Studies by Leung et al. shows that the therapeutic
concentrations of raloxifene (1e3 nM) augment endothelial func-
tions through up-regulation of eNOS activity by increased eNOS
phosphorylation in porcine coronary arteries [230]. Y-27632 and
fasudil, are some of the emerging drugs the physiological role of
Rho-Kinases (ROCKs), particularly in cardiovascular disease
[232e234]. Inhibition of RhoA or ROCKs leads to the rapid activa-
tion of PI3K/Akt and phosphorylation of eNOS [235,236], suggest-
ing the potential role of ROCKs in regulating eNOS activation in
addition to eNOS expression. Fasudil and Y-27632 were shown to
be effective Rho-kinase inhibitors for the treatment of a wide range
of cardiovascular disease, including cerebral and coronary vaso-
spasm, angina, hypertension, pulmonary hypertension, and heart
failure. They rapidly increase the endothelial eNOS activity through
phosphorylation and exert cardiovascular protection [234,236].
Rho/Rho-kinase pathway plays an important role in various intra-
cellular functions that are involved in the pathogenesis of cardio-
vascular disease [237].
4. Future challenges
The pattern of eNOS phosphorylation and dephosphorylation
later evolved as exceedingly complex. The challenge for future
studies will be to examine the relative contribution of each regu-
latory site on both the level and the time course of NO production.
Hopefully, this may help to design smarter eNOS constructs that, on
delivering in cardiovascular tissue, would drive NO release where
and when required. RBC-eNOS is another emerging area of
research, which is gaining prominence for fields like atheroscle-
rosis, vasculocirculatory diseases, RBC aging and storage and high
altitude pathophysiology. Novel diagnostic approaches for quanti-
fying RBC-eNOS and identification of phosphorylation dysfunction
under diseased condition help in the development of new diag-
nostic and therapeutic strategies for these diseases. Regulatory
mechanisms of various phosphorylation sites of eNOS, are to be
further elucidated. Interactive studies of these phosphorylation
residues with other regulatory protein of eNOS have to be carried
out extensively. This may provide valuable information in thera-
peutics of vascular anomalies related to eNOS/NO dysfunctions.
Acknowledgement
This work was supported by a grant from KBC-RF. The authors
also acknowledge Ms. Puja Kumari for compilation of the infor-
mation in Table 1.
References
[1] I. Fleming, R. Busse, Molecular mechanisms involved in the regulation of the
endothelial nitric oxide synthase. Am. J. Physiol. Regul. Integr. Comp. Physiol.
284 (2003) 1e12.
[2] J.A. Panza, C.E. Garcia, C.M. Kilcoyne, A.A. Quyyumi, R.O. Cannon, Evidence
that endothelial dysfunction in patients with hypercholesterolemia is not
due to increased extracellular nitric oxide breakdown by superoxide anions.
3rd Circul. 91 (1995) 1732e1738.
[3] J.P. Cooke, D.W. Losordo, Nitric oxide and angiogenesis. Circulation 105
(2002) 2133e2135.
[4] K.M. Naseem, The role of nitric oxide in cardiovascular diseases. Mol. Aspect.
Med. 26 (2005) 33e65.
[5] P. Kubes, M. Suzuki, D.N. Granger, Nitric oxide: an endogenous modulator of
leukocyte adhesion. Proc. Natl. Acad. Sci. U.S.A. 88 (1991) 4651e4655.
[6] D.J. Lefer, S.P. Jones, W.G. Girod, et al., Leukocyte-endothelial cell interactions
in nitric oxide synthase-deficient mice. Am. J. Physiol. 276 (1999)
1943e1950.
[7] S.P. Jones, W.G. Girod, A.J. Palazzo, D.N. Granger, M.B. Grisham,
D. Jourd’Heuil, et al., Myocardial ischemia-reperfusion injury is exacerbated
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1193
in absence of endothelial cell nitric oxide synthase. Am. J. Physiol. 276 (1999)
1567e1573.
[8] P. Heeringa, H. van Goor, Y. Itoh-Lindstrom, N. Maeda, R.J. Falk, K.J. Assmann,
et al., Lack of endothelial nitric oxide synthase aggravates murine accelerated
anti-glomerular basement membrane glomerulonephritis. Am. J. Pathol. 156
(2000) 879e888.
[9] V. Malhotra, L. Orci, B.S. Glick, M.R. Block, J.E. Rothman, Role of an N-ethyl-
maleimide-sensitive transport component in promoting fusion of transport
vesicles with cisternae of the Golgi stack. Cell 54 (1988) 221e227.
[10] S.C. Tai, G.B. Robb, P.A. Marsden, Arterioscler. Thromb. Vasc. Biol. 24 (2004)
405e412.
[11] D. Fulton, J. Fontana, G. Sowa, J.P. Gratton, M. Lin, K.X. Li, B. Michell, B.E. Kemp,
D. Rodman, W.C. Sessa, Localization of endothelial nitric-oxide synthase
phosphorylated on serine 1179 and nitric oxide in Golgi and plasma membrane
defines the existence of two pools of active enzyme. J. Biol. Chem. 277 (2002)
4277e4284.
[12] T. Michel, O. Feron, Nitric oxide synthases: which, where, how, and why?
J. Clin. Invest. 100 (1997) 2146e2152.
[13] W.K. Alderton, C.E. Cooper, R.G. Knowles, Nitric oxide synthases: structure,
function and inhibition. Biochem. J. 357 (2001) 593e615.
[14] C. Nathan, Q.-W. Xie, Nitric oxide synthase: roles, tolls and controls. Cell 78
(1994) 915e918.
[15] A. Papapetropoulos, R.D. Rudic, W.C. Sessa, Molecular control of nitric oxide
synthases in the cardiovascular system. Cardiovasc. Res. 43 (1999) 509e520.
[16] Y. Wang, P.A. Marsden, Nitric oxide synthases: gene structure and regulation.
Adv. Pharmacol. 34 (1995) 71e90.
[17] U. Forstermann, J.P. Boissel, H. Kleinert, Expressional control of the “consti-
tutive” isoforms of nitric oxide synthase (NOS I and NOS III). FASEB J. 12
(1998) 773e790.
[18] C. Nathan, Q.W. Xie, Regulation of biosynthesis of nitric oxide. J. Biol. Chem.
269 (1994) 3725e3728.
[19] A. Papapetropoulos, G. García-Cardeña, J.A. Madri, W.C. Sessa, Nitric oxide
production contributes to the angiogenic properties of vascular endothelial
growth factor in human endothelial cells. J. Clin. Invest. 100 (1997)
3131e3139.
[20] T. Simoncini, G. Varone, L. Fornari, P. Mannella, M. Luisi, F. Labrie, A.R. Genazzani,
Genomic and nongenomic mechanisms of nitric oxide synthesis induction in
endothelial cells by a fourth generation selective estrogen modulator. Endo-
crinology 143 (2002) 2052e2061.
[21] D. Fulton, J.P. Gratton, W.C. Sessa, Post-translational control of endothelial
nitric oxide synthase: why isn’t calcium/calmodulin enough? J. Pharmacol.
Exp. Ther. 299 (2001) 818e824.
[22] D. Fulton, P. Gratton, T.J. McCabe, J. Fontana, Y. Fujio, K. Walsh, T.F. Franke,
A. Papapetropoulos, W.C. Sessa, Regulation of endothelium derived nitric
oxide production by protein kinase Akt. Nature 399 (1999) 597e601.
[23] S. Dimmeler, I. Fleming, B. Fisslthaler, C. Hermann, R. Busse, A.M. Zeiher,
Activation of nitric oxide synthase in endothelial cells by Akt-dependent
phosphorylation. Nature 399 (1999) 601e605.
[24] J. Igarashi, H.S. Thatte, P. Prabhakar, D.E. Golan, T. Michel, Calcium-inde-
pendent activation of endothelial nitric oxide synthase by ceramide. Proc.
Natl. Acad. Sci. U.S.A. 96 (1999) 12583e12588.
[25] M.A. Marletta, Nitric oxide synthase: aspects concerning structure and
catalysis. Cell 78 (1994) 927e930.
[26] M.J. Kuchan, J.A. Frangos, Role of calcium and calmodulin in flow-induced
nitric oxide production in endothelial cells. Am. J. Physiol. Cell Physiol. 266
(1994) C628eC636.
[27] I. Fleming, J. Bauersachs, B. Fisslthaler, R. Busse, Ca2þ
-independent activation
of the endothelial nitric oxide synthase in response to tyrosine phosphatase
inhibitors and fluid shear stress. Circ. Res. 82 (1998) 686e695.
[28] W.C. Sessa, eNOS at a glance. J. Cell Sci. 17 (2004) 2427e2429.
[29] T.J. McCabe, D. Fulton, L.J. Roman, W.C. Sessa, Enhanced electron flux and
reduced calmodulin dissociation may explain “calcium-independent” eNOS
activation by phosphorylation. J. Biol. Chem. 275 (2000) 6123e6128.
[30] J. Mombouli, P. Vanhoutte, Kinins and endothelial control of vascular smooth
muscle. Annu. Rev. Pharmacol. Toxicol. 35 (1995) 679e705.
[31] D. Jagnandan, W.C. Sessa, D. Fulton, Intracellular location regulates calcium-
calmodulin-dependent activation of organelle-restricted eNOS. Am. J. Phys-
iol. Cell Physiol. 289 (2005) 1024e1033.
[32] W.C. Sessa, K. Pritchard, N. Seyedi, et al., Chronic exercise in dogs increases
coronary vascular nitric oxide production and endothelial cell nitric oxide
synthase gene expression. Circ Res. 74 (1994) 349e353.
[33] E.M. Prence, Effects of calcium on phosphatidylserine- and saposin C-stim-
ulated glucosylceramide beta-glucosidase activity. Biochem. J. 310 (1995)
571e575.
[34] D. Gattullo, P. Pagliaro, N.A. Marsh, G. Losano, New insights into nitric oxide
and coronary circulation. Life Sci. 65 (1999) 2167e2174.
[35] A. Mülsch, E. Bassenge, R. Busse, Nitric oxide synthesis in endothelial cytosol:
evidence for a calcium-dependent and a calcium-independent mechanism.
Naunyn Schmiedeberg’s Arch. Pharmacol. 340 (1989) 767e770.
[36] K. Ayajiki, M. Kindermann, M. Hecker, I. Fleming, R. Busse, Intracellular pH
and tyrosine phosphorylation but not calcium determine shear stress-
induced nitric oxide production in native endothelial cells. Circ. Res. 78
(1996) 750e758.
[37] I. Fleming, J. Bauersachs, R. Busse, Calcium-dependent and independent
activation of the endothelial NO synthase. J. Vasc. Res. 34 (1997) 165e174.
[38] T. Michel, Phosphorylation and subcellular translocation of endothelial nitric
oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 90 (1993) 6252e6256.
[39] M.A. Corson, N.L. James, S.E. Latta, R.M. Nerem, B.C. Berk, D.G. Harrison,
Phosphorylation of endothelial nitric oxide synthase in response to fluid
shear stress. Circ Res. 79 (1996) 984e991.
[40] G. Garcia-Cardena, R. Fan, D.F. Stern, J. Liu, W.C. Sessa, Endothelial nitric
oxide is regulated by tyrosine phosphorylation and interacts with caveolin-1.
J. Biol. Chem. 271 (1996) 27237e27240.
[41] B. Gallis, G.L. Corthals, D.R. Goodlett, H. Ueba, F. Kim, S.R. Presnell, D. Figeys,
D.G. Harrison, B.C. Berk, R. Aebersold, M.A. Corson, Identification of flow
dependent endothelial nitric-oxide synthase phosphorylation sites by mass
spectrometry and regulation of phosphorylation and nitric oxide production
by the phosphatidylinositol 3-kinase inhibitor LY294002. J. Biol. Chem. 274
(1999) 30101e30108.
[42] E. Butt, A. M.BernhardtSmolenski, P. Kotsonis, L.G. Frohlich, A. Sickmann,
H.E. Meyer, S.M. Lohmann, H.H. Schmidt, Endothelial nitric-oxide synthase
(type III) is activated and becomes calcium independent upon phosphory-
lation by cyclic nucleotide-dependent protein kinases. J. Biol. Chem. 275
(2000) 5179e5187.
[43] Z.P. Chen, K.I. Mitchelhill, B.J. Michell, D. Stapleton, I. Rodriguez-Crespo,
L.A. Witters, D.A. Power, P.R. Ortiz de Montellano, B.E. Kemp, AMP-activated
protein kinase phosphorylation of endothelial NO synthase. FEBS Lett. 443
(1999) 285e289.
[44] I. Fleming, B. Fisslthaler, S. Dimmeler, B.E. Kemp, R. Busse, Phosphorylation of
Thr495 regulates Ca2þ
/calmodulin-dependent endothelial nitric oxide syn-
thase activity. Circ. Res. 88 (2001) 68e75.
[45] B.J. Michell, J.E. Griffiths, K.I. Mitchelhill, I. Rodriguez-Crespo, T. Tiganis,
S. Bozinovski, P.R. de Montellano, B.E. Kemp, R.B. Pearson, The Akt kinase
signals directly to endothelial nitric oxide synthase. Curr. Biol. 9 (1999)
845e848.
[46] B.J. Michell, Z. Chen, T. Tiganis, D. Stapleton, F. Katsis, D.A. Power, A.T. Sim,
B.E. Kemp, Coordinated control of endothelial nitric-oxide synthase phos-
phorylation by protein kinase C and the cAMP-dependent protein kinase.
J. Biol. Chem. 276 (2001) 17625e17628.
[47] M.B. Harris, H. Ju, V.J. Venema, H. Liang, R. Zou, B.J. Michell, Z.P. Chen,
B.E. Kemp, R.C. Venema, Reciprocal phosphorylation and regulation of
endothelial nitric-oxide synthase in response to bradykinin stimulation.
J. Biol. Chem. 276 (2001) 16587e16591.
[48] R. Kou, D. Greif, T. Michel, Dephosphorylation of endothelial nitric-oxide
synthase by vascular endothelial growth factor. Implications for the vascular
responses to cyclosporin A. J. Biol. Chem. 277 (2002) 29669e29673.
[49] D. Fulton, J.E. Church, L. Ruan, C. Li, S.G. Sood, B.E. Kemp, et al., Src kinase
activates endothelial nitric oxide synthase by phosphorylating Tyr 83. J. Biol.
Chem. 43 (2005) 35943e35952.
[50] D. Fulton, L. Ruan, S.G. Sood, C. Li, Q. Zhang, R.C. Venema, Agonist-stimulated
endothelial nitric oxide synthase activation and vascular relaxation: role of
eNOS phosphorylation at Tyr83. Circ. Res. 102 (2008) 497e504.
[51] M. Okuda, M. Takahashi, J. Suero, C.E. Murry, O. Traub, H. Kawakatsu, B.C. Berk,
Shear stress stimulation of p130(cas) tyrosine phosphorylation requires
calcium-dependent c-Src activation. J. Biol. Chem. 274 (1999) 26803e26809.
[52] L.K. Tai, M. Okuda, J. Abe, C. Yan, B.C. Berk, Fluid shear stress activates
proline-rich tyrosine kinase via reactive oxygen species-dependent pathway.
Arterioscler. Thromb. Vasc. Biol. 22 (2002) 1790e1796.
[53] B. Fisslthaler, A.E. Loot, A. Mohamed, R. Busse, I. Fleming, Inhibition of
endothelial nitric oxide synthase activity by proline-rich tyrosine kinase 2 in
response to fluid shear stress and insulin. Circ. Res. 102 (2008) 1520e1528.
[54] D.S. Bredt, S.H. Snyder, Isolation of nitric oxide synthetase, a calmodulin
requiring enzyme. Proc. Natl. Acad. Sci. U.S.A. 87 (1990) 682e685.
[55] U. Forstermann, et al., Calmodulin-dependent endothelium-derived relaxing
factor/nitric oxide synthase activity is present in the particulate and cytosolic
fractions of bovine aortic endothelial cells. Proc. Natl. Acad. Sci. U.S.A. 88
(1991) 1788e1792.
[56] W.N. Durán, A. Seyama, K. Yoshimura, D.R. González, P.I. Jara, X.F. Figueroa,
M.P. Boric, Stimulation of NO production and of eNOS phosphorylation in the
microcirculation in vivo. Microvasc. Res. 60 (2000) 104e111.
[57] Y. Li, J. Zheng, I.M. Bird, R.R. Magness, Mechanisms of shear stress-induced
endothelial nitric-oxide synthase phosphorylation and expression in ovine
fetoplacental artery endothelial cells. Biol. Reprod. 70 (2004) 785e796.
[58] R.S. Scotland, M. Morales-Ruiz, Y. Chen, J. Yu, R.D. Rudic, D. Fulton,
J.P. Gratton, W.C. Sessa, Functional reconstitution of endothelial nitric oxide
synthase reveals the importance of serine 1179 in endothelium-dependent
vasomotion. Circ. Res. 90 (2002) 904e910.
[59] Y. Wang, S. Nagase, A. Koyama, Stimulatory effect of IGF-I and VEGF on eNOS
message, protein expression, eNOS phosphorylation and nitric oxide
production in rat glomeruli, and the involvement of PI3-K signaling pathway.
Nitric Oxide 10 (2004) 25e35.
[60] S.W. Bae, H.S. Kim, Y.N. Cha, Y.S. Park, S.A. Jo, I. Jo, Rapid increase in endo-
thelial nitric oxide production by bradykinin is mediated by protein kinase A
signaling pathway. Biochem. Biophys. Res. Commun. 306 (2003) 981e987.
[61] B.J. Michell, M.B. Harris, Z.P. Chen, H. Ju, et al., Identification of regulatory
sites of phosphorylation of the bovine endothelial nitric-oxide synthase at
serine 617 and serine 635. J. Biol. Chem. 277 (2002) 42344e42351.
[62] H. Chen, M. Montagnani, T. Funahashi, I. Shimomura, M.J. Quon, Adiponectin
stimulates production of nitric oxide in vascular endothelial cells. J. Biol.
Chem. 278 (2003) 45021e45026.
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981194
[63] O. Rocks, A. Peyker, M. Kahms, P.J. Verveer, C. Koerner, M. Lumbierres,
J. Kuhlmann, H. Waldmann, A. Wittinghofer, P.I. Bastiaens, An acylation cycle
regulates localization and activity of palmitoylated Ras isoforms. Science 307
(2005) 1746e1752.
[64] G. Sowa, J. Liu, A. Papapetropoulos, M. Rex-Haffner, T.E. Hughes, W.C. Sessa,
Trafficking of endothelial nitric-oxide synthase in living cells: quantitative
evidence supporting the role of palmitoylation as a kinetic trapping mech-
anism limiting membrane diffusion. J. Biol. Chem. 274 (1999) 22524e22531.
[65] T. Sakoda, K. Hirata, R. Kuroda, N. Miki, M. Suematsu, S. Kawashima,
M. Yokoyama, Myristoylation of endothelial cell nitric oxide synthase is
important for extracellular release of nitric oxide. Mol. Cell Biochem. 152
(1995) 143e148.
[66] L. Busconi, T. Michel, Endothelial nitric oxide synthase. N-terminal myr-
istoylation determines subcellular localization. J. Biol. Chem. 268 (1993)
8410e8413.
[67] I. Liu, T.E. Hughes, W.C. Sessa, The first 35 amino acids and fatty acylation
sites determine the molecular targeting of endothelial nitric oxide synthase
into the Golgi region of cells: a green fluorescent protein study. J. Cell Biol.
137 (1997) 1525e1535.
[68] L.J. Robinson, T. Michel, Mutagenesis of palmitoylation sites in endothelial
nitric oxide synthase identifies a novel motif for dual acylation and subcel-
lular targeting. Proc. Natl. Acad. Sci. U.S.A. 92 (1995) 11776e11780.
[69] L.J. Robinson, L. Busconi, T. Michel, Agonist-modulated palmitoylation of
endothelial nitric oxide synthase. J. Biol. Chem. 270 (1995) 995e998.
[70] R.G. Parton, Caveolae and caveolins. Curr. Opin. Cell Biol. 8 (1996) 542e548.
[71] E.J. Smart, G.A. Graf, M.A. McNiven, W.C. Sessa, J.A. Engelman, P.E. Scherer,
T. Okamoto, M.P. Lisanti, Caveolins, liquid-ordered domains, and signal
transduction. Mol. Cell. Biol. 19 (1999) 7289e7304.
[72] G. Garcia-Cardena, P. Oh, J. Liu, J.E. Schnitzer, W.C. Sessa, Targeting of nitric
oxide synthase to endothelial cell caveolae via palmitoylation: implications
for nitric oxide signaling. Proc. Natl. Acad. Sci. U.S.A. 93 (1996) 6448e6453.
[73] O. Feron, L. Belhassen, L. Kobzik, T.W. Smith, R.A. Kelly, T. Michel, Endothelial
nitric oxide synthase targeting to caveolae: specific interactions with cav-
eolin isoforms in cardiac myocytes and endothelial cells. J. Biol. Chem. 271
(1996) 22810e22814.
[74] G. Garcia-Cardena, P. Martasek, B.S. Masters, P.M. Skidd, J. Couet, S. Li,
M.P. Lisanti, W.C. Sessa, Dissecting the interaction between nitric oxide
synthase (NOS) and caveolin: functional significance of the NOS caveolin
binding domain in vivo. J. Biol. Chem. 272 (1997) 25437e25440.
[75] S. Bulotta, A. Cerullo, R. Barsacchi, C.D. Palma Rotiroti, E. Clementi,
N. Borgese, Endothelial nitric oxide synthase is segregated from caveolin-1
and localizes to the leading edge of migrating cells. Exp. Cell Res. 312 (2006)
877e889.
[76] N.A. Maniatis, V. Brovkovych, S.E. Allen, T.A. John, A.N. Shajahan,
C. Tiruppathi, S.M. Vogel, R.A. Skidgel, A.B. Malik, R.D. Minshall, Novel
mechanism of endothelial nitric oxide synthase activation mediated by
caveolae internalization in endothelial cells. Circ. Res. 99 (2006) 870e877.
[77] H.F. Heijnen, S. Waaijenborg, J.D. Crapo, R.P. Bowler, J.W. Akkerman,
J.W. Slot, Colocalization of eNOS and the catalytic subunit of PKA in endo-
thelial cell junctions: a clue for regulated NO production. J. Histochem.
Cytochem. 52 (2004) 1277e1285.
[78] G. Sowa, M. Pypaert, W.C. Sessa, Distinction between signaling mechanisms
in lipid rafts vs. caveolae. Proc. Natl. Acad. Sci. U.S.A. 98 (2001)
14072e14077.
[79] I. Jiang, D. Cyr, R.W. Babbitt, W.C. Sessa, C. Patterson, Chaperone-dependent
regulation of endothelial nitric-oxide synthase intracellular trafficking by the
cochaperone/ubiquitin ligase CHIP. J. Biol. Chem. 278 (2003) 49332e49341.
[80] P.I. Nedvetsky, W.C. Sessa, H.H. Schmidt, There’s NO binding like NOS
binding:proteineprotein interactions in NO/cGMP signaling. Proc. Natl. Acad.
Sci. U.S.A. 99 (2002) 16510e16512.
[81] U. Zabel, C. Kleinschnitz, P. Oh, P. Nedvetsky, A. Smolenski, H. Müller,
P. Kronich, P. Kugler, U. Walter, J.E. Schnitzer, H.H. Schmidt, Calcium-
dependent membrane association sensitizes soluble guanylyl cyclase to
nitric oxide. Nat. Cell Biol. 4 (2002) 307e311.
[82] W.C. Sessa, G. Garcia-Cardena, J. Liu, A. Keh, J.S. Pollock, J. Bradley, S. Thiru,
I.M. Braverman, K.M. Desai, The golgi association of endothelial nitric oxide
synthase is necessary for the efficient synthesis of nitric oxide. J. Biol. Chem.
270 (1995) 17641e17644.
[83] P.W.Shaul,E.J.Smart,L.J.Robinson,Z.German,I.S.Yuhanna,Y.Ying,R.G.Anderson,
T. Michel, Acylation targets endothelial nitric-oxide synthase to plasmalemmal
caveolae. J. Biol. Chem. 271 (1996) 6518e6522.
[84] I. Hecker, A. Mulsch, E. Bassenge, U. Förstermann, R. Busse, Subcellular
localization and characterization of nitric oxide synthase(s) in endothelial
cells: physiological implications. Biochem. J. 299 (1994) 247e252.
[85] V. Rizzo, D.P. McIntosh, P. Oh, J.E. Schnitzer, In situ flow activates endothelial
nitric oxide synthase in luminal caveolae of endothelium with rapid caveolin
dissociation and calmodulin association. J. Biol. Chem. 273 (1998)
34724e34729.
[86] Y. Feng, V.J. Venema, R.C. Venema, N. Tsai, R.B. Caldwell, VEGF induces
nuclear translocation of Flk-1/KDR, endothelial nitric oxide synthase, and
caveolin-1 in vascular endothelial cells. Biochem. Biophys. Res. Commun.
256 (1999) 192e197.
[87] R.M. Goetz, H.S. Thatte, P. Prabhakar, M.R. Cho, T. Michel, D.E. Golan, Estra-
diol induces the calcium-dependent translocation of endothelial nitric oxide
synthase. Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 2788e2793.
[88] J.B. Michel, O. Feron, D. Sacks, T. Michel, Reciprocal regulation of endothelial
nitric-oxide synthase by Ca2þ
-calmodulin and caveolin. J. Biol. Chem. 272
(1997) 15583e15586.
[89] P. Prabhakar, H.S. Thatte, R.M. Goetz, M.R. Cho, D.E. Golan, T. Michel,
Receptor-regulated translocation of endothelial nitric-oxide synthase. J. Biol.
Chem. 273 (1998) 27383e27388.
[90] V.J. Venema, M.B. Marrero, R.C. Venema, Bradykinin-stimulated protein
tyrosine phosphorylation promotes endothelial nitric oxide synthase trans-
location to the cytoskeleton. Biochem. Biophys. Res. Commun. 226 (1996)
703e710.
[91] Q. Wang, W.F. Patton, H.B. Hechtman, D. Shepro, A novel anti-inflammatory
peptide inhibits endothelial cell cytoskeletal rearrangement, nitric oxide
synthase translocation, and paracellular permeability increases. J. Cell
Physiol. 172 (1997) 171e182.
[92] S. Kai, D. Toomre, Lipid rafts and signal transduction. Nat. Rev. 1 (2000)
31e43.
[93] F.A. Sánchez, N.B. Savalia, R.G. Durán, B.K. Lal, M.P. Boric, W.N. Durán,
Functional significance of differential eNOS translocation. Am. J. Physiol.
Heart Circ. Physiol. 291 (2006) 1058e1064.
[94] B. Yang, T.N. Oo, V. Rizzo, Lipid rafts mediate H2O2 prosurvival effects in
cultured endothelial cells. FASEB J. 20 (2006) 1501e1503.
[95] C. Wann-Hansson, I.R. Hallberg, B. Risberg, A. Lundell, R. Klevsgard, Health-
related quality of life after revascularization for peripheral arterial occlusive
disease: long-term follow-up. J. Adv. Nurs. 51 (2005) 227e235.
[96] I. Jialal, S. Devaraj, S.K. Venugopal, C-reactive protein: risk marker or medi-
ator in atherothrombosis? Hypertension 44 (2004) 6e11.
[97] V. Pasceri, J.T. Willerson, E.T. Yeh, Direct proinflammatory effect of C-reactive
protein on human endothelial cells. Circulation 102 (2000) 2165e2168.
[98] S.K. Venugopal, S. Devaraj, I. Yuhanna, P. Shaul, I. Jialal, Demonstration that
C-reactive protein decreases eNOS expression and bioactivity in human
aortic endothelial cells. Circulation 106 (2002) 1439e1441.
[99] A. Napp, K. Brixius, C. Pott, C. Ziskoven, B. Boelck, U. Mehlhorn, R.H. Schwinger,
W. Bloch, Effects of the beta3-adrenergic agonist BRL 37344 on endothelial
nitric oxide synthase phosphorylation and force of contraction in human failing
myocardium. J. Card. Fail. 15 (2009) 57e67.
[100] S. Tanaka, Y. Toh, R. Mori, K. Komori, K. Okadome, K. Sugimachi, Possible role
of cytomegalovirus in the pathogenesis of inflammatory aortic diseases:
a preliminary report. J. Vasc. Surg. 16 (1992) 274e279.
[101] S. Tanaka, K. Komori, K. Okadome, K. Sugimachi, R. Mori, Detection of active
cytomegalovirus infection in inflammatory aortic aneurysms with RNA
polymerase chain reaction. J. Vasc. Surg. 20 (1994) 235e243.
[102] S. Pampou, S.N. Gnedoy, V.B. Bystrevskaya, V.N. Smirnov, E.I. Chazov,
J.L. Melnick, et al., Cytomegalovirus genome and the immediateeearly
antigen in cells of different layers of human aorta. Virchows Arch. 436 (2000)
539e552.
[103] S.J. Duffy, J.F. Keaney Jr., M. Holbrook, N. Gokce, P.L. Swerdloff, B. Frei,
J.A. Vita, Acute and chronic tea consumption reverses endothelial dysfunc-
tion in patients with coronary artery disease. Circulation 104 (2001)
151e156.
[104] E. Anter, S.R. Thomas, E. Schulz, O.M. Shapira, J.A. Vita, J.F. Keaney Jr., Acti-
vation of endothelial nitric-oxide synthase by the p38 MAPK in response to
black tea polyphenols. J. Biol. Chem. 279 (2004) 46637e46643.
[105] R. Hambrecht, V. Adams, S. Erbs, A. Linke, N. Krankel, Y. Shu, Y. Baither,
S. Gielen, H. Thiele, J.F. Gummert, F.W. Mohr, G. Schuler, Regular physical
activity improves endothelial function in patients with coronary artery
disease by increasing phosphorylation of endothelial nitric oxide synthase.
Circulation 107 (2003) 3152e3158.
[106] B. Ashley, J. Billig, L. Ambrecht, M. Wolfert, Robin Looft-Wilson, Preservation
of EDHF-dependent vasodilation and connexin expression in hyper-
homocysteinemic mice. FASEB J. 21 (2007) 759e763.
[107] X. Jiang, F. Yang, H. Tan, D. Liao, R.M. Bryan Jr., J.K. Randhawa, R.E. Rumbaut,
W. Durante, A.I. Schafer, X. Yang, H. Wang, Hyperhomocystinemia impairs
endothelial function and eNOS activity via PKC activation. Arterioscler.
Thromb. Vasc. Biol. 25 (2005) 2515e2521.
[108] C. Robin, B.S. Looft-Wilson, Ashley, et al., Chronic diet-induced hyper-
homocysteinemia impairs eNOS regulation in mouse mesenteric arteries.
Am. J. Physiol. Regul. Integr. Comp. Physiol. 295 (2008) 59e66.
[109] M.G. Signorello, A. Segantin, M. Passalacqua, G. Leoncini, Homocysteine
decreases platelet NO level via protein kinase C activation. Nitric Oxide 20
(2009) 104e113.
[110] I. Scherrer-Crosbie, R. Ullrich, K.D. Bloch, H. Nakajima, B. Nasseri, H.T. Aretz,
M.L. Lindsey, A.C. Vancon, P.L. Huang, R.T. Lee, W.M. Zapol, M.H. Picard,
Endothelial nitric oxide synthase limits left ventricular remodeling after
myocardial infarction in mice. Circulation 104 (2001) 1286e1291.
[111] C.L. Sartorio, D. Fraccarollo, M. Leutke, J. Bauersachs, The selective aldoste-
rone receptor antagonist eplerenone improves vasomotor dysfunction and
vascular oxidative stress early after myocardial infarction. Circulation 110
(Suppl. II) (2005) 262.
[112] D. Fraccarollo, P. Galuppo, S. Hildemann, M. Christ, G. Ertl, J. Bauersachs,
Additive improvement of left ventricular remodeling and neurohormonal
activation by aldosterone receptor blockade with eplerenone and ACE
inhibition in rats with myocardial infarction. J. Am. Coll. Cardiol. 42 (2003)
1666e1673.
[113] J.W. Calvert, S. Gundewar, S. Jha, J.J.M. Greer, W.H. Bestermann, R. Tian, David
J. Lefer, Acute metformin therapy confers cardioprotection against
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1195
myocardial infarction via AMPK-eNOS-mediated signaling. Diabetes 57
(2008) 696e705.
[114] C. Iadecola, Bright and dark sides of nitric oxide in ischaemic brain injury.
Trends Neurosci. 20 (1997) 132e139.
[115] A.F. Samdani, et al., Nitric oxide synthase in models of focal ischemia. Stroke
28 (1997) 1283e1288.
[116] Z. Huang, et al., Enlarged infarcts in endothelial nitric oxide synthase
knockout mice are attenuated by nitro-L-arginine. J. Cereb. Blood Flow
Metab. 16 (1996) 981e987.
[117] T. Dalkara, et al., Blood flow-dependent functional recovery in a rat model of
focal cerebral ischemia. Am. J. Physiol. 267 (1994) 678e683.
[118] K. Osuka, Y. Watanabe, N. Usuda, A. Nakazawa, M. Tokuda, M.J. Yoshida,
Modification of endothelial no synthase through protein phosphorylation
after forebrain cerebral ischemia/reperfusion. Stroke(35) (2004) 2582.
[119] A. Hashiguchi, S. Yano, M. Morioka, J. Hamada, M. Kochi, K. Fukunaga,
Dephosphorylation of eNOS on Thr495 after transient forebrain ischemia in
gerbil hippocampus. Brain Res. Mol. Brain Res. 133 (2005) 317e319.
[120] D.N. Atochin, A. Wang, V.W. Liu, J.D. Critchlow, A.P. Dantas, R. Looft-Wilson,
T. Murata, S. Salomone, H.K. Shin, C. Ayata, M.A. Moskowitz, T. Michel,
W.C. Sessa, P.L. Huang, The phosphorylation state of eNOS modulates
vascular reactivity and outcome of cerebral ischemia in vivo. J. Clin. Invest.
117 (2007) 1961e1967.
[121] K. Izuishi, A. Tsung, M.A. Hossain, M. Fujiwara, H. Wakabayashi, T. Masaki,
Ischemic preconditioning of the murine liver protects through the Akt kinase
pathway. Hepatology 44 (2006) 573e580.
[122] Y. Wang, Y. Vodovotz, P.K. Kim, R. Zamora, T.R. Billiar, Mechanisms of
hepatoprotection by nitric oxide. Ann. N.Y. Acad. Sci. 962 (2002) 415e422.
[123] H. Taniai, I.N. Hines, S. Bharwani, R.E. Maloney, Y. Nimura, B. Gao, S.C. Flores,
J.M. McCord, M.B. Grisham, T.Y. Aw, Susceptibility of murine periportal
hepatocytes to hypoxia-reoxygenation: role for NO and Kupffer cell-derived
oxidants. Hepatology 39 (2004) 1544e1552.
[124] F. Roviezzo, S. Cuzzocrea, A. Di Lorenzo, V. Brancaleone, E. Mazzon, R. Di
Paola, M. Bucci, G. Cirino, Protective role of PI3-kinase-Akt-eNOS signalling
pathway in intestinal injury associated with splanchnic artery occlusion
shock. Br. J. Pharmacol. 151 (2007) 377e383.
[125] Y. Shi, J.E. Baker, C. Zhang, J.S. Tweddell, J. Su, A. Kirkwood, A. Pritchard Jr.,
Chronic hypoxia increases endothelial nitric oxide synthase generation of
nitric oxide by increasing heat shock protein 90 association and serine
phosphorylation. Circ. Res. 91 (2002) 300e306.
[126] R.F. Furchgott, J.V. Zawadzki, The obligatory role of endothelial cells in the
relaxation of arterial smooth muscle by acetylcholine. Nature 288 (1980)
373e376.
[127] T. Murata, K. Sato, M. Hori, H. Ozaki, H. Karaki, Decreased endothelial nitric-
oxide synthase (eNOS) activity resulting from abnormal interaction between
eNOS and its regulatory proteins in hypoxia-induced pulmonary hyperten-
sion. J. Biol. Chem. 277 (2002) 44085e44092.
[128] F. Coulet, S. Nadaud, M. Agrapart, F. Soubrier, Identification of hypoxia-
response element in the human endothelial nitric oxide synthase gene
promoter. J. Biol. Chem. 278 (2003) 46230e46240.
[129] J. Liu, Y. Gao, S. Negash, L.D. Longo, J.U. Raj, Long-term effects of prenatal
hypoxia on endothelium-dependent relaxation responses in pulmonary
arteries of adult sheep. Am. J. Physiol. Lung Cell. Mol. Physiol. 296 (2009)
547e554.
[130] K.I. Rother, Diabetes treatmentdbridging the divide. N. Engl. J. Med. 356
(2007) 1499e1501.
[131] L.M. Tierney, S.J. McPhee, M.A. Papadakis, Current Medical Diagnosis &
Treatment, International Edition. Lange Medical Books, McGraw-Hill, New
York, 2002, 1203e1215.
[132] D.W. Laight, M.J. Carrier, E.E. Anggard, Antioxidants, diabetes and endothelial
dysfunction. Cardiovasc. Res. 47 (2000) 457e464.
[133] B. Tesfamariam, M.L. Brown, D. Deykin, et al., Elevated glucose promotes
generation of endothelium-derived vasoconstrictor prostanoids in rabbit
aorta. J. Clin. Invest. 85 (1990) 929e932.
[134] H.G. Bohlen, J.M. Lash, Topical hyperglycemia rapidly suppresses EDRF-
mediated vasodilation of normal rat arterioles. Am. J. Physiol. 265 (1993)
219e225.
[135] S. Meraji, L. Jayakody, M.P. Senaratne, et al., Endothelium-dependent relax-
ation in aorta of BB rat. Diabetes 36 (1987) 978e981.
[136] G.M. Pieper, D.A. Meier, S.R. Hager, Endothelial dysfunction in a model of
hyperglycemia and hyperinsulinemia. Am. J. Physiol. 269 (1995) 845e850.
[137] M.T. Johnstone, S.J. Creager, K.M. Scales, et al., Impaired endothelium-
dependent vasodilation in patients with insulin-dependent diabetes melli-
tus. Circulation 88 (1993) 2510e2516.
[138] S.B. Williams, J.A. Cusco, M.A. Roddy, et al., Impaired nitric oxide-mediated
vasodilation in patients with non-insulin-dependent diabetes mellitus. J. Am.
Coll. Cardiol. 27 (1996) 567e574.
[139] P. Clarkson, D.S. Celermajer, A.E. Donald, et al., Impaired vascular reactivity in
insulin-dependent diabetes mellitus is related to disease duration and low
density lipoprotein cholesterol levels. J. Am. Coll. Cardiol. 28 (1996) 573e579.
[140] G.E. McVeigh, G.M. Brennan, G.D. Johnston, et al., Impaired endothelium-
dependent and independent vasodilation in patients with type 2 (non-
insulin-dependent) diabetes mellitus. Diabetologia 35 (1992) 771e776.
[141] X.L. Du, D. Edelstein, S. Dimmeler, Q. Ju, C. Sui, M. Brownlee, Hyperglycemia
inhibits endothelial nitric oxide synthase activity by posttranslational
modification at the Akt site. J. Clin. Invest. 108 (2001) 1341e1348.
[142] E.B. Okon, A.W. Chung, P. Rauniyar, E. Padilla, T. Tejerina, B.M. McManus,
H. Luo, C. van Breemen, Compromised arterial function in human type 2
diabetic patients. Diabetes 54 (2005) 2415e2423.
[143] H. Okumura, N. Nagaya, T. Itoh, I. Okano, J. Hino, K. Mori, Y. Tsukamoto,
H. Ishibashi-Ueda, S. Miwa, K. Tambara, S. Toyokuni, C. Yutani, K. Kangawa,
Adrenomedullin infusion attenuates myocardial ischemia/reperfusion injury
through the phosphatidylinositol 3-kinase/Akt-dependent pathway. Circu-
lation 109 (2004) 242e248.
[144] S. Yang, L. Lin, J.X. Chen, C.R. Lee, J.M. Seubert, Y. Wang, H. Wang, Z.R. Chao,
D.D. Tao, J.P. Gong, Z.Y. Lu, D.W. Wang, D.C. Zeldin, Cytochrome P-450
epoxygenases protect endothelial cells from apoptosis induced by tumor
necrosis factor-alpha via MAPK and PI3K/Akt signaling pathways. Am. J.
Physiol. Heart Circ. Physiol. 293 (2007) 142e151.
[145] K. Kawasaki, R.S. Smith Jr., C.M. Hsieh, J. Sun, J. Chao, J.K. Liao, Activation of
the phosphatidylinositol 3-kinase/protein kinase Akt pathway mediates
nitric oxide-induced endothelial cell migration and angiogenesis. Mol. Cell
Biol. 23 (2003) 5726e5737.
[146] Y. Wang, X. Wei, X. Xiao, R. Hui, J.W. Card, M.A. Carey, D.W. Wang, D.C. Zeldin,
Arachidonic acid epoxygenase metabolites stimulate endothelial cell growth
and angiogenesis via mitogen-activated protein kinase and phosphatidyli-
nositol 3-kinase/Akt signaling pathways. J. Pharmacol. Exp. Ther. 314 (2005)
522e532.
[147] R.R. Shankar, Y. Wu, H.Q. Shen, J.S. Zhu, A.D. Baron, Mice with gene
disruption of both endothelial and neuronal nitric oxide synthase exhibit
insulin resistance. Diabetes 49 (2000) 684e687.
[148] H. Duplain, R. Burcelin, C. Sartori, S. Cook, M. Egli, M. Lepori, P. Vollenweider,
T. Pedrazzini, P. Nicod, B. Thorens, U. Scherrer, Insulin resistance, hyperlip-
idemia, and hypertension in mice lacking endothelial nitric oxide synthase.
Circulation 104 (2001) 342e345.
[149] S. Cook, O. Hugli, M. Egli, B. Menard, S. Thalmann, C. Sartori, C. Perrin,
P. Nicod, B. Thorens, P. Vollenweider, U. Scherrer, R. Burcelin, Partial gene
deletion of endothelial nitric oxide synthase predisposes to exaggerated
high-fat diet-induced insulin resistance and arterial hypertension. Diabetes
53 (2004) 2067e2072.
[150] G. Zeng, F.H. Nystrom, L.V. Ravichandran, L.N. Cong, M. Kirby, H. Mostowski,
M.J. Quon, Roles for insulin receptor, PI3-kinase, and Akt in insulin-signaling
pathways related to production of nitric oxide in human vascular endothelial
cells. Circulation 101 (2000) 1539e1545.
[151] M. Montagnani, H. Chen, V.A. Barr, M.J. Quon, Insulin-stimulated activation
of eNOS is independent of Ca2þ
but requires phosphorylation by Akt at Ser
(1179). J. Biol. Chem. 276 (2001) 30392e30398.
[152] B. Musicki, M.F. Kramer, R.E. Becker, L. Arthur Burnett, Inactivation of
phosphorylated endothelial nitric oxide synthase (Ser-1177) by O-GlcNAc in
diabetes-associated erectile dysfunction. Proc. Natl. Acad. Sci. U.S.A 102
(2005) 11870e11875.
[153] J.W. Elrod, M.R. Duranski, W. Langston, J.J. Greer, L. Tao, T.R. Dugas, C.G. Kevil,
H.C. Champion, D.J. Lefer, eNOS gene therapy exacerbates hepatic ischemia-
reperfusion injury in diabetes: a role for eNOS uncoupling. Circ. Res. 99
(2006) 78e85.
[154] D.S. Marks, J.A. Vita, J.D. Folts, J.F.J. Keaney, G.N. Welch, J. Loscalzo, Inhibition
of neointimal proliferation in rabbits after vascular injury by a single treat-
ment with a protein adduct of nitric oxide. J. Clin. Invest. 96 (1995)
2630e2638.
[155] K.S. Masters, E.A. Lipke, E.E. Rice, M.S. Liel, H.A. Myler, C. Zygourakis,
D.A. Tulis, J.L. West, Nitric oxide-generating hydrogels inhibit neointima
formation. J. Biomater. Sci. Polym. Ed. 16 (2005) 659e672.
[156] M.R. Kapadia, L.W. Chow, N.D. Tsihlis, S.S. Ahanchi, J.W. Eng, J. Murar,
J. Martinez, D.A. Popowich, Q. Jiang, J.A. Hrabie, J.E. Saavedra, L.K. Keefer,
J.F. Hulvat, S.I. Stupp, M.R. Kibbe, Nitric oxide and nanotechnology: a novel
approach to inhibit neointimal hyperplasia. J. Vasc. Surg. 47 (2008)
173e182.
[157] Y. Li, P.I. Lee, Controlled nitric oxide delivery platform based on s-nitro-
sothiol conjugated interpolymer complexes for diabetic wound healing. Mol.
Pharm. 7 (2010) 254e266.
[158] S.S. Ahanchi, V.N. Varu, N.D. Tsihlis, J. Martinez, C.G. Pearce, M.R. Kapadia,
Q. Jiang, J.E. Saavedra, L.K. Keefer, J.A. Hrabie, M.R. Kibbe, Heightened efficacy
of nitric oxide-based therapies in type II diabetes mellitus and metabolic
syndrome. Am. J. Physiol. Heart Circ. Physiol. 295 (2008) 2388e2398.
[159] C.G. Pearce, S.F. Najjar, M.R. Kapadia, J. Murar, J. Eng, B. Lyle, O.O. Aalami,
Q. Jiang, J.A. Hrabie, J.E. Saavedra, L.K. Keefer, M.R. Kibbe, Beneficial effect of
a short-acting NO donor for the prevention of neointimal hyperplasia. Free
Radic. Biol. Med. 44 (2008) 73e81.
[160] T. Nakagawa, W. Sato, O. Glushakova, M. Heinig, T. Clarke, M. Campbell-
Thompson, Y. Yuzawa, M.A. Atkinson, R.J. Johnson, B. Croker, Diabetic
endothelial nitric oxide synthase knockout mice develop advanced diabetic
nephropathy. J. Am. Soc. Nephrol 18 (2007) 539e550.
[161] J. Molnar, S. Yu, N. Mzhavia, C. Pau, I. Chereshnev, H.M. Dansky, Diabetes
induces endothelial dysfunction but does not increase neointimal formation
in high-fat diet fed C57BL/6J mice. Circ. Res. 96 (2005) 1178e1184.
[162] S. Pennathur, J.W. Heinecke, Oxidative stress and endothelial dysfunction in
vascular disease. Curr. Diab. Rep. 7 (2007) 257e264.
[163] U. Hink, H. Li, H. Mollnau, M. Oelze, E. Matheis, M. Hartmann, M. Skatchkov,
F. Thaiss, R.A. Stahl, A. Warnholtz, T. Meinertz, K. Griendling, D.G. Harrison,
U. Forstermann, T. Munzel, Mechanisms underlying endothelial dysfunction
in diabetes mellitus. Circ. Res. 88 (2001) 14e22.
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981196
[164] K. Chalupsky, H. Cai, Endothelial dihydrofolate reductase: critical for nitric
oxide bioavailability and role in angiotensin II uncoupling of endothelial
nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 102 (2005) 9056e9061.
[165] N.J. Alp, S. Mussa, J. Khoo, S. Cai, T. Guzik, A. Jefferson, N. Goh, K.A. Rockett,
K.M. Channon, Tetrahydrobiopterin-dependent preservation of nitric
oxidemediated endothelial function in diabetes by targeted transgenic
GTPcyclohydrolase I overexpression. J. Clin. Invest. 112 (2003) 725e735.
[166] J.B. Laursen, M. Somers, S. Kurz, L. McCann, A. Warnholtz, B.A. Freeman,
M. Tarpey, T. Fukai, D.G. Harrison, Endothelial regulation of vasomotion in
apoE-deficient mice: implications for interactions between peroxynitrite and
tetrahydrobiopterin. Circulation 103 (2001) 1282e1288.
[167] U. Landmesser, S. Dikalov, S.R. Price, L. McCann, T. Fukai, S.M. Holland,
W.E. Mitch, D.G. Harrison, Oxidation of tetrahydrobiopterin leads to
uncoupling of endothelial cell nitric oxide synthase in hypertension. J. Clin.
Invest. 111 (2003) 1201e1209.
[168] T. Munzel, A. Daiber, V. Ullrich, A. Mulsch, Vascular consequences of endo-
thelial nitric oxide synthase uncoupling for the activity and expression of the
soluble guanylyl cyclase and the cGMP-dependent protein kinase. Arte-
rioscler. Thromb. Vasc. Biol. 25 (2005) 1551e1557.
[169] M.B. Harris, H. Ju, V.J. Venema, H. Liang, R. Zou, B.J. Michell, et al., Reciprocal
phosphorylation and regulation of endothelial nitric-oxide synthase in
response to bradykinin stimulation. J. Biol. Chem. 276 (2001) 16587e16591.
[170] C.A. Chen, L.J. Druhan, S. Varadharaj, Y.R. Chen, J.L. Zweier, Phosphorylation
of endothelial nitric-oxide synthase regulates superoxide generation from
the enzyme. J. Biol. Chem. 283 (2008) 27038e27047.
[171] Y. O’Hara, T.E. Peterson, D.G. Harrison, Hypercholesterolemia increases
endothelial superoxide anion production. J. Clin. Invest. 91 (1993)
2546e2551.
[172] T.E. Peterson, V. Poppa, H. Ueba, A. Wu, C. Yan, B.C. Berk, Opposing effects of
reactive oxygen species and cholesterol on endothelial nitric oxide synthase
and endothelial cell caveolae. Circ. Res. 85 (1999) 29e37.
[173] U. Landmesser, H. Cai, S. Dikalov, L. McCann, J. Hwang, et al., Role of p47
(phox) in vascular oxidative stress and hypertension caused by angiotensin
II. Hypertension 40 (2002) 511e515.
[174] S.R. Thomas, K. Chen, J.F. Keaney Jr., Hydrogen peroxide activates en-
dothelial nitric-oxide synthase through coordinated phosphorylation and
dephosphorylation via a phosphoinositide 3- kinase-dependent signaling
pathway. J. Biol. Chem. 277 (2002) 6017e6024.
[175] H. Zhang, J. Zhang, Z. Ungvari, C. Zhang, Resveratrol improves endothelial
function: role of TNF {alpha} and vascular oxidative stress. Arterioscler.
Thromb. Vasc. Biol. 29 (2009) 1164e1171.
[176] W. Insull, The pathology of atherosclerosis: plaque development and plaque
responses to medical treatment. Am. J. Med. 122 (2009) 3e14.
[177] C.D. Liapis, E.D. Avgerinos, N.P. Kadoglou, J.D. Kakisis, What a vascular
surgeon should know and do about atherosclerotic risk factors. J. Vasc. Surg.
49 (2009) 1348e1354.
[178] M.E. Safar, P. Jankowski, Central blood pressure and hypertension: role in
cardiovascular risk assessment. Clin. Sci. 116 (2009) 273e282.
[179] C.M. Werner, M. Böhm, The therapeutic role of RAS blockade in chronic heart
failure. Ther. Adv. Cardiovasc. Dis. 2 (2008) 167e177.
[180] K.K. Gaddam, A. Verma, M. Thompson, R. Amin, H. Ventura, Hypertension
and cardiac failure in its various forms. Med. Clin. North Am. 93 (2009)
665e680.
[181] E. Agabiti-Rosei, From macro- to microcirculation: benefits in hypertension
and diabetes. J. Hypertens. 26 (2008) 15e21.
[182] B.P. Murphy, T. Stanton, F.G. Dunn, Hypertension and myocardial ischemia.
Med. Clin. North Am. 93 (2009) 681e695.
[183] L. Tylicki, B. Rutkowski, Hypertensive nephropathy: pathogenesis, diagnosis
and treatment. Pol. Merkuriusz. Lek. 14 (2003) 168e173.
[184] L.D. Truong, S.S. Shen, M.H. Park, B. Krishnan, Diagnosing nonneoplastic
lesions in nephrectomy specimens. Archiv. Pathol. Lab Med. 133 (2009)
189e200.
[185] R.E. Tracy, S. White, A method for quantifying adrenocortical nodular
hyperplasia at autopsy: some use of the method in illuminating hyperten-
sion and atherosclerosis. Ann. Diagn. Pathol. 6 (2002) 20e29.
[186] D.R. Singer, A. Kite, Management of hypertension in peripheral arterial
disease: does the choice of drugs matter? Eur. J. Vasc. Endovasc. Surg. 35
(2008) 701e708.
[187] W.S. Aronow, Hypertension and the older diabetic. Clin. Geriatr. Med. 24
(2008) 489e501.
[188] A.W. Gardner, A. Afaq, Management of lower extremity peripheral arterial
disease. J. Cardiopulm. Rehab. Prev. 28 (6) (2008) 349e357.
[189] P.A. Cahill, C. Foster, E.M. Redmond, C. Gingalewski, Y. Wu, J.V. Sitzmann,
Enhanced nitric oxide synthase activity in portal hypertensive rabbits.
Hepatology 22 (1995) 598e606.
[190] P.Y. Martin, D.L. Xu, M. Niederberger, A. Weigert, P. Tsai, J. St John, P. Gines,
R.W. Schrier, Upregulation of endothelial constitutive NOS: a major role in
the increased NO production in cirrhotic rats. Am. J. Physiol. Renal Fluid
Electrolyte Physiol. 270 (1996) 494e499.
[191] M. Niederberger, P. Gines, P.Y. Martin, P. Tsai, K. Morris, I. McMurtry, R.W. Schrier,
Comparison of vascular nitric oxide production and systemic hemodynamics in
cirrhosis versus prehepatic portal hypertension in rats. Hepatology 24 (1996)
947e951.
[192] M. Niederberger, P.Y. Martin, P. Gines, K. Morris, P. Tsai, D.L. Xu, I. McMurtry,
R.W. Schrier, Normalization of nitric oxide production corrects arterial
vasodilation and hyperdynamic circulation in cirrhotic rats. Gastroenter-
ology 109 (1995) 1624e1630.
[193] Y. Iwakiri, M. Tsai, J. Timothy, T.J. McCabe, J.P. Gratton, D. Fulton, J.R. Groszmann,
W.C. Sessa, Phosphorylation of eNOS initiates excessive NO production in early
phases of portal hypertension. Am. J. Physiol. Heart Circ. Physiol. 282 (2002)
2084e2090.
[194] V. Shah, Molecular mechanisms in the pathogenesis of cirrhotic portal
hypertension: focus on nitric oxide. J. Gastroenterol. Hepatol. 19 (2004)
145e149.
[195] T.J. Bivalacqua, F.Mustafa Usta, C.H. Champion, J.P. Kadowitz, J.Wayne
G. Hellstrom, Endothelial dysfunction in erectile dysfunction: role of the
endothelium in erectile physiology and disease. J. Androl. 24 (2003).
[196] B. Musicki, A.l. Burnett, eNOS function and dysfunction in the penis. Exp. Biol.
Med. 231 (2006) 154e165.
[197] J. Hidetaka, E. Masato, M.R. Kano, S. Ogawa, K. Iijima, M. Akishita, Y. Ouchi,
Cilostazol inhibits oxidative stress-induced premature senescence via
upregulation of sirt1 in human endothelial cells. Arterioscler. Thromb. Vasc.
Biol. 28 (2008) 1634.
[198] T. Minamino, H. Miyauci, T. Yoshida, Y. Ishida, H. Yoshida, I. Komuro,
Endothelial cell senescence in human atherosclerosis: role of telomere in
endothelial dysfunction. Circulation 105 (2002) 1541e1544.
[199] K.F. Burrig, The endothelium of advanced arteriosclerotic plaques in humans.
Arterioscler. Thromb. 11 (1991) 1678e1689.
[200] T. Hayashi, H. Matsui-Hirai, A. Miyazaki-Akita, A. Fukatsu, J. Funami, Q.F. Ding,
S. Kamalanathan, Y. Hattori, L.J. Ignarro, A. Iguchi, Endothelial cellular
senescence is inhibited by nitric oxide: implications in atherosclerosis asso-
ciated with menopause and diabetes. Proc. Natl. Acad. Sci. U.S.A. 103 (2006)
17018e17023.
[201] M. Vasa, K. Breitschopf, A.M. Zeiher, S. Dimmeler, Nitric oxide activates
telomerase and delays endothelial cell senescence. Circ. Res. 87 (2000)
540e542.
[202] J. Kambayashi, Y. Liu, B. Sun, Y. Shakur, M. Yoshitake, F. Czerwiec,
Cilostazol as a unique antithrombotic agent. Curr. Pharm. Des. 9 (2003)
2289e2302.
[203] T. Hayashi, K. Yano, H. Matsui-Hirai, H. Yokoo, Y. Hattori, A. Iguchi, Nitric
oxide and endothelial cellular senescence. Pharmacol. Ther. 120 (2008)
333e339.
[204] A. Hashimoto, G. Miyakoda, Y. Hirose, T. Mori, Activation of endothelial nitric
oxide synthase by cilostazol via a cAMP/protein kinase A- and phosphati-
dylinositol 3-kinase/Akt-dependent mechanism. Atherosclerosis 189 (2006)
350e357.
[205] M. Braunstein, A.B. Rose, S.G. Holmes, C.D. Allis, J.R. Broach, Transcriptional
silencing in yeast is associated with reduced nucleosome acetylation. Genes
Dev. 7 (1993) 592e604.
[206] H. Vaziri, S.K. Dessain, N.E. Eaton, S.I. Imai, R.A. Frye, T.K. Pandita, L. Guarente,
R.A. Weinberg, hSIR2 (SIRT1) functions as an NAD-dependent p53 deacety-
lase. Cell 107 (2001) 149e159.
[207] E. Langley, M. Pearson, M. Faretta, U.M. Bauer, R.A. Frye, S. Minucci, P.G. Pelicci,
T. Kouzarides, Human SIR2 deacetylates p53 and antagonizes PML/p53-
induced cellular senescence. EMBO J. 21 (2002) 2383e2396.
[208] F. Picard, M. Kurtev, N. Chung, A. Topark-Ngarm, T. Senawong, R. Machado
De Oliveira, M. Leid, M.W. McBurney, L. Guarente, Sirt1 promotes fat
mobilization in white adipocytes by repressing PPAR-gamma. Nature 429
(2004) 771e776.
[209] E. Nisoli, C. Tonello, A. Cardile, V. Cozzi, R. Bracale, L. Tedesco, S. Falcone,
A. Valerio, O. Cantoni, E. Clementi, S. Moncada, M.O. Carruba, Calorie
restriction promotes mitochondrial biogenesis by inducing the expression of
eNOS. Science 310 (2005) 314e317.
[210] M. Potente, S. Dimmeler, NO targets SIRT1: a novel signaling network in
endothelial senescence. Arterioscler. Thromb. Vasc. Biol. 28 (2008)
1577e1579.
[211] Y. Korkmaz, The basal phosphorylation sites of endothelial nitric oxide
synthase at serine (Ser)1177, Ser116, and threonine (Thr)495 in rat molar
epithelial rests of malassez. J. Periodontol. 76 (2005) 1513e1519.
[212] M.T. Gentile, C. Vecchione, A. Maffei, A. Aretini, G. Marino, R. Poulet,
L. Capobianco, G. Selvetella, G. Lembo, Mechanisms of soluble b-amyloid
impairment of endothelial function. J. Biol. Chem. 279 (2004)
48135e48142.
[213] J.S. Stamler, L. Jia, J.P. Eu, T.J. McMahon, I.T. Demchenko, J. Bonaventura,
K. Gernert, C.A. Piantadosi, Blood flow regulation by S-nitro-
sohemoglobin in the physiological oxygen gradient. Science 276 (1997)
2034e2037.
[214] J. Kleinbongard, R. Schulz, T. Rassaf, et al., Red blood cells express a func-
tional endothelial nitric oxide synthase. Blood 107 (2006) 2943e2951.
[215] T. Anuk, E.B. Assayag, R. Rotstein, R. Fusman, D. Zeltser, S. Berliner,
D. Avitzour, I. Shapira, N. Arber, N.M. Bornstein, Prognostic implications of
admission inflammatory profile in acute ischemic neurological events. Acta
Neurol. Scand. 106 (2002) 196e199.
[216] P. Muda, P. Kampus, M. Zilmer, K. Zilmer, C. Kairane, T. Ristimäe, K. Fischer,
R. Teesalu, Homocysteine and red blood cell glutathione as indices for
middle-aged untreated essential hypertension patients. J. Hypertens. 21
(2003) 2329e2333.
[217] D. Mihov, J. Vogel, M. Gassmann, A. Bogdanova, Erythropoietin activates
nitric oxide synthase in murine erythrocytes. Am. J. Physiol. Cell. Physiol. 297
(2009) 378e388.
G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1197
Acknowledgment in Research article at AUKBC

More Related Content

What's hot

Test article 2
Test article 2Test article 2
Test article 2
Greg
 
Postranslational Modification
Postranslational ModificationPostranslational Modification
Postranslational Modification
Bahaddin Ahmad
 
Key pointsmedicalsciencesataglance
Key pointsmedicalsciencesataglanceKey pointsmedicalsciencesataglance
Key pointsmedicalsciencesataglance
Elsa von Licy
 

What's hot (20)

pharmaceutics
pharmaceuticspharmaceutics
pharmaceutics
 
Nitric oxide and its role in reproduction
Nitric oxide and its role in reproductionNitric oxide and its role in reproduction
Nitric oxide and its role in reproduction
 
Wp seidman 10
Wp seidman 10Wp seidman 10
Wp seidman 10
 
Wp seidman 05
Wp seidman 05Wp seidman 05
Wp seidman 05
 
Post translational modification
Post translational modificationPost translational modification
Post translational modification
 
Histone modifications.ppt
Histone modifications.pptHistone modifications.ppt
Histone modifications.ppt
 
Test article 2
Test article 2Test article 2
Test article 2
 
Post translational modification of protein
Post translational modification of proteinPost translational modification of protein
Post translational modification of protein
 
Role of nitric oxide in cell signaling
Role of nitric oxide in cell signalingRole of nitric oxide in cell signaling
Role of nitric oxide in cell signaling
 
Post tranlational modification
Post tranlational modificationPost tranlational modification
Post tranlational modification
 
Postranslational Modification
Postranslational ModificationPostranslational Modification
Postranslational Modification
 
Key pointsmedicalsciencesataglance
Key pointsmedicalsciencesataglanceKey pointsmedicalsciencesataglance
Key pointsmedicalsciencesataglance
 
Post translational modification in plants.
Post translational modification in plants.Post translational modification in plants.
Post translational modification in plants.
 
Oxidative stress and Liver fibrosis
Oxidative stress and Liver fibrosisOxidative stress and Liver fibrosis
Oxidative stress and Liver fibrosis
 
Post Translational Modification
Post Translational ModificationPost Translational Modification
Post Translational Modification
 
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
Flavocytochrome p450 bm3 mutant a264 e undergoes substrate dependent formatio...
 
Biotech 2012 spring_8_post-trans
Biotech 2012 spring_8_post-transBiotech 2012 spring_8_post-trans
Biotech 2012 spring_8_post-trans
 
Post translational-modification-creative-peptides
Post translational-modification-creative-peptidesPost translational-modification-creative-peptides
Post translational-modification-creative-peptides
 
co and post translation modification
co and post translation modificationco and post translation modification
co and post translation modification
 
Ptms sonika chouhan
Ptms   sonika chouhanPtms   sonika chouhan
Ptms sonika chouhan
 

Similar to Acknowledgment in Research article at AUKBC

Dissecting NOS electronics_Biochem J 2013
Dissecting NOS electronics_Biochem J 2013Dissecting NOS electronics_Biochem J 2013
Dissecting NOS electronics_Biochem J 2013
Luciana Hannibal, Ph.D.
 
Synthesis of proteins modified
Synthesis of proteins   modifiedSynthesis of proteins   modified
Synthesis of proteins modified
Bruno Mmassy
 

Similar to Acknowledgment in Research article at AUKBC (20)

NO homeostasis_Hannibal_CAR 2016 (1)
NO homeostasis_Hannibal_CAR 2016 (1)NO homeostasis_Hannibal_CAR 2016 (1)
NO homeostasis_Hannibal_CAR 2016 (1)
 
Biochemistry of nitric oxide
Biochemistry of nitric oxideBiochemistry of nitric oxide
Biochemistry of nitric oxide
 
Dissecting NOS electronics_Biochem J 2013
Dissecting NOS electronics_Biochem J 2013Dissecting NOS electronics_Biochem J 2013
Dissecting NOS electronics_Biochem J 2013
 
Single Nucleotide Polymorphism - The new generation therapy
Single Nucleotide Polymorphism - The new generation therapySingle Nucleotide Polymorphism - The new generation therapy
Single Nucleotide Polymorphism - The new generation therapy
 
2046-1682-6-1
2046-1682-6-12046-1682-6-1
2046-1682-6-1
 
Ed endotlemium
Ed endotlemiumEd endotlemium
Ed endotlemium
 
JBC2
JBC2JBC2
JBC2
 
Diabetic nephropathy
Diabetic nephropathyDiabetic nephropathy
Diabetic nephropathy
 
Endoplasmic Reticulum: Structure, Roles, and Importance in Cellular Biology |...
Endoplasmic Reticulum: Structure, Roles, and Importance in Cellular Biology |...Endoplasmic Reticulum: Structure, Roles, and Importance in Cellular Biology |...
Endoplasmic Reticulum: Structure, Roles, and Importance in Cellular Biology |...
 
Nitric oxide .pdfckjgcgchfcfhxcfhcfhchfch
Nitric oxide .pdfckjgcgchfcfhxcfhcfhchfchNitric oxide .pdfckjgcgchfcfhxcfhcfhchfch
Nitric oxide .pdfckjgcgchfcfhxcfhcfhchfch
 
CcP_IJQC_2001
CcP_IJQC_2001CcP_IJQC_2001
CcP_IJQC_2001
 
Papadie 2
Papadie 2Papadie 2
Papadie 2
 
BBL2
BBL2BBL2
BBL2
 
Synthesis of proteins modified
Synthesis of proteins   modifiedSynthesis of proteins   modified
Synthesis of proteins modified
 
CELLULAR REPROGRAMMING: Current Technology, Perspectives and Generation of iP...
CELLULAR REPROGRAMMING: Current Technology, Perspectives and Generation of iP...CELLULAR REPROGRAMMING: Current Technology, Perspectives and Generation of iP...
CELLULAR REPROGRAMMING: Current Technology, Perspectives and Generation of iP...
 
endoplasmic reticulum
endoplasmic reticulum endoplasmic reticulum
endoplasmic reticulum
 
Nrf2 Transcription Factor- Nuclear Factor- Erythroid 2 related factor)
Nrf2 Transcription Factor- Nuclear Factor- Erythroid 2 related factor)Nrf2 Transcription Factor- Nuclear Factor- Erythroid 2 related factor)
Nrf2 Transcription Factor- Nuclear Factor- Erythroid 2 related factor)
 
Epigenetics
EpigeneticsEpigenetics
Epigenetics
 
Nitric oxide
Nitric oxideNitric oxide
Nitric oxide
 
Soal dan Pembahasan Farmakologi Molekular - PPAR dan Reseptor Estrogen
Soal dan Pembahasan Farmakologi Molekular - PPAR dan Reseptor EstrogenSoal dan Pembahasan Farmakologi Molekular - PPAR dan Reseptor Estrogen
Soal dan Pembahasan Farmakologi Molekular - PPAR dan Reseptor Estrogen
 

Acknowledgment in Research article at AUKBC

  • 1. Review eNOS phosphorylation in health and disease Gopi Krishna Kolluru, Jamila H. Siamwala, Suvro Chatterjee* Vascular Biology Lab, AU-KBC Research Centre, MIT Campus, Anna University, Chennai 600 044, TN, India a r t i c l e i n f o Article history: Received 13 February 2010 Accepted 29 March 2010 Available online 2 April 2010 Keywords: eNOS Serine phosphorylation Nitric oxide a b s t r a c t Endothelium plays a fundamental role in maintaining the vascular tone by releasing various biochemical factors that modulate the contractile and relaxatory behavior of the underlying vascular smooth muscle, regulation of inflammation, immunomodulation, platelet aggregation, and thrombosis. Endothelium regulates these cellular processes by activating endothelial nitric oxide synthase (eNOS) responsible for nitric oxide (NO) production. eNOS is constitutively expressed in ECs in response to humoral, mechanical or pharmacological stimulus. eNOS activity is regulated mainly by protein-protein interactions and multisite phosphorylations. The phosphorylation state of specific serine, threonine and tyrosine residues of the enzyme plays a pivotal role in regulation of eNOS activity. Perturbations of eNOS phosphorylation have been reported in a number of diseases thereby emphasizing the importance of regulation of eNOS activity. This review summarizes the mechanism of eNOS regulation through multi-site phosphorylation in different pathologies. Attempts have been made to highlight phosphorylation of eNOS at various residues, regulation of the enzyme activity via posttranslational modifications and its implications on health and disease. Ó 2010 Elsevier Masson SAS. All rights reserved. 1. Introduction There has been a drastic transformation from times past when eNOS regulation was thought to be a very simple process. Many studies and research groups working on eNOS regulation have come a long way in discovering the increasing number of inter- acting protein kinases and the role of eNOS phosphorylation on specific serine (Ser) and threonine (Thr) residues. The plethora of studies on eNOS phosphorylation and control of NO synthesis has furthered the understanding of pathophysiological conditions of endothelial dysfunctions and rectification. Inspite of several studies on eNOS phosphorylation over a decade, the complexity of eNOS regulation still exists and poses to be a topic of interest for researchers. The changes in eNOS activity in relation to eNOS phosphorylation pattern will provide information regarding attenuated NO availability and the pathogenic mechanisms involved in multiple disease conditions and their potential treat- ment conditions related to NO deficiency. This review attempts to summarize the present knowledge of the multi-site phosphoryla- tion of eNOS under different physiological and pathological conditions. 1.1. NO in physiology and pathology Nitric oxide (NO) is one of the simplest gaseous free radicals and mediates a significant and diverse number of signaling functions in nearly every organ system in the body. NO produced in the endo- thelium via the enzyme eNOS is an important vasoactive compound [1,2]. NO is responsible to regulate a diverse range of physiological and cellular processes including endothelial cell migration, prolif- eration, extracellular matrix degradation, and angiogenesis [3]. NO is a key component in endothelium-dependent regulation of vascular tone, platelet function, angiogenesis and mitogenesis that are crucial in cardiovascular physiology [4]. A potent anti- inflammatory agent, NO inhibits leukocyte interactions with the vessel wall, thereby reducing pathological inflammation and thrombosis [5]. eNOS knockout mice studies showed that leukocyte adherence to the vessel wall is elevated 10-fold [6]. Also NO inhibits inflammation in various vascular disease models like myocardial infarction, glomerulonephritis, lung injury, and stroke [7,8]. Thereby NO controls inflammation and thrombosis in part by regulating vesicle trafficking. NO decreases granule trafficking from the Golgi to the plasma membrane by targeting a key component of exocytic machinery [9]. Loss of endothelial function due to phos- phorylation and sub-cellular localization of eNOS [11,12], has been implicated in a number of cardiovascular diseases, making regula- tion of eNOS through posttranslational modifications, a promising therapeutic target [2,10]. * Corresponding author. Tel.: þ91 44 2223 4885/2711x48; fax: þ91 44 2223 1034. E-mail address: soovro@yahoo.ca (S. Chatterjee). Contents lists available at ScienceDirect Biochimie journal homepage: www.elsevier.com/locate/biochi 0300-9084/$ e see front matter Ó 2010 Elsevier Masson SAS. All rights reserved. doi:10.1016/j.biochi.2010.03.020 Biochimie 92 (2010) 1186e1198
  • 2. 1.2. Expression and regulation of eNOS NO is an important signaling molecule but is also highly reactive and highly diffusible. It is therefore important that there is strict control and regulation of NO production. NO is produced by a group of enzymes called neuronal NOS (nNOS), inducible NOS (iNOS) and endothelial NOS (eNOS) [13,14]. These enzymes convert arginine into citrulline, producing NO in the process. The activity of NOS enzymes is subject to a discreet and multiple interconnected mechanisms of regulation. There are many ways by which regulation of NOS occurs, such as the intracellular distribution, gene expression, protein binding, enzymatic activa- tion by phosphorylation and cellular inhibitors of NOS activity. iNOS expression is controlled by inflammatory mediators and cytokines and produces large, unregulated quantities of NO, whereas nNOS and eNOS produce low amounts of NO. nNOS and especially eNOS activity are induced by physiological and patho- physiological stimuli [15e17]. Various chemical stimuli like VEGF, estrogen receptor modulator and sphingosine-1-phosphate or mechanical factors like shear stress activate eNOS [18e20]. eNOS is regulated by multiple interdependent control mechanisms and signaling pathways which can be calcium-dependent and/or -independent [13,21e24]. 1.3. Mechansim of eNOS activation It is well-established that eNOS is a calmodulin dependent enzyme [6,25]. eNOS is activated by elevation of intracellular calcium and subsequent activation of CaM-binding domain by calmodulin [26,27]. eNOS comprises of a C-terminal reductase domain, which transfers the electron flux via NADPH to the N-terminal oxygenase domain, which contains a heme, and binding sites for arginine, tetrahydrobiopterin and CaM. Further the CaM- binding domain facilitates electron flux from the reductase domain to the oxygenase domain [28]. The phosphorylation of eNOS allows active flux of electrons from the reductase domain to NO generation in the oxygenase domain by guarding the steric hindrance due to some non-catalytic inserts [29]. eNOS activity can be induced by several agonists like acetylcholine, bradykinin B2 receptor, thrombin and ATP, through intracellular elevation of calcium levels [30]. Studies [1,31] have shown that the contribution of calcium and substrate/co-factor have been crucial to the activity of eNOS in intracellular organelles like Golgi, mitochondria, nuclei and plasma membrane (Figs. 1 and 2). 1.4. Calcium independent eNOS activation Apart from canonical calmodulin/Ca-dependent eNOS activa- tion, there is a Ca-independent regulation of eNOS. Substantial increase in NO release induced by hemodynamic shear stress appears to be largely independent of intracellular calcium [26]. Shear stress induces eNOS mRNA and protein levels consistent with eNOS-derived NO, which is important for flow-dependent vasodi- lation and remodeling of vessels [32,33]. Increase in blood flow through exercise upregulates eNOS, which is part of the cardio- and athero-protective phenomenon [34]. It was also reported that ceramide promotes eNOS activation and enzyme translocation in endothelial cells [24]. Various studies have reinforced that substantial amount of NO is produced in unstimulated endothelial cells by a Ca-independent mechanism [35e37]. Mechanisms like phosphorylation and asso- ciated regulatory proteins like Hsp90; gp60, Caveolin-1, a major caveolae-localized protein through caveolae internalization, play a crucial role in Ca2þ -independent activation of eNOS and conse- quent NO production (discussed below). 1.5. eNOS phosphorylation Protein phosphorylation, a posttranslational modification and a key regulator of eNOS activity [6,38e41] is modulated by kinases, phosphatases [42e45] and proteineprotein interactions [21]. The requisite substrates and cofactors for eNOS function are L-arginine, tetrahydrobiopterin (BH4), iron (Fe), FMN, FAD and NADPH [45]. Though eNOS activity is coupled to changes in Ca2þ levels, it is not the sole factor required for the regulation of the enzyme activity. The binding of calmodulin (CaM) and the flow of electrons from the reductase to the oxygenase domain of the enzyme is also depen- dent on protein phosphorylation and dephosphorylation [1]. The primary sites where eNOS gets phosphorylated are serine residues and, to a lesser extent, on tyrosine (Tyr) and threonine residues. While phosphorylation of Ser617, 635, and 1179 results in the activation of eNOS, the phosphorylation of Ser116 and Thr497 reduces the eNOS function [44,46e48]. Recently few reports have shown that tyrosine residues present within oxygenase domain of eNOS regulate the production of NO, as the modulation of eNOS activity by tyrosine phosphorylation is supposedly prominent in primary endothelial cells [49,50]. Tyr81 and Tyr657 are the residues, which gets phosphorylated due to oxidative-stress and overexpression of v-Src leading to the regulation of eNOS activity [46]. It appears that Tyr81 mediates phosphorylation of eNOS through Src mediated mechanism under different agonists [47]. While the role played by the other Tyr residue Tyr657 is on the contrary. This particular tyrosine residue, which is prominent during shear stress works through PYK2/c-Src in phosphorylating eNOS [48,49,51e53]. It seems to provide a negative feed back to eNOS phosphorylation for NO production. However the proper mechanistic role through which the two Tyro- sine residues target the phosphorylation is an open area to address. 1.6. Kinases involved in eNOS phosphorylation Calmodulin is one of the very first protein involved in the regulation of eNOS [54,55]. Increase in intracellular calcium levels activates CaM that in turn activates CaM kinase II, which phos- phorylates Ser1179 on eNOS. Ser1177 (human eNOS) and Ser1179 (bovine eNOS) are phosphorylated by the serine/threonine kinase Akt (protein kinase B) [6,22,23]. Under physiological conditions, esatcudeResanegyxO Calmodulin COOH P T495 / 497 S116 S1177/1179 P PS615/ 617 S633/ 635 P P HDL Shear stress Shear stress, VEGF, IGF-1 bradykinin, insulin, estrogen sphingosine 1-phosphate, adiponectin, leptin, statins 8-Br-cAMP Shear stress, VEGF, bradykinin, statins 8-Br-cAMP PMA NH2 Fig. 1. Regulators of eNOS acting on various phosphorylation sites. The eNOS phos- phorylation sites are numbered according to human and bovine eNOS sequence. Abbreviations: VEGF, vascular endothelial cell growth factor; 8-Br-cAMP, 8-bromoa- denosine-30,50-cyclic monophosphate; S-1-P, sphingosine 1-phosphate. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1187
  • 3. shear stress, hormones and autacoids activate eNOS by phosphor- ylation/dephosphorylation mechanisms [22,39,40,46,56e59]. Humoral factors like Bradykinin stimulate NO synthesis by increasing phosphorylation of multi-sites at eNOS-Ser1177 [45,47,60], eNOS-Ser633 and eNOS-Ser615 [61], and stimulating dephosphorylation of eNOS-Thr495 [45,47] as well. The phos- phorylation of Ser1179 is common for multiple diverse signaling systems, such as insulin and adipokines [56,61,62]. Apart from these, kinases like Akt [47], and PKA, phosphatases like PP1 and calcineurin (PP2B) [45] have been implicated in the regulation of phosphorylation. 2. eNOS e posttranslational modifications 2.1. Myristoylation and palmitoylation eNOS comprises two types of fatty acid modifications: irre- versible acylation i.e., myristoylation and reversible palmitoylation. Myristoylation provides general membrane association, while palmitoylation directs proteins specifically to the plasma membrane (PM). The presence of eNOS at the PM may bring eNOS to interact with the factors that are required for its proper function, such as arginine, calcium, and cofactor BH4. Proteins during acyl- ation get depalmitoylated at the PM and are redistributed to other intracellular membranes [40]. Repalmitoylation occurs specifically at the Golgi, which further directs the proteins to the PM [63,64]. Studies have demonstrated that the myristoyl moiety is an absolute requirement for the membrane localization and activity of eNOS [65,66]. Without this modification, eNOS is almost completely cytosolic and lacks palmitoyl moieties [67,68]. While myr- istoylation is an absolute necessity for eNOS membrane activity, palmitoylation determines the sub-cellular localization of the eNOS. Palmitoylation is in turn regulated by eNOS activators like bradykinin [69]. The cycle of depalmitoylation and repalmitoyla- tion, results in a constant shuttling of the enzyme between Golgi and PM, redistributing eNOS to distinct membrane microdomains. 2.2. Other interacting proteins of eNOS eNOS interacts with the proteins like caveolae coat protein caveolin-1 (Cav-1) and heat shock protein 90 (Hsp90) during its active and inactive states which regulates its activity. Caveolae are specialized invaginations of the PM composed of cholesterol, glycosphingolipids, and some structural proteins, such as caveolin [70,71]. eNOS is bound to caveolin-1 in caveolae facilitated by myristoylation and palmitoylation, due to which the enzyme activity of eNOS is basally repressed [40,72e75]. Caveolin-1 medi- ated regulation of eNOS through caveolae-mediated endocytosis gets initiated once the glycoprotein gp60 is activated in endothelial cells [76]. Maniatis et al., had shown that NO production in pulmonary endothelial cells is significantly mediated by caveolae internalization and is independent of the increase of intracellular Ca2þ . Compartmentalization of eNOS in caveolae is necessary for its interaction with regulatory proteins, and calcium- and phosphor- ylation-dependent signal transduction events that modify the response of the enzyme to extracellular stimuli. Stimuli, such as shear stress, induce calcium increase thereby displacing eNOS from caveolin-1 and activating eNOS by redistributing it from plasma membrane caveolae [77,78]. The chaperone protein Hsp90 can interact with eNOS and positively influence its function. The C-terminal Hsp70-interacting protein (CHIP) interacts with both Hsp70 and Hsp90, and nega- tively regulates eNOS trafficking into the Golgi complex. By contrast, the NOS interacting protein (NOSIP) and the NOS traffic inducer (NOSTRIN) can negatively regulate eNOS localization in the plasma membrane [28,79e81]. 2.3. eNOS localization eNOS, the key endothelial isoform has a unique sub-cellular localization pattern. In addition to phosphorylation, the location of eNOS is important for its activation [74,82,83]. Functional eNOS was found to be primarily located in the plasma membrane regions of endothelial cells [84]. The eNOS expression and the site of its synthesis as well, have a major influence on the biological activity of the molecule. Complex mechanisms that include the cell specificity of eNOS expression, and co- and posttranslational processing lead to trafficking of the enzyme to plasma membrane caveolae [21]. In the plasma membrane, eNOS is mainly targeted to the cav- eolae [72,83], where it is inhibited by binding to caveolin-1 (cav-1) through a consensus site [11,74]. Calcium-calmodulin and intra- cellular calcium can dissociate eNOS from cav-1, allowing activation of the enzyme [74,85]. On external stimulus, eNOS shuttles between caveolae and sub-cellular compartments such as cytosol, Golgi, and/or perinuclear structures [86,87e91]. This sub-cellular T495S116 S1177 Atherosclerosis Coronary Artery Disease Hyperhomocysteinaemia Myocardial infarction Cerebral Ischaemia Severe subarachnoid haemorrhage Liver Ischemia Hypoxia Diabetes Portal hypertension Erectile dysfunction Hertwig's epithelial root sheath Alzheimers Disease Cancer esatcudeResanegyxO Calmodulin Hyperhomocysteinaemia Cerebral Ischaemia Hypoxia Diabetes Erectile dysfunction Alzheimers Disease P PP Atherosclerosis Coronary Artery Disease Erectile dysfunction Hertwig's epithelial root sheath Alzheimers Disease Fig. 2. Effect of various pathophysiological and pathological conditions on the phosphorylation sites of eNOS. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981188
  • 4. localization of eNOS plays an important role in the regulation of its activity possibly along with phosphorylation [86e91]. Lipid rafts, which are rich cholesterol and sphingolipid regions found in surface invaginations of cell, polymerize the caveolins to form caveoli, which are implicated in endocytosis [92]. Many studies implicate a coordinated role played by lipid rafts and caveolins in eNOS sub-cellular signaling [71,78,93]. Studies provide evidence that lipid rafts are implicated in oxidative-stress signaling on endothelial cells [93,94]. This information hints a possible regulatory mechanism to control selective movement of eNOS signaling by modifying the signaling activity of lipid raft or caveolae, to counteract and protect the cellular sensitivity to redox stress. 2.4. eNOS phosphorylation in human pathology 2.4.1. Artherosclerosis e endothelial dysfunction Atherosclerosis is the major cause of chronic vascular diseases such as coronary artery disease, cerebrovascular disease and peripheral arterial occlusive disease [95]. Coronary artery disease is characterized by accumulation of atheromatous plagues in lumen of arteries which supply nutrients, oxygen to the myocardium. Causes of the disease are obesity, diabetes, hypertension and smoking. Upon progress of the disease, the lumen becomes obstructed completely, leading to myocardial infarction, chronic coronary ischemia, angina and flash edema. Coronary artery disease starts with the formation of atherosclerotic lesions. It is characterized by decreased eNOS activity and NO bioavailability and increased expression of cell adhesion molecules such as VCAM-1 and ICAM-1 [96e98]. eNOS inhibition has also been shown to accelerate atherosclerosis suggesting that NO may inhibit several key steps in the atherosclerotic process. Thus, eNOS could be a candidate gene, which is implicated in atherosclerosis. Modulation of eNOS activity by dynamic changes in phosphorylation of eNOS has been of considerable interest because of its pathophysiological role. In failing cardiomyocytes Napp et al. showed beta (3) adrenergic stimulation seems to deactivate rather than activate eNOS. During stimulation with beta (3) adrenergic agonist BRL improvises a further dephosphorylation of eNOS (Ser1177) and Akt, while the treatment increases the Ser114 phosphorylation of eNOS. In atherosclerosis caused by cytomegalovirus infection Ser1177 phosphorylation was shown to be decreased dramatically in infected cells [99e102]. Black tea was found to increase NO bioavailability in patients with artherosclerosis [103]. ERa (estrogen receptor alpha) plays a key role in mediating the activation of eNOS in response to black tea polyphenols. In cultured cells, the black tea polyphenolic frac- tion promotes both eNOS catalytic activity and NO bioactivity. This effect is because of activation of phosphoinositol 3-kinase (PI3-K) and Akt via a p38 MAPK-dependent mechanism [104]. This was supported by Anter et al. 2004 who showed that black tea poly- phenols induced time-dependent phosphorylation of ERa on Ser118, which was inhibited by ER antagonist ICI 182,780. Ser118 phosphorylation was observed in response to black tea polyphenols and mutation of this residue abrogated polyphenol-induced eNOS activation in COS cells. In addition to S118 phosphorylation, Ser1177 was also found to be phosphorylated by 300% in CAD after 4 weeks of daily aerobic exercise training compared with sedentary controls [105]. Therefore two key residues S118 and Ser1177 phosphoryla- tion were required for regulation of NO production in patients with coronary artery disease (CAD). 2.4.2. Hyperhomocysteinaemia Homocysteine produces endothelial injury and stimulates platelet aggregation. Hyperhomocysteinaemia has been associated with increased risk of thrombosis and atherosclerosis. Hyperhomocysteinemia (HHcy) impairs endothelium-dependent vasodilation by increasing reactive oxygen species, thereby reducing NO bioavailability [106]. HHcy impairs endothelial func- tion and eNOS activity, primarily through PKC activation [107]. Robin et al. showed significantly less basal eNOS and phospho- Ser1179-eNOS/eNOS in mesenteric arteries from HHcy mice but no difference in phospho-thr495-eNOS/eNOS [108]. Signerello et al. has shown that PKC stimulates the eNOS phosphorylation of the negative regulatory residue thr495 and the dephosphorylation of the positive regulatory site Ser1177 [109]. 2.4.3. Myocardial infarction Myocardial infarction (MI) or acute myocardial infarction (AMI) occurs due to blockage in coronary artery following rupture of a vulnerable atherosclerotic plague. In chronic myocardial infarc- tion infarct size in eNOSÀ/À mice was unchanged but evident remodeling with less capillary density and hypertrophy accompa- nied with subsequent systolic and diastolic dysfunction and increased mortality at 28 days was observed [110], emphasizing the beneficial effects of eNOS-derived NO on ventricular remodeling after myocardial infarction. Recent data showed that blocking mineralocorticoid receptor improved endothelial dysfunction and oxidative stress by increasing reduced eNOS phosphorylation at Ser1177 thereby making NO available in experimental myocardial infarction [111]. Furthermore, in rats with chronic heart failure after large myocardial infarction [112] eplerenone and metformin [113] in combination with an angiotensin-converting enzyme inhibitor increased myocardial eNOS phosphorylation at Ser1177. 2.4.4. Ischemia Ischemia occurs due to shortage of oxygen, glucose and other blood born nutrients and is caused due to constriction or blockage of blood vessels Myocardial infarction, stroke, organ trans- plantation, and cardiopulmonary bypass causes reperfusion injury which is leading cause of tissue damage leading to ischemia. In all these conditions, the initial trigger of the damage is the transient disruption of the normal blood supply to target organs followed by reperfusion. From a clinical viewpoint, no therapy is currently available to limit reperfusion injury, which emphasizes the importance of a better understanding of its underlying pathological mechanisms, to devise potential future therapeutic strategies. 2.4.5. Cerebral ischemia Prominent roles of eNOS and vascular NO in maintaining cere- bral blood flow and prevention of neuronal injury have been shown in by pharmacological and genetical approaches in animal models of cerebral ischemia [114e117]. Vascular NO production protects against stroke regulates and cerebrovascular perfusion by increasing collateral flow to the ischemic area. eNOS knockout mice show decreased blood flow in the ischemic border zone and develop larger cerebral infarctions [116]. The vascular endothelial growth factor (VEGF) contributes to activation of eNOS by Ca2þ /calmodulin and also stimulates the protein kinase Akt, which directly phosphorylate eNOS on Ser1177 and increases enzyme activity. Increase in Ser1177 phospho-eNOS occurs in endothelial cells of microvessels after ischemic episodes with temporal expression of VEGF [118]. Osuka et al. (2004) showed 0.5e2 h transient increase in phospho-Akt at Ser 437 after reper- fusion, whereas after 6 h there was an elevation of phospho-eNOS at Ser1177. Hashiguchi et al. (2005) showed in the gerbil hippo- campal microvasculature transient ischemia model that Ser1177 phosphorylation was unchanged by 24 h after reperfusion, despite post-ischemic up-regulation of eNOS protein [119]. However, Thr495 phosphorylation significantly and persistently decreased by 48 h [119]. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1189
  • 5. Stroke is loss of brain functions due to ischemia caused by thrombosis or embolism or haemorrhage. Stroke can lead to permanent neurological damage, complications and death. Risk factors include advance age, hypertension, transient ischemic attack, diabetes, high-cholesterol, cigarette smoking and atrial fibrillation. Stroke can be ischemic stroke or haemorrhagic stroke. Dmitriy et al. determined the effects of modulation of the Ser1179 phosphorylation site on vascular reactivity, cerebral blood flow, and outcome in a middle cerebral artery (MCA) occlusion model of stroke [120]. Their results indicate that modulation of the Ser1179 phosphorylation site affects endothelium-dependent vasodilation and cerebral blood flow and that these effects determine outcome of stroke in vivo. 2.4.6. Liver ischemia Ischemic hepatitis is caused by diffusion ischemia while hepatic infarction results from hepatic artery disorders. Ischemic hepatitis is characterized by diffuse liver damage due to an inadequate blood or O2 supply. Causes are most often systemic e impaired hepatic perfusion (due to heart failure or acute hypotension), Hypoxemia (due to respiratory failure or carbon monoxide toxicity) or increased metabolic demand (due to sepsis). Recent findings demonstrate that the acute activation of the serine-threonine kinase Akt is cardioprotective and PI3K/Akt pathway activation has protective effects on hepatic I/R injury [121]. Activation of PI3K leads to phosphorylation of membrane phosphatidylinositol 3,4 biphosphate, which recruits Akt to the cell membrane leading to phosphorylation and activation of Akt. Activated Akt in turn promotes eNOS phosphorylation. In the liver PI3K-Akt-eNOS thus appears to play a central role in pro- tecting against ischemiaereperfusion (IeR) injury [122,123]. Roviezzo et al. showed activation of the Akt pathway in ischemic regions of reperfused ileum through an increased S1179 phos- phorylation in reperfused intestinal tissue coupled to Akt activation [124]. 2.4.7. Severe subarachnoid haemorrhage Severe subarachnoid haemorrhage (SAH) is a form of stroke that induces dysfunction of endothelial nitirc oxide synthase (eNOS), resulting in severe vasospasm. Clinically, however, some portions of cerebral arteries may show only mild vasospasm. Although severe vasospastic arteries after SAH have been intensively studied, activity of eNOS with the mild form of the disease has received less attention. Osuka et al. 2004 showed that SAH induces a temporary activation of AMPK alpha, which phosphorylates eNOS at Ser1177 in endothelial cells of mild vasospastic basilar arteries. This signal transduction might play an important role in controlling cerebral blood flow after SAH [118]. 2.4.8. Hypoxia Hypoxia is a pathological condition in which body as a whole or a region of the body is deprived of adequate oxygen supply. Chronic hypoxia increases endothelial nitric oxide synthase (eNOS) production of nitric oxide (NO) and cardioprotection in neonatal rabbit hearts [125]. The effect of hypoxia on eNOS expression remains controversial. The presence of inflammatory mediators and cytokines under hypoxic conditions further decreases expres- sion of eNOS and production of NO [126]. Ser1177 phosphorylation under hypoxia was observed by a number of workers. Shi et al. 2002 showed that association of Hsp90 with eNOS is important for increasing NO production by Ser1177 phosphorylation and limiting eNOS-dependent superoxide anion generation. Hsp90 is associated with eNOS and the extent to which the enzyme is activated is based on phosphorylation of eNOS at Ser1177 [126]. Additionally, hypoxic conditions attenuate Akt-mediated phos- phorylation at Ser1177, alter calcium metabolism, and alter the balance of eNOS proteineprotein interactions with caveolin and calmodulin [127]. Coulet et al. 2003 showed that human eNOS is a hypoxia-inducible gene, whose transcription is stimulated through HIF-2 interaction with two contiguous sites located at À5375 to À5366 of the human eNOS promoter [128]. Liu J et al. 2009 have shown that phosphorylation of Ser1177 in eNOS decreased, whereas phosphorylation of Thr495 increased, in the prenatally hypoxic pulmonary arteries [129]. These data demon- strate that prenatal hypoxia results in persistent abnormalities in endothelium-dependent relaxation responses of pulmonary arteries in adult sheep due to decreased eNOS activity resulting from altered posttranslational regulation. 2.4.9. Diabetes In diabetes, the body either fails to properly respond to its own insulin, does not make enough insulin, or both. This causes glucose to accumulate in the blood, often leading to various complications [130,131]. Endothelial dysfunction is a critical initial factor 2 in the development of diabetic vascular disease [132]. Endothelial dysfunction, as represented by impaired endothelium functions, and NO-mediated relaxation, occurs in cellular and experimental models of diabetes [133e136]. Similarly, many, but not all, clinical studies have found that endothelium-dependent vasodilation is abnormal in patients with type 1 or type 2 diabetes [137e140]. Thus, decreased levels of NO in diabetes may underlie its athero- genic predisposition. The defect of Akt/eNOS signaling may play a primary role in endothelial dysfunction in type 2 diabetes mellitus. In aorta from diabetic animals, as well as in type 2 diabetic patients, Akt/eNOS phosphorylation has been shown to be decreased [141,142], at least partly via hyperglycemia-induced GlcNAc modification of the enzyme, which may explain the development and progression of diabetes-associated atherosclerosis [141]. Insulin also mediates its effects through binding to insulin receptors and triggering down- stream signaling pathways, of which the most important is the phosphatidylinositol 3-kinase (PI3K)-Akt/protein kinase B (Akt/ PKB) pathway. This pathway is involved in a variety of insulin responses including transport of glucose through cell membranes, myocardial survival, and anti-apoptotic effects in endothelial cells [143,144]. As such, PI3K-Akt pathway activation by eNOS-derived NO may result in improved endothelial function and rescue of impaired myocardial cells [145,146]. Recent studies on eNOS gene disruption studies in mice revealed that, this leads to insulin resistance, resulting in hypertension and hyperlipidemia [147e149]. Further biochemical studies in insulin-responsive cells have uncovered phosphorylation-dependent signaling role in insulin-stimulated activation of eNOS [150,151]. In diabetic rat model, Biljana Musicki et al. showed that eNOS dysfunction in the penis in diabetes by O-GlcNAc modification affecting phosphorylation at Ser1177 residue, and rendering it incapable of activation by normal fluid shear stress stimuli and VEGF signaling [152]. Such eNOS impairment could contribute to erectile dysfunction and affect long-term penile health in diabetes. In diabetic liver Elrod et al. showed that although there was no difference in total hepatic eNOS protein between non-diabetic and diabetic animals, they discovered decreased phosphorylation of both Thr495 and Ser1177 residues in diabetic mice as compared with non-diabetic control animals [153]. In another study of dia- betic animals [141] it has also been shown that this regulatory site is subject to posttranslational modification, thereby decreasing eNOS activity. NO-based therapies have been proved by numerous investiga- tions in various animal models [154e156]. Evidence also suggest G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981190
  • 6. that NO is involved in the diabetic pathology [157]. In Type II dia- betes mellitus (DM) where neointimal hyperplasia conditions lead to vascular interventions, external delivery of NO proved to recover arterial injury from neointimal hyperplasia [158,159]. Uncoupling of eNOS has been demonstrated in animal models of diabetes [160]. Together, these data indicate that diabetes and insulin resistance are characterized, at least in part, by endothelial dysfunction and potentially by altered eNOS expression and NO production. Insulin treatments have stimulated phosphorylation of eNOS and protein kinase B (Akt) in arteries of diabetic mice [161]. We can interpret from these studies that activation or inhibition of eNOS would rectify diabetes pathology. We presume that phosphoryla- tion of eNOS play key role in rectifying the defects under this diseased condition and further sites involved in eNOS phosphory- lation under these conditions should be identified and extensively studied. 2.4.10. Oxidative stress Oxidative stress, a pathological imbalance due to anomalous reactive oxygen species (ROS) production, is a common feature of many of the pathological conditions like cardiovascular anomalies, hypertension, diabetes, atherosclerosis, cancer, which lead to endothelial dysfunction [162]. ROS production involves inactivation of the signaling molecule nitric oxide (NO), leading to endothelial dysfunction. Uncoupled endothelial nitric oxide synthase (eNOS) could be another important source of ROS in the endothelium. Uncoupling of eNOS is now recognized as important in the patho- physiology of several cardiovascular disorders and a predictor of future adverse vascular events such as hypertension, atheroscle- rosis and diabetes [163,164]. Under deficiency of the NOS cofactor tetrahydrobiopterin (BH4), the enzyme can become uncoupled and generate superoxide instead of NO. Furthermore, BH4 itself is prone to degradation by oxidation, which leads to further amplification of ROS production due to NOS uncoupling [164,165]. (ONOOÀ ) oxidizes BH4 and reduces the availability of BH4 resulting in eNOS uncoupling, thereby increased superoxide and decreased NO production [166,167]. Role of ROS in oxidative stress under pathological conditions lead by decoupling of eNOS and consequent reduction of NO production is very much evident from the literature [165,166]. It has been studied that Thr495 dephosphorylation has been impli- cated in the uncoupling of eNOS resulting to increased superoxide (O2 À ) stress instead of NO, which has resulted in atherosclerosis [168]. But the increased NO production even under Thr495 dephosphorylation under inducers like bradykinin is still a point of study [44,169,170]. It has to be explored further as such why Thr495 dephosphorylation behaving in a different manner under various conditions of pathology and physiology, and is there any feed back mechanism from this event to the NO signaling system? Tyrosine phosphorylation is a novel mechanism of eNOS regu- lation, which can affect the activity of eNOS and also influence the distribution of eNOS in the cellular compartments. Tyrosine may interact with other intercellular regulatory proteins like caveolin-1 that may influence the activity of eNOS [40]. However, studies have to be carried out to elucidate the proper mechanism of tyrosine phosphorylation and its interaction with other proteins in regu- lating eNOS activity and thereby NO production, which may provide a valuable information in therapeutic studies related to vascular anomalies. 2.4.11. Oxidative stress in type II diabetes Superoxide production contributes to atherogenesis as evi- denced by increased superoxide anion content in vessels from animals fed with high-cholesterol diets [171]. Superoxide reduces expression of eNOS protein, as well as decrease the number of caveolae in endothelial cells [172]. The most significant source of superoxide in the vascular wall is NADPH oxidase [173]. The enzyme superoxide dismutase acts to detoxify superoxide and releases hydrogen peroxide (H2O2). H2O2 induces phosphorylation of eNOS-Ser1177 and dephosphorylation of eNOS-Thr495, by tyrosine kinaseedependent PI3K/Akt mechanisms [174]. Oxidative stress plays an important role in type 2 diabetes-related endothelial dysfunction. Zhang H et al. 2009 showed that Resveratrol restored endothelial function in type 2 diabetes by inhibiting TNF-alpha- induced activation of NAD(P)H oxidase and preserving eNOS phosphorylation (Ser1177) [175]. 2.4.12. Hypertension Hypertension is a risk factor for all clinical manifestations of atherosclerosis since it is a risk factor for atherosclerosis itself [176e180]. It is an independent predisposing factor for heart failure [181], coronary artery disease [182e184] stroke, renal disease [185e187], and peripheral arterial disease [188,189]. Cirrhosis is one of the main causes of portal hypertension. Adrenal cortical abnormalities, kidney diseases, neuroendocrin tumors are also known to cause secondary hypertension. Defect in eNOS phos- phorylation has been accounted for endothelial dysfunction further leading to hypertension, hyperlipidemia, and other diseased conditions. Atochin et al. has shown modulation of S1179 phos- phorylation as an approach for treating cardiovascular diseases, particularly influenced by diabetes, obesity, metabolic syndrome, hyperlipidemia, and hypertension [120]. However the effects of phosphorylation sites of eNOS apart from S1179 on vascular path- ogenesis have to be further studied in detail. 2.4.13. Portal hypertension and liver cirrhosis Normally, blood from the intestines and spleen is carried to the liver through the portal vein. But cirrhosis slows the normal flow of blood, which increases the pressure in the portal vein resulting in the condition called portal hypertension. Overproduction of vascular NO plays a central role in both systemic and splanchnic vasodilatation, which is a hallmark of portal hypertension [190e193]. In the chronic model of portal hypertension increased eNOS expression and enzyme activity are well-established events [191e194]. However, the mechanism of the early induction of excessive NO production by eNOS remains to be elucidated. The phosphorylation of eNOS by Akt activates the enzyme and may be the first step in increasing NO production in portal hypertension. Iwakiri et al. showed that the phosphorylation of eNOS at Ser1176 was significantly increased in the PVL group [193]. Furthermore, PVL significantly increased Akt phosphorylation (an active form of Akt). When vessels were treated with wortmannin (10 nM) to block the PI3K/Akt pathway, NO-induced vasodilatation was significantly reduced. 2.4.14. Erectile dysfunction Erectile dysfunction (ED) is defined as consistent inability to obtain or maintain an erection for satisfactory sexual intercourse. ED is predominantly the disease of vascular origin. The incidence of ED dramatically increases in men with diabetes mellitus, hyper- cholesterolemia, and cardiovascular disease. Loss of the functional integrity of the endothelium and subsequent endothelial dysfunc- tion plays an integral role in the occurrence of ED in this cohort of men [195]. Age related ED is associated with eNOS activation through dysregulation of its phosphorylation. PI3-K/Aktedependent eNOS activation has recently been shown in the penis. In rats and mice, both neuro- and agonist-induced penile erection produced rapid increase in Akt and eNOS phosphorylation at Ser1177 in the penis, which remain elevated after the termination of the initial stimulus G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1191
  • 7. [196]. In another study Akt dependent phosphorylation of a posi- tive regulatory site Ser1177 on eNOS was shown to be decreased whereas phosphorylation of a negative regulatory site thr495 on eNOS increased in the aged rat penis. In diabetic rat penis decreased eNOS phosphorylation on Ser1177 caused by O-linked N-acetyl- glucoseamine (O-GlcNAc) modification of eNOS. 2.4.15. eNOS and human aging Hidetaka Ota et al. 2008 showed the phenomenon of human aging is known to be a critical cardiovascular risk factor [197]. Cellular senescence of endothelial cells has been proposed to be involved in endothelial dysfunction and atherogenesis [198]. The lesions of human atherosclerosis have been extensively studied histologically, and these studies have demonstrated that there are vascular cells that exhibit the morphological features of cellular senescence [199]. Although there is substantial evidence demon- strating an aging-associated development of cardiac and vascular dysfunction, the mechanisms responsible for this phenomenon have not yet been clearly established. Aging is associated with erectile dysfunction (ED) attributed to reduced nitric oxide syn- thase (NOS) activity and nitric oxide bioavailability. A lot of work has been carried out on endothelial cellular senescence [200e202]. Endothelial NO can protect against a state of oxidative stress, and activation of eNOS and subsequent production of NO delay endothelial cellular senescence [200,201]. A PDE3 inhibitor, cil- ostazol, used as a vasodilating antiplatelet drug for treating inter- mittent claudication, and in preclinical studies was shown to have a protective effect on endothelial cells by increasing eNOS activity [202]. Cilostazol increases intracellular cAMP content accordingly and activates protein kinase A (PKA) or PI3K/Akt signaling [203]. In yeast, Sir2 (silent information regulator-2) has been identified as an NADþ -dependent histone deacetylase [204]. Mammalian sirtuin 1 (Sirt1), the closest homolog of Sir2, regulates the cell cycle, senes- cence, apoptosis, and metabolism, by interacting with a number of molecules, including p53, PML, and PPAR-&UnknownEntity; [205e207]. A recent study showed that production of NO by caloric restriction increases Sirt1 expression and suggested that eNOS may be involved in regulating the expression of Sirt1 in murine white adipocytes [208]. HUVECs treated with other cAMP-elevating agents and DETA-NO showed a reduction of SA-bgal-positive cells as well [209]. Cilostazol increased phosphorylation of Akt at Ser473 and of endothelial nitric oxide synthase (eNOS) at Ser1177, with a dose-dependent increase in Sirt1 expression. Moreover, the effect of cilostazol on premature senescence was abrogated through inhibition of Sirt1 [210]. 2.4.16. eNOS phosphorylation and Hertwig’s epithelial root sheath Epithelial cell rests of Malassez (ERM) are quiescent epithelial remnants of Hertwig’s epithelial root sheath (HERS) that are involved in the formation of tooth roots. After completion of crown formation, HERS are converted from cervical loop cells, which have the potential to generate enamel for tooth crown formation. Korkmaz 2005 has shown that low concentrations of NO produced by nNOS, and eNOS activate intracellular soluble guanylate cyclase (sGC) to produce intracellular cyclic guanosine 30:50-mono- phosphate (cGMP), which triggers rapid cellular responses such as cell proliferation, cell differentiation, and apoptosis under physio- logical conditions [211]. He later showed the basal production of NO by eNOS in the ERM is modulated by phosphorylation of eNOS at Ser1177 and Ser116 residues, while the basal activity of the eNOS is not influenced by phosphorylation of eNOS at Thr495 residue. 2.4.17. Alzheimers disease Alzheimer’s disease (AD) is associated to a cerebral amyloid angiopathy with dysregulation of cerebral blood flow (CBF). Vascular risk factors (VRFs) are probably determinants of incidence and course of AD. b-amyloid peptides (AbP), the main component of senile plaques typical of AD, circulate in soluble globular form in bloodstream. Interestingly, AbP is able to induce endothelial dysfunction, and this effect may represent the link between vascular and neuronal pathophysiological factors involved in AD. Gentile et al. 2004 showed that AbP enhances eNOS phosphorylation on thr495 and Ser116 and reduces acetylcholine-induced phosphorylation on Ser1177 [212]. Such an effect depends on a PKC signaling pathway, as suggested by its phosphorylation on Ser 660. 2.4.18. RBC-eNOS RBC-NOS-generated NO contributes to the overall NO pool and also play prominent role in regulation of blood flow and platelet Table 1 List of publications showing the effect of different diseases on the phosphorylation status of eNOS. Phosphorylation site Origin Mode of induction/ inhibition References Ser1177; Thr495 Hypoxic pulmonary arteries of sheep Chronic hypoxia Liu J, et al./Am. J. Physiol. Lung Cell. Mol. Physiol. 2009;296(3):L547e54. Ser1177 Diabetic Mice Aortic EC Diabetes Zhang H, et al./Arterioscler. Thromb. Vasc. Biol. 2009;29(8):1164e71. Ser1177 Diabetic animal’s aorta Telmisartan Wenzel P, et al./Free Radic. Biol. Med. 2008;45(5):619e26. Ser1177 Aortas from DOCA rats & unineph- rectomized rats PugNAc Lima VV, et al./Hypertension 2009;53(2):166e74. Ser1177 Ischemic myocardium affected rats WY-14643 Bulhak AA, et al./Am. J. Physiol. Heart Circ. Physiol. 2009; 296(3):H719e27. Ser1177; Thr495 Mouse aorta Tacrolimus Cook LG, et al./Kidney Int. 2009;75(7):719e26. Ser1177; Ser114 eNOS in failng myocardium BRL37344 Napp A, et al./J Card Fail. 2009;15(1):57e67. Ser1177 HAECs TNF-alpha-neutralizing antibody Yuen DY, et.al/Diabetes 2009 May;58(5):1086e95. Ser1177; Thr495 Recombinant BH(4)-free eNOS VEGF/PKC Chen CA, et.al/J. Biol. Chem. 2008 Oct 3;283(40):27038e47. Ser1177; Thr495 Renal vascular -eNOS Radicicol Ramírez V, et.al/Am. J. Physiol. Renal Physiol. 2008;295(4):F1044e51. Ser1177 ECV-D8eNOSGFP and CVEC PAF Sánchez FA, et.al/Proc. Natl. Acad. Sci. U. S. A. 2009 21;106(16):6849e53. Ser1179 Cerebral artery Iscemic stroke Dmitriy et al./Clin. Invest. 2007;117:1961e7. Ser1177 Forebrain Cerebral ischemia Osuka et al./Stroke 2004; (35) 2582 Ser1177 Mice Myocardial infarction NOS3 deficiency Scherrer-Crosbie et al./Circulation 2001;104:1286e1291 Ser1177 Diabetic myocardial tissue Eplerenone and metformin Calvert et al./Diabetes 2008;57:696e705 Ser1177 Human dermal microvascular endothelial cells HIF-1, HIF-2 Coulet et al./J. Biol. Chem. 2003;278: 46230e46240. Ser1177; Ser116; Thr495 Vessels of cerebral amyloid aorta AbP Gentile et al./J. Biol. Chem. 2004; 279: 48135e48142 G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981192
  • 8. aggregation [213,214]. RBCs play a role in the pathogenesis of hypertension and stroke [215,216]. Recently, functionally active endothelial type NO synthase was discovered in mature murine and human red blood cells (RBC-eNOS). Nikolaev et al. found that the treatment of mouse erythrocytes with rHuEpo resulted in a time- and dose-dependent up-regulation of NO production, mediated via activation of the PI3K/Akt pathway and RBC-eNOS phosphorylation at Ser1177 [217]. Red blood cell (RBC)-derived NOS has common but also distinct regulatory mechanisms when compared with eNOS [218], which depends on intracellular Ca2þ level and phosphorylation at Ser1177, regulated by PI3K [215]. Suhr et al. 2009 has investigated the influence of intensive exercise on eNOS content and the phosphorylation states of the eNOS at Ser116, Ser1177, and Thr495 in human erythrocytes [219]. Evidence from the literature suggests that in vivo flow dynamics are more affected by RBC aggregation on to endothelial cells leading to pathological conditions [220]. RBC adhesiveness and aggregation have proved to be useful markers for detecting vascular inflam- mation and atherosclerosis progression in patients with coronary artery disease [221]. As RBC-derived NO significantly contributes to the intravascular NO pool, a lack of NOS activity and eNOS phos- phorylation seems to be likely in these diseases. Rosuvastatin has shown an improved RBC-eNOS activity through phosphorylation of eNOS at Ser1177 residue. This improvement resembled the in vivo situation of RBCs passing through capillaries [222]. RBC-NOS will serve as a key factor for fields such as atherosclerosis, microcircu- latory diseases, RBC aging and storage and adaptation mechanisms for high altitude [218]. 2.4.19. eNOS in cancer eNOS has been detected in tumour cells [223], and catalyzes the synthesis of NO, which can facilitate S-nitrosylation of the thiol group of cysteines (Cys) in proteins [224], such as that of Cys118 of HRas, which enhances the dissociation of guanine nucleotides thereby increasing GTP-bound HRas [225]. Wildtype Ras proteins can be required for activation of the MAPK pathway by oncogenic Ras [226], and membrane-targeting of RasGAP, which inhibits wildtype but not oncogenic Ras, reverts oncogenic Ras trans- formation of NIH3T3 cells [227], suggesting that wildtype Ras proteins may facilitate oncogenic signaling. Kian-Huat Lim et al. has shown that AKT activation of eNOS maintains tumour growth in the absence of oncogenic Ras by activating wildtype Ras through S-nitrosylation of Cys118. Activated KRas and Ser1177 phosphorylated eNOS were also elevated in the tumour specimens compared to matched and unmatched normal tissue controls [228], with the caveat that biopsies also contain stromal tissue that could contribute to detected eNOS phosphorylation. 3. Therapeutic challenges Phosphorylation of eNOS is a key mechanism responsible for eNOS activity and subsequent NO production [22,23,27,39,41]. Anomalies like cardiovascular dysfunctions, erectile dysfunctions, stroke and several vascular abnormalities have been implicated for erratic NO production. In the recent years eNOS phosphorylation has been a potential target as a new therapeutic area of interven- tion in many pathological situations, to develop a way of promoting NO production [120,229]. Dmitriy et al. has shown in vivo that eNOS phosphorylation is an important determinant of vascular function, blood flow, and cerebral ischemia. This work also provides proof of concept for the modulation of Ser1179 phosphorylation as an approach to prevent cardiovascular disease, particularly influ- enced by risk factors of diabetes, metabolic syndrome, hyperlipid- emia, and hypertension. Recent advances have shown that pharmacological agents like Raloxifene, Fasudil and Y27632 improve cardiovascular function [231e234]. Studies by Leung et al. shows that the therapeutic concentrations of raloxifene (1e3 nM) augment endothelial func- tions through up-regulation of eNOS activity by increased eNOS phosphorylation in porcine coronary arteries [230]. Y-27632 and fasudil, are some of the emerging drugs the physiological role of Rho-Kinases (ROCKs), particularly in cardiovascular disease [232e234]. Inhibition of RhoA or ROCKs leads to the rapid activa- tion of PI3K/Akt and phosphorylation of eNOS [235,236], suggest- ing the potential role of ROCKs in regulating eNOS activation in addition to eNOS expression. Fasudil and Y-27632 were shown to be effective Rho-kinase inhibitors for the treatment of a wide range of cardiovascular disease, including cerebral and coronary vaso- spasm, angina, hypertension, pulmonary hypertension, and heart failure. They rapidly increase the endothelial eNOS activity through phosphorylation and exert cardiovascular protection [234,236]. Rho/Rho-kinase pathway plays an important role in various intra- cellular functions that are involved in the pathogenesis of cardio- vascular disease [237]. 4. Future challenges The pattern of eNOS phosphorylation and dephosphorylation later evolved as exceedingly complex. The challenge for future studies will be to examine the relative contribution of each regu- latory site on both the level and the time course of NO production. Hopefully, this may help to design smarter eNOS constructs that, on delivering in cardiovascular tissue, would drive NO release where and when required. RBC-eNOS is another emerging area of research, which is gaining prominence for fields like atheroscle- rosis, vasculocirculatory diseases, RBC aging and storage and high altitude pathophysiology. Novel diagnostic approaches for quanti- fying RBC-eNOS and identification of phosphorylation dysfunction under diseased condition help in the development of new diag- nostic and therapeutic strategies for these diseases. Regulatory mechanisms of various phosphorylation sites of eNOS, are to be further elucidated. Interactive studies of these phosphorylation residues with other regulatory protein of eNOS have to be carried out extensively. This may provide valuable information in thera- peutics of vascular anomalies related to eNOS/NO dysfunctions. Acknowledgement This work was supported by a grant from KBC-RF. The authors also acknowledge Ms. Puja Kumari for compilation of the infor- mation in Table 1. References [1] I. Fleming, R. Busse, Molecular mechanisms involved in the regulation of the endothelial nitric oxide synthase. Am. J. Physiol. Regul. Integr. Comp. Physiol. 284 (2003) 1e12. [2] J.A. Panza, C.E. Garcia, C.M. Kilcoyne, A.A. Quyyumi, R.O. Cannon, Evidence that endothelial dysfunction in patients with hypercholesterolemia is not due to increased extracellular nitric oxide breakdown by superoxide anions. 3rd Circul. 91 (1995) 1732e1738. [3] J.P. Cooke, D.W. Losordo, Nitric oxide and angiogenesis. Circulation 105 (2002) 2133e2135. [4] K.M. Naseem, The role of nitric oxide in cardiovascular diseases. Mol. Aspect. Med. 26 (2005) 33e65. [5] P. Kubes, M. Suzuki, D.N. Granger, Nitric oxide: an endogenous modulator of leukocyte adhesion. Proc. Natl. Acad. Sci. U.S.A. 88 (1991) 4651e4655. [6] D.J. Lefer, S.P. Jones, W.G. Girod, et al., Leukocyte-endothelial cell interactions in nitric oxide synthase-deficient mice. Am. J. Physiol. 276 (1999) 1943e1950. [7] S.P. Jones, W.G. Girod, A.J. Palazzo, D.N. Granger, M.B. Grisham, D. Jourd’Heuil, et al., Myocardial ischemia-reperfusion injury is exacerbated G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1193
  • 9. in absence of endothelial cell nitric oxide synthase. Am. J. Physiol. 276 (1999) 1567e1573. [8] P. Heeringa, H. van Goor, Y. Itoh-Lindstrom, N. Maeda, R.J. Falk, K.J. Assmann, et al., Lack of endothelial nitric oxide synthase aggravates murine accelerated anti-glomerular basement membrane glomerulonephritis. Am. J. Pathol. 156 (2000) 879e888. [9] V. Malhotra, L. Orci, B.S. Glick, M.R. Block, J.E. Rothman, Role of an N-ethyl- maleimide-sensitive transport component in promoting fusion of transport vesicles with cisternae of the Golgi stack. Cell 54 (1988) 221e227. [10] S.C. Tai, G.B. Robb, P.A. Marsden, Arterioscler. Thromb. Vasc. Biol. 24 (2004) 405e412. [11] D. Fulton, J. Fontana, G. Sowa, J.P. Gratton, M. Lin, K.X. Li, B. Michell, B.E. Kemp, D. Rodman, W.C. Sessa, Localization of endothelial nitric-oxide synthase phosphorylated on serine 1179 and nitric oxide in Golgi and plasma membrane defines the existence of two pools of active enzyme. J. Biol. Chem. 277 (2002) 4277e4284. [12] T. Michel, O. Feron, Nitric oxide synthases: which, where, how, and why? J. Clin. Invest. 100 (1997) 2146e2152. [13] W.K. Alderton, C.E. Cooper, R.G. Knowles, Nitric oxide synthases: structure, function and inhibition. Biochem. J. 357 (2001) 593e615. [14] C. Nathan, Q.-W. Xie, Nitric oxide synthase: roles, tolls and controls. Cell 78 (1994) 915e918. [15] A. Papapetropoulos, R.D. Rudic, W.C. Sessa, Molecular control of nitric oxide synthases in the cardiovascular system. Cardiovasc. Res. 43 (1999) 509e520. [16] Y. Wang, P.A. Marsden, Nitric oxide synthases: gene structure and regulation. Adv. Pharmacol. 34 (1995) 71e90. [17] U. Forstermann, J.P. Boissel, H. Kleinert, Expressional control of the “consti- tutive” isoforms of nitric oxide synthase (NOS I and NOS III). FASEB J. 12 (1998) 773e790. [18] C. Nathan, Q.W. Xie, Regulation of biosynthesis of nitric oxide. J. Biol. Chem. 269 (1994) 3725e3728. [19] A. Papapetropoulos, G. García-Cardeña, J.A. Madri, W.C. Sessa, Nitric oxide production contributes to the angiogenic properties of vascular endothelial growth factor in human endothelial cells. J. Clin. Invest. 100 (1997) 3131e3139. [20] T. Simoncini, G. Varone, L. Fornari, P. Mannella, M. Luisi, F. Labrie, A.R. Genazzani, Genomic and nongenomic mechanisms of nitric oxide synthesis induction in endothelial cells by a fourth generation selective estrogen modulator. Endo- crinology 143 (2002) 2052e2061. [21] D. Fulton, J.P. Gratton, W.C. Sessa, Post-translational control of endothelial nitric oxide synthase: why isn’t calcium/calmodulin enough? J. Pharmacol. Exp. Ther. 299 (2001) 818e824. [22] D. Fulton, P. Gratton, T.J. McCabe, J. Fontana, Y. Fujio, K. Walsh, T.F. Franke, A. Papapetropoulos, W.C. Sessa, Regulation of endothelium derived nitric oxide production by protein kinase Akt. Nature 399 (1999) 597e601. [23] S. Dimmeler, I. Fleming, B. Fisslthaler, C. Hermann, R. Busse, A.M. Zeiher, Activation of nitric oxide synthase in endothelial cells by Akt-dependent phosphorylation. Nature 399 (1999) 601e605. [24] J. Igarashi, H.S. Thatte, P. Prabhakar, D.E. Golan, T. Michel, Calcium-inde- pendent activation of endothelial nitric oxide synthase by ceramide. Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 12583e12588. [25] M.A. Marletta, Nitric oxide synthase: aspects concerning structure and catalysis. Cell 78 (1994) 927e930. [26] M.J. Kuchan, J.A. Frangos, Role of calcium and calmodulin in flow-induced nitric oxide production in endothelial cells. Am. J. Physiol. Cell Physiol. 266 (1994) C628eC636. [27] I. Fleming, J. Bauersachs, B. Fisslthaler, R. Busse, Ca2þ -independent activation of the endothelial nitric oxide synthase in response to tyrosine phosphatase inhibitors and fluid shear stress. Circ. Res. 82 (1998) 686e695. [28] W.C. Sessa, eNOS at a glance. J. Cell Sci. 17 (2004) 2427e2429. [29] T.J. McCabe, D. Fulton, L.J. Roman, W.C. Sessa, Enhanced electron flux and reduced calmodulin dissociation may explain “calcium-independent” eNOS activation by phosphorylation. J. Biol. Chem. 275 (2000) 6123e6128. [30] J. Mombouli, P. Vanhoutte, Kinins and endothelial control of vascular smooth muscle. Annu. Rev. Pharmacol. Toxicol. 35 (1995) 679e705. [31] D. Jagnandan, W.C. Sessa, D. Fulton, Intracellular location regulates calcium- calmodulin-dependent activation of organelle-restricted eNOS. Am. J. Phys- iol. Cell Physiol. 289 (2005) 1024e1033. [32] W.C. Sessa, K. Pritchard, N. Seyedi, et al., Chronic exercise in dogs increases coronary vascular nitric oxide production and endothelial cell nitric oxide synthase gene expression. Circ Res. 74 (1994) 349e353. [33] E.M. Prence, Effects of calcium on phosphatidylserine- and saposin C-stim- ulated glucosylceramide beta-glucosidase activity. Biochem. J. 310 (1995) 571e575. [34] D. Gattullo, P. Pagliaro, N.A. Marsh, G. Losano, New insights into nitric oxide and coronary circulation. Life Sci. 65 (1999) 2167e2174. [35] A. Mülsch, E. Bassenge, R. Busse, Nitric oxide synthesis in endothelial cytosol: evidence for a calcium-dependent and a calcium-independent mechanism. Naunyn Schmiedeberg’s Arch. Pharmacol. 340 (1989) 767e770. [36] K. Ayajiki, M. Kindermann, M. Hecker, I. Fleming, R. Busse, Intracellular pH and tyrosine phosphorylation but not calcium determine shear stress- induced nitric oxide production in native endothelial cells. Circ. Res. 78 (1996) 750e758. [37] I. Fleming, J. Bauersachs, R. Busse, Calcium-dependent and independent activation of the endothelial NO synthase. J. Vasc. Res. 34 (1997) 165e174. [38] T. Michel, Phosphorylation and subcellular translocation of endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 90 (1993) 6252e6256. [39] M.A. Corson, N.L. James, S.E. Latta, R.M. Nerem, B.C. Berk, D.G. Harrison, Phosphorylation of endothelial nitric oxide synthase in response to fluid shear stress. Circ Res. 79 (1996) 984e991. [40] G. Garcia-Cardena, R. Fan, D.F. Stern, J. Liu, W.C. Sessa, Endothelial nitric oxide is regulated by tyrosine phosphorylation and interacts with caveolin-1. J. Biol. Chem. 271 (1996) 27237e27240. [41] B. Gallis, G.L. Corthals, D.R. Goodlett, H. Ueba, F. Kim, S.R. Presnell, D. Figeys, D.G. Harrison, B.C. Berk, R. Aebersold, M.A. Corson, Identification of flow dependent endothelial nitric-oxide synthase phosphorylation sites by mass spectrometry and regulation of phosphorylation and nitric oxide production by the phosphatidylinositol 3-kinase inhibitor LY294002. J. Biol. Chem. 274 (1999) 30101e30108. [42] E. Butt, A. M.BernhardtSmolenski, P. Kotsonis, L.G. Frohlich, A. Sickmann, H.E. Meyer, S.M. Lohmann, H.H. Schmidt, Endothelial nitric-oxide synthase (type III) is activated and becomes calcium independent upon phosphory- lation by cyclic nucleotide-dependent protein kinases. J. Biol. Chem. 275 (2000) 5179e5187. [43] Z.P. Chen, K.I. Mitchelhill, B.J. Michell, D. Stapleton, I. Rodriguez-Crespo, L.A. Witters, D.A. Power, P.R. Ortiz de Montellano, B.E. Kemp, AMP-activated protein kinase phosphorylation of endothelial NO synthase. FEBS Lett. 443 (1999) 285e289. [44] I. Fleming, B. Fisslthaler, S. Dimmeler, B.E. Kemp, R. Busse, Phosphorylation of Thr495 regulates Ca2þ /calmodulin-dependent endothelial nitric oxide syn- thase activity. Circ. Res. 88 (2001) 68e75. [45] B.J. Michell, J.E. Griffiths, K.I. Mitchelhill, I. Rodriguez-Crespo, T. Tiganis, S. Bozinovski, P.R. de Montellano, B.E. Kemp, R.B. Pearson, The Akt kinase signals directly to endothelial nitric oxide synthase. Curr. Biol. 9 (1999) 845e848. [46] B.J. Michell, Z. Chen, T. Tiganis, D. Stapleton, F. Katsis, D.A. Power, A.T. Sim, B.E. Kemp, Coordinated control of endothelial nitric-oxide synthase phos- phorylation by protein kinase C and the cAMP-dependent protein kinase. J. Biol. Chem. 276 (2001) 17625e17628. [47] M.B. Harris, H. Ju, V.J. Venema, H. Liang, R. Zou, B.J. Michell, Z.P. Chen, B.E. Kemp, R.C. Venema, Reciprocal phosphorylation and regulation of endothelial nitric-oxide synthase in response to bradykinin stimulation. J. Biol. Chem. 276 (2001) 16587e16591. [48] R. Kou, D. Greif, T. Michel, Dephosphorylation of endothelial nitric-oxide synthase by vascular endothelial growth factor. Implications for the vascular responses to cyclosporin A. J. Biol. Chem. 277 (2002) 29669e29673. [49] D. Fulton, J.E. Church, L. Ruan, C. Li, S.G. Sood, B.E. Kemp, et al., Src kinase activates endothelial nitric oxide synthase by phosphorylating Tyr 83. J. Biol. Chem. 43 (2005) 35943e35952. [50] D. Fulton, L. Ruan, S.G. Sood, C. Li, Q. Zhang, R.C. Venema, Agonist-stimulated endothelial nitric oxide synthase activation and vascular relaxation: role of eNOS phosphorylation at Tyr83. Circ. Res. 102 (2008) 497e504. [51] M. Okuda, M. Takahashi, J. Suero, C.E. Murry, O. Traub, H. Kawakatsu, B.C. Berk, Shear stress stimulation of p130(cas) tyrosine phosphorylation requires calcium-dependent c-Src activation. J. Biol. Chem. 274 (1999) 26803e26809. [52] L.K. Tai, M. Okuda, J. Abe, C. Yan, B.C. Berk, Fluid shear stress activates proline-rich tyrosine kinase via reactive oxygen species-dependent pathway. Arterioscler. Thromb. Vasc. Biol. 22 (2002) 1790e1796. [53] B. Fisslthaler, A.E. Loot, A. Mohamed, R. Busse, I. Fleming, Inhibition of endothelial nitric oxide synthase activity by proline-rich tyrosine kinase 2 in response to fluid shear stress and insulin. Circ. Res. 102 (2008) 1520e1528. [54] D.S. Bredt, S.H. Snyder, Isolation of nitric oxide synthetase, a calmodulin requiring enzyme. Proc. Natl. Acad. Sci. U.S.A. 87 (1990) 682e685. [55] U. Forstermann, et al., Calmodulin-dependent endothelium-derived relaxing factor/nitric oxide synthase activity is present in the particulate and cytosolic fractions of bovine aortic endothelial cells. Proc. Natl. Acad. Sci. U.S.A. 88 (1991) 1788e1792. [56] W.N. Durán, A. Seyama, K. Yoshimura, D.R. González, P.I. Jara, X.F. Figueroa, M.P. Boric, Stimulation of NO production and of eNOS phosphorylation in the microcirculation in vivo. Microvasc. Res. 60 (2000) 104e111. [57] Y. Li, J. Zheng, I.M. Bird, R.R. Magness, Mechanisms of shear stress-induced endothelial nitric-oxide synthase phosphorylation and expression in ovine fetoplacental artery endothelial cells. Biol. Reprod. 70 (2004) 785e796. [58] R.S. Scotland, M. Morales-Ruiz, Y. Chen, J. Yu, R.D. Rudic, D. Fulton, J.P. Gratton, W.C. Sessa, Functional reconstitution of endothelial nitric oxide synthase reveals the importance of serine 1179 in endothelium-dependent vasomotion. Circ. Res. 90 (2002) 904e910. [59] Y. Wang, S. Nagase, A. Koyama, Stimulatory effect of IGF-I and VEGF on eNOS message, protein expression, eNOS phosphorylation and nitric oxide production in rat glomeruli, and the involvement of PI3-K signaling pathway. Nitric Oxide 10 (2004) 25e35. [60] S.W. Bae, H.S. Kim, Y.N. Cha, Y.S. Park, S.A. Jo, I. Jo, Rapid increase in endo- thelial nitric oxide production by bradykinin is mediated by protein kinase A signaling pathway. Biochem. Biophys. Res. Commun. 306 (2003) 981e987. [61] B.J. Michell, M.B. Harris, Z.P. Chen, H. Ju, et al., Identification of regulatory sites of phosphorylation of the bovine endothelial nitric-oxide synthase at serine 617 and serine 635. J. Biol. Chem. 277 (2002) 42344e42351. [62] H. Chen, M. Montagnani, T. Funahashi, I. Shimomura, M.J. Quon, Adiponectin stimulates production of nitric oxide in vascular endothelial cells. J. Biol. Chem. 278 (2003) 45021e45026. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981194
  • 10. [63] O. Rocks, A. Peyker, M. Kahms, P.J. Verveer, C. Koerner, M. Lumbierres, J. Kuhlmann, H. Waldmann, A. Wittinghofer, P.I. Bastiaens, An acylation cycle regulates localization and activity of palmitoylated Ras isoforms. Science 307 (2005) 1746e1752. [64] G. Sowa, J. Liu, A. Papapetropoulos, M. Rex-Haffner, T.E. Hughes, W.C. Sessa, Trafficking of endothelial nitric-oxide synthase in living cells: quantitative evidence supporting the role of palmitoylation as a kinetic trapping mech- anism limiting membrane diffusion. J. Biol. Chem. 274 (1999) 22524e22531. [65] T. Sakoda, K. Hirata, R. Kuroda, N. Miki, M. Suematsu, S. Kawashima, M. Yokoyama, Myristoylation of endothelial cell nitric oxide synthase is important for extracellular release of nitric oxide. Mol. Cell Biochem. 152 (1995) 143e148. [66] L. Busconi, T. Michel, Endothelial nitric oxide synthase. N-terminal myr- istoylation determines subcellular localization. J. Biol. Chem. 268 (1993) 8410e8413. [67] I. Liu, T.E. Hughes, W.C. Sessa, The first 35 amino acids and fatty acylation sites determine the molecular targeting of endothelial nitric oxide synthase into the Golgi region of cells: a green fluorescent protein study. J. Cell Biol. 137 (1997) 1525e1535. [68] L.J. Robinson, T. Michel, Mutagenesis of palmitoylation sites in endothelial nitric oxide synthase identifies a novel motif for dual acylation and subcel- lular targeting. Proc. Natl. Acad. Sci. U.S.A. 92 (1995) 11776e11780. [69] L.J. Robinson, L. Busconi, T. Michel, Agonist-modulated palmitoylation of endothelial nitric oxide synthase. J. Biol. Chem. 270 (1995) 995e998. [70] R.G. Parton, Caveolae and caveolins. Curr. Opin. Cell Biol. 8 (1996) 542e548. [71] E.J. Smart, G.A. Graf, M.A. McNiven, W.C. Sessa, J.A. Engelman, P.E. Scherer, T. Okamoto, M.P. Lisanti, Caveolins, liquid-ordered domains, and signal transduction. Mol. Cell. Biol. 19 (1999) 7289e7304. [72] G. Garcia-Cardena, P. Oh, J. Liu, J.E. Schnitzer, W.C. Sessa, Targeting of nitric oxide synthase to endothelial cell caveolae via palmitoylation: implications for nitric oxide signaling. Proc. Natl. Acad. Sci. U.S.A. 93 (1996) 6448e6453. [73] O. Feron, L. Belhassen, L. Kobzik, T.W. Smith, R.A. Kelly, T. Michel, Endothelial nitric oxide synthase targeting to caveolae: specific interactions with cav- eolin isoforms in cardiac myocytes and endothelial cells. J. Biol. Chem. 271 (1996) 22810e22814. [74] G. Garcia-Cardena, P. Martasek, B.S. Masters, P.M. Skidd, J. Couet, S. Li, M.P. Lisanti, W.C. Sessa, Dissecting the interaction between nitric oxide synthase (NOS) and caveolin: functional significance of the NOS caveolin binding domain in vivo. J. Biol. Chem. 272 (1997) 25437e25440. [75] S. Bulotta, A. Cerullo, R. Barsacchi, C.D. Palma Rotiroti, E. Clementi, N. Borgese, Endothelial nitric oxide synthase is segregated from caveolin-1 and localizes to the leading edge of migrating cells. Exp. Cell Res. 312 (2006) 877e889. [76] N.A. Maniatis, V. Brovkovych, S.E. Allen, T.A. John, A.N. Shajahan, C. Tiruppathi, S.M. Vogel, R.A. Skidgel, A.B. Malik, R.D. Minshall, Novel mechanism of endothelial nitric oxide synthase activation mediated by caveolae internalization in endothelial cells. Circ. Res. 99 (2006) 870e877. [77] H.F. Heijnen, S. Waaijenborg, J.D. Crapo, R.P. Bowler, J.W. Akkerman, J.W. Slot, Colocalization of eNOS and the catalytic subunit of PKA in endo- thelial cell junctions: a clue for regulated NO production. J. Histochem. Cytochem. 52 (2004) 1277e1285. [78] G. Sowa, M. Pypaert, W.C. Sessa, Distinction between signaling mechanisms in lipid rafts vs. caveolae. Proc. Natl. Acad. Sci. U.S.A. 98 (2001) 14072e14077. [79] I. Jiang, D. Cyr, R.W. Babbitt, W.C. Sessa, C. Patterson, Chaperone-dependent regulation of endothelial nitric-oxide synthase intracellular trafficking by the cochaperone/ubiquitin ligase CHIP. J. Biol. Chem. 278 (2003) 49332e49341. [80] P.I. Nedvetsky, W.C. Sessa, H.H. Schmidt, There’s NO binding like NOS binding:proteineprotein interactions in NO/cGMP signaling. Proc. Natl. Acad. Sci. U.S.A. 99 (2002) 16510e16512. [81] U. Zabel, C. Kleinschnitz, P. Oh, P. Nedvetsky, A. Smolenski, H. Müller, P. Kronich, P. Kugler, U. Walter, J.E. Schnitzer, H.H. Schmidt, Calcium- dependent membrane association sensitizes soluble guanylyl cyclase to nitric oxide. Nat. Cell Biol. 4 (2002) 307e311. [82] W.C. Sessa, G. Garcia-Cardena, J. Liu, A. Keh, J.S. Pollock, J. Bradley, S. Thiru, I.M. Braverman, K.M. Desai, The golgi association of endothelial nitric oxide synthase is necessary for the efficient synthesis of nitric oxide. J. Biol. Chem. 270 (1995) 17641e17644. [83] P.W.Shaul,E.J.Smart,L.J.Robinson,Z.German,I.S.Yuhanna,Y.Ying,R.G.Anderson, T. Michel, Acylation targets endothelial nitric-oxide synthase to plasmalemmal caveolae. J. Biol. Chem. 271 (1996) 6518e6522. [84] I. Hecker, A. Mulsch, E. Bassenge, U. Förstermann, R. Busse, Subcellular localization and characterization of nitric oxide synthase(s) in endothelial cells: physiological implications. Biochem. J. 299 (1994) 247e252. [85] V. Rizzo, D.P. McIntosh, P. Oh, J.E. Schnitzer, In situ flow activates endothelial nitric oxide synthase in luminal caveolae of endothelium with rapid caveolin dissociation and calmodulin association. J. Biol. Chem. 273 (1998) 34724e34729. [86] Y. Feng, V.J. Venema, R.C. Venema, N. Tsai, R.B. Caldwell, VEGF induces nuclear translocation of Flk-1/KDR, endothelial nitric oxide synthase, and caveolin-1 in vascular endothelial cells. Biochem. Biophys. Res. Commun. 256 (1999) 192e197. [87] R.M. Goetz, H.S. Thatte, P. Prabhakar, M.R. Cho, T. Michel, D.E. Golan, Estra- diol induces the calcium-dependent translocation of endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 2788e2793. [88] J.B. Michel, O. Feron, D. Sacks, T. Michel, Reciprocal regulation of endothelial nitric-oxide synthase by Ca2þ -calmodulin and caveolin. J. Biol. Chem. 272 (1997) 15583e15586. [89] P. Prabhakar, H.S. Thatte, R.M. Goetz, M.R. Cho, D.E. Golan, T. Michel, Receptor-regulated translocation of endothelial nitric-oxide synthase. J. Biol. Chem. 273 (1998) 27383e27388. [90] V.J. Venema, M.B. Marrero, R.C. Venema, Bradykinin-stimulated protein tyrosine phosphorylation promotes endothelial nitric oxide synthase trans- location to the cytoskeleton. Biochem. Biophys. Res. Commun. 226 (1996) 703e710. [91] Q. Wang, W.F. Patton, H.B. Hechtman, D. Shepro, A novel anti-inflammatory peptide inhibits endothelial cell cytoskeletal rearrangement, nitric oxide synthase translocation, and paracellular permeability increases. J. Cell Physiol. 172 (1997) 171e182. [92] S. Kai, D. Toomre, Lipid rafts and signal transduction. Nat. Rev. 1 (2000) 31e43. [93] F.A. Sánchez, N.B. Savalia, R.G. Durán, B.K. Lal, M.P. Boric, W.N. Durán, Functional significance of differential eNOS translocation. Am. J. Physiol. Heart Circ. Physiol. 291 (2006) 1058e1064. [94] B. Yang, T.N. Oo, V. Rizzo, Lipid rafts mediate H2O2 prosurvival effects in cultured endothelial cells. FASEB J. 20 (2006) 1501e1503. [95] C. Wann-Hansson, I.R. Hallberg, B. Risberg, A. Lundell, R. Klevsgard, Health- related quality of life after revascularization for peripheral arterial occlusive disease: long-term follow-up. J. Adv. Nurs. 51 (2005) 227e235. [96] I. Jialal, S. Devaraj, S.K. Venugopal, C-reactive protein: risk marker or medi- ator in atherothrombosis? Hypertension 44 (2004) 6e11. [97] V. Pasceri, J.T. Willerson, E.T. Yeh, Direct proinflammatory effect of C-reactive protein on human endothelial cells. Circulation 102 (2000) 2165e2168. [98] S.K. Venugopal, S. Devaraj, I. Yuhanna, P. Shaul, I. Jialal, Demonstration that C-reactive protein decreases eNOS expression and bioactivity in human aortic endothelial cells. Circulation 106 (2002) 1439e1441. [99] A. Napp, K. Brixius, C. Pott, C. Ziskoven, B. Boelck, U. Mehlhorn, R.H. Schwinger, W. Bloch, Effects of the beta3-adrenergic agonist BRL 37344 on endothelial nitric oxide synthase phosphorylation and force of contraction in human failing myocardium. J. Card. Fail. 15 (2009) 57e67. [100] S. Tanaka, Y. Toh, R. Mori, K. Komori, K. Okadome, K. Sugimachi, Possible role of cytomegalovirus in the pathogenesis of inflammatory aortic diseases: a preliminary report. J. Vasc. Surg. 16 (1992) 274e279. [101] S. Tanaka, K. Komori, K. Okadome, K. Sugimachi, R. Mori, Detection of active cytomegalovirus infection in inflammatory aortic aneurysms with RNA polymerase chain reaction. J. Vasc. Surg. 20 (1994) 235e243. [102] S. Pampou, S.N. Gnedoy, V.B. Bystrevskaya, V.N. Smirnov, E.I. Chazov, J.L. Melnick, et al., Cytomegalovirus genome and the immediateeearly antigen in cells of different layers of human aorta. Virchows Arch. 436 (2000) 539e552. [103] S.J. Duffy, J.F. Keaney Jr., M. Holbrook, N. Gokce, P.L. Swerdloff, B. Frei, J.A. Vita, Acute and chronic tea consumption reverses endothelial dysfunc- tion in patients with coronary artery disease. Circulation 104 (2001) 151e156. [104] E. Anter, S.R. Thomas, E. Schulz, O.M. Shapira, J.A. Vita, J.F. Keaney Jr., Acti- vation of endothelial nitric-oxide synthase by the p38 MAPK in response to black tea polyphenols. J. Biol. Chem. 279 (2004) 46637e46643. [105] R. Hambrecht, V. Adams, S. Erbs, A. Linke, N. Krankel, Y. Shu, Y. Baither, S. Gielen, H. Thiele, J.F. Gummert, F.W. Mohr, G. Schuler, Regular physical activity improves endothelial function in patients with coronary artery disease by increasing phosphorylation of endothelial nitric oxide synthase. Circulation 107 (2003) 3152e3158. [106] B. Ashley, J. Billig, L. Ambrecht, M. Wolfert, Robin Looft-Wilson, Preservation of EDHF-dependent vasodilation and connexin expression in hyper- homocysteinemic mice. FASEB J. 21 (2007) 759e763. [107] X. Jiang, F. Yang, H. Tan, D. Liao, R.M. Bryan Jr., J.K. Randhawa, R.E. Rumbaut, W. Durante, A.I. Schafer, X. Yang, H. Wang, Hyperhomocystinemia impairs endothelial function and eNOS activity via PKC activation. Arterioscler. Thromb. Vasc. Biol. 25 (2005) 2515e2521. [108] C. Robin, B.S. Looft-Wilson, Ashley, et al., Chronic diet-induced hyper- homocysteinemia impairs eNOS regulation in mouse mesenteric arteries. Am. J. Physiol. Regul. Integr. Comp. Physiol. 295 (2008) 59e66. [109] M.G. Signorello, A. Segantin, M. Passalacqua, G. Leoncini, Homocysteine decreases platelet NO level via protein kinase C activation. Nitric Oxide 20 (2009) 104e113. [110] I. Scherrer-Crosbie, R. Ullrich, K.D. Bloch, H. Nakajima, B. Nasseri, H.T. Aretz, M.L. Lindsey, A.C. Vancon, P.L. Huang, R.T. Lee, W.M. Zapol, M.H. Picard, Endothelial nitric oxide synthase limits left ventricular remodeling after myocardial infarction in mice. Circulation 104 (2001) 1286e1291. [111] C.L. Sartorio, D. Fraccarollo, M. Leutke, J. Bauersachs, The selective aldoste- rone receptor antagonist eplerenone improves vasomotor dysfunction and vascular oxidative stress early after myocardial infarction. Circulation 110 (Suppl. II) (2005) 262. [112] D. Fraccarollo, P. Galuppo, S. Hildemann, M. Christ, G. Ertl, J. Bauersachs, Additive improvement of left ventricular remodeling and neurohormonal activation by aldosterone receptor blockade with eplerenone and ACE inhibition in rats with myocardial infarction. J. Am. Coll. Cardiol. 42 (2003) 1666e1673. [113] J.W. Calvert, S. Gundewar, S. Jha, J.J.M. Greer, W.H. Bestermann, R. Tian, David J. Lefer, Acute metformin therapy confers cardioprotection against G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1195
  • 11. myocardial infarction via AMPK-eNOS-mediated signaling. Diabetes 57 (2008) 696e705. [114] C. Iadecola, Bright and dark sides of nitric oxide in ischaemic brain injury. Trends Neurosci. 20 (1997) 132e139. [115] A.F. Samdani, et al., Nitric oxide synthase in models of focal ischemia. Stroke 28 (1997) 1283e1288. [116] Z. Huang, et al., Enlarged infarcts in endothelial nitric oxide synthase knockout mice are attenuated by nitro-L-arginine. J. Cereb. Blood Flow Metab. 16 (1996) 981e987. [117] T. Dalkara, et al., Blood flow-dependent functional recovery in a rat model of focal cerebral ischemia. Am. J. Physiol. 267 (1994) 678e683. [118] K. Osuka, Y. Watanabe, N. Usuda, A. Nakazawa, M. Tokuda, M.J. Yoshida, Modification of endothelial no synthase through protein phosphorylation after forebrain cerebral ischemia/reperfusion. Stroke(35) (2004) 2582. [119] A. Hashiguchi, S. Yano, M. Morioka, J. Hamada, M. Kochi, K. Fukunaga, Dephosphorylation of eNOS on Thr495 after transient forebrain ischemia in gerbil hippocampus. Brain Res. Mol. Brain Res. 133 (2005) 317e319. [120] D.N. Atochin, A. Wang, V.W. Liu, J.D. Critchlow, A.P. Dantas, R. Looft-Wilson, T. Murata, S. Salomone, H.K. Shin, C. Ayata, M.A. Moskowitz, T. Michel, W.C. Sessa, P.L. Huang, The phosphorylation state of eNOS modulates vascular reactivity and outcome of cerebral ischemia in vivo. J. Clin. Invest. 117 (2007) 1961e1967. [121] K. Izuishi, A. Tsung, M.A. Hossain, M. Fujiwara, H. Wakabayashi, T. Masaki, Ischemic preconditioning of the murine liver protects through the Akt kinase pathway. Hepatology 44 (2006) 573e580. [122] Y. Wang, Y. Vodovotz, P.K. Kim, R. Zamora, T.R. Billiar, Mechanisms of hepatoprotection by nitric oxide. Ann. N.Y. Acad. Sci. 962 (2002) 415e422. [123] H. Taniai, I.N. Hines, S. Bharwani, R.E. Maloney, Y. Nimura, B. Gao, S.C. Flores, J.M. McCord, M.B. Grisham, T.Y. Aw, Susceptibility of murine periportal hepatocytes to hypoxia-reoxygenation: role for NO and Kupffer cell-derived oxidants. Hepatology 39 (2004) 1544e1552. [124] F. Roviezzo, S. Cuzzocrea, A. Di Lorenzo, V. Brancaleone, E. Mazzon, R. Di Paola, M. Bucci, G. Cirino, Protective role of PI3-kinase-Akt-eNOS signalling pathway in intestinal injury associated with splanchnic artery occlusion shock. Br. J. Pharmacol. 151 (2007) 377e383. [125] Y. Shi, J.E. Baker, C. Zhang, J.S. Tweddell, J. Su, A. Kirkwood, A. Pritchard Jr., Chronic hypoxia increases endothelial nitric oxide synthase generation of nitric oxide by increasing heat shock protein 90 association and serine phosphorylation. Circ. Res. 91 (2002) 300e306. [126] R.F. Furchgott, J.V. Zawadzki, The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature 288 (1980) 373e376. [127] T. Murata, K. Sato, M. Hori, H. Ozaki, H. Karaki, Decreased endothelial nitric- oxide synthase (eNOS) activity resulting from abnormal interaction between eNOS and its regulatory proteins in hypoxia-induced pulmonary hyperten- sion. J. Biol. Chem. 277 (2002) 44085e44092. [128] F. Coulet, S. Nadaud, M. Agrapart, F. Soubrier, Identification of hypoxia- response element in the human endothelial nitric oxide synthase gene promoter. J. Biol. Chem. 278 (2003) 46230e46240. [129] J. Liu, Y. Gao, S. Negash, L.D. Longo, J.U. Raj, Long-term effects of prenatal hypoxia on endothelium-dependent relaxation responses in pulmonary arteries of adult sheep. Am. J. Physiol. Lung Cell. Mol. Physiol. 296 (2009) 547e554. [130] K.I. Rother, Diabetes treatmentdbridging the divide. N. Engl. J. Med. 356 (2007) 1499e1501. [131] L.M. Tierney, S.J. McPhee, M.A. Papadakis, Current Medical Diagnosis & Treatment, International Edition. Lange Medical Books, McGraw-Hill, New York, 2002, 1203e1215. [132] D.W. Laight, M.J. Carrier, E.E. Anggard, Antioxidants, diabetes and endothelial dysfunction. Cardiovasc. Res. 47 (2000) 457e464. [133] B. Tesfamariam, M.L. Brown, D. Deykin, et al., Elevated glucose promotes generation of endothelium-derived vasoconstrictor prostanoids in rabbit aorta. J. Clin. Invest. 85 (1990) 929e932. [134] H.G. Bohlen, J.M. Lash, Topical hyperglycemia rapidly suppresses EDRF- mediated vasodilation of normal rat arterioles. Am. J. Physiol. 265 (1993) 219e225. [135] S. Meraji, L. Jayakody, M.P. Senaratne, et al., Endothelium-dependent relax- ation in aorta of BB rat. Diabetes 36 (1987) 978e981. [136] G.M. Pieper, D.A. Meier, S.R. Hager, Endothelial dysfunction in a model of hyperglycemia and hyperinsulinemia. Am. J. Physiol. 269 (1995) 845e850. [137] M.T. Johnstone, S.J. Creager, K.M. Scales, et al., Impaired endothelium- dependent vasodilation in patients with insulin-dependent diabetes melli- tus. Circulation 88 (1993) 2510e2516. [138] S.B. Williams, J.A. Cusco, M.A. Roddy, et al., Impaired nitric oxide-mediated vasodilation in patients with non-insulin-dependent diabetes mellitus. J. Am. Coll. Cardiol. 27 (1996) 567e574. [139] P. Clarkson, D.S. Celermajer, A.E. Donald, et al., Impaired vascular reactivity in insulin-dependent diabetes mellitus is related to disease duration and low density lipoprotein cholesterol levels. J. Am. Coll. Cardiol. 28 (1996) 573e579. [140] G.E. McVeigh, G.M. Brennan, G.D. Johnston, et al., Impaired endothelium- dependent and independent vasodilation in patients with type 2 (non- insulin-dependent) diabetes mellitus. Diabetologia 35 (1992) 771e776. [141] X.L. Du, D. Edelstein, S. Dimmeler, Q. Ju, C. Sui, M. Brownlee, Hyperglycemia inhibits endothelial nitric oxide synthase activity by posttranslational modification at the Akt site. J. Clin. Invest. 108 (2001) 1341e1348. [142] E.B. Okon, A.W. Chung, P. Rauniyar, E. Padilla, T. Tejerina, B.M. McManus, H. Luo, C. van Breemen, Compromised arterial function in human type 2 diabetic patients. Diabetes 54 (2005) 2415e2423. [143] H. Okumura, N. Nagaya, T. Itoh, I. Okano, J. Hino, K. Mori, Y. Tsukamoto, H. Ishibashi-Ueda, S. Miwa, K. Tambara, S. Toyokuni, C. Yutani, K. Kangawa, Adrenomedullin infusion attenuates myocardial ischemia/reperfusion injury through the phosphatidylinositol 3-kinase/Akt-dependent pathway. Circu- lation 109 (2004) 242e248. [144] S. Yang, L. Lin, J.X. Chen, C.R. Lee, J.M. Seubert, Y. Wang, H. Wang, Z.R. Chao, D.D. Tao, J.P. Gong, Z.Y. Lu, D.W. Wang, D.C. Zeldin, Cytochrome P-450 epoxygenases protect endothelial cells from apoptosis induced by tumor necrosis factor-alpha via MAPK and PI3K/Akt signaling pathways. Am. J. Physiol. Heart Circ. Physiol. 293 (2007) 142e151. [145] K. Kawasaki, R.S. Smith Jr., C.M. Hsieh, J. Sun, J. Chao, J.K. Liao, Activation of the phosphatidylinositol 3-kinase/protein kinase Akt pathway mediates nitric oxide-induced endothelial cell migration and angiogenesis. Mol. Cell Biol. 23 (2003) 5726e5737. [146] Y. Wang, X. Wei, X. Xiao, R. Hui, J.W. Card, M.A. Carey, D.W. Wang, D.C. Zeldin, Arachidonic acid epoxygenase metabolites stimulate endothelial cell growth and angiogenesis via mitogen-activated protein kinase and phosphatidyli- nositol 3-kinase/Akt signaling pathways. J. Pharmacol. Exp. Ther. 314 (2005) 522e532. [147] R.R. Shankar, Y. Wu, H.Q. Shen, J.S. Zhu, A.D. Baron, Mice with gene disruption of both endothelial and neuronal nitric oxide synthase exhibit insulin resistance. Diabetes 49 (2000) 684e687. [148] H. Duplain, R. Burcelin, C. Sartori, S. Cook, M. Egli, M. Lepori, P. Vollenweider, T. Pedrazzini, P. Nicod, B. Thorens, U. Scherrer, Insulin resistance, hyperlip- idemia, and hypertension in mice lacking endothelial nitric oxide synthase. Circulation 104 (2001) 342e345. [149] S. Cook, O. Hugli, M. Egli, B. Menard, S. Thalmann, C. Sartori, C. Perrin, P. Nicod, B. Thorens, P. Vollenweider, U. Scherrer, R. Burcelin, Partial gene deletion of endothelial nitric oxide synthase predisposes to exaggerated high-fat diet-induced insulin resistance and arterial hypertension. Diabetes 53 (2004) 2067e2072. [150] G. Zeng, F.H. Nystrom, L.V. Ravichandran, L.N. Cong, M. Kirby, H. Mostowski, M.J. Quon, Roles for insulin receptor, PI3-kinase, and Akt in insulin-signaling pathways related to production of nitric oxide in human vascular endothelial cells. Circulation 101 (2000) 1539e1545. [151] M. Montagnani, H. Chen, V.A. Barr, M.J. Quon, Insulin-stimulated activation of eNOS is independent of Ca2þ but requires phosphorylation by Akt at Ser (1179). J. Biol. Chem. 276 (2001) 30392e30398. [152] B. Musicki, M.F. Kramer, R.E. Becker, L. Arthur Burnett, Inactivation of phosphorylated endothelial nitric oxide synthase (Ser-1177) by O-GlcNAc in diabetes-associated erectile dysfunction. Proc. Natl. Acad. Sci. U.S.A 102 (2005) 11870e11875. [153] J.W. Elrod, M.R. Duranski, W. Langston, J.J. Greer, L. Tao, T.R. Dugas, C.G. Kevil, H.C. Champion, D.J. Lefer, eNOS gene therapy exacerbates hepatic ischemia- reperfusion injury in diabetes: a role for eNOS uncoupling. Circ. Res. 99 (2006) 78e85. [154] D.S. Marks, J.A. Vita, J.D. Folts, J.F.J. Keaney, G.N. Welch, J. Loscalzo, Inhibition of neointimal proliferation in rabbits after vascular injury by a single treat- ment with a protein adduct of nitric oxide. J. Clin. Invest. 96 (1995) 2630e2638. [155] K.S. Masters, E.A. Lipke, E.E. Rice, M.S. Liel, H.A. Myler, C. Zygourakis, D.A. Tulis, J.L. West, Nitric oxide-generating hydrogels inhibit neointima formation. J. Biomater. Sci. Polym. Ed. 16 (2005) 659e672. [156] M.R. Kapadia, L.W. Chow, N.D. Tsihlis, S.S. Ahanchi, J.W. Eng, J. Murar, J. Martinez, D.A. Popowich, Q. Jiang, J.A. Hrabie, J.E. Saavedra, L.K. Keefer, J.F. Hulvat, S.I. Stupp, M.R. Kibbe, Nitric oxide and nanotechnology: a novel approach to inhibit neointimal hyperplasia. J. Vasc. Surg. 47 (2008) 173e182. [157] Y. Li, P.I. Lee, Controlled nitric oxide delivery platform based on s-nitro- sothiol conjugated interpolymer complexes for diabetic wound healing. Mol. Pharm. 7 (2010) 254e266. [158] S.S. Ahanchi, V.N. Varu, N.D. Tsihlis, J. Martinez, C.G. Pearce, M.R. Kapadia, Q. Jiang, J.E. Saavedra, L.K. Keefer, J.A. Hrabie, M.R. Kibbe, Heightened efficacy of nitric oxide-based therapies in type II diabetes mellitus and metabolic syndrome. Am. J. Physiol. Heart Circ. Physiol. 295 (2008) 2388e2398. [159] C.G. Pearce, S.F. Najjar, M.R. Kapadia, J. Murar, J. Eng, B. Lyle, O.O. Aalami, Q. Jiang, J.A. Hrabie, J.E. Saavedra, L.K. Keefer, M.R. Kibbe, Beneficial effect of a short-acting NO donor for the prevention of neointimal hyperplasia. Free Radic. Biol. Med. 44 (2008) 73e81. [160] T. Nakagawa, W. Sato, O. Glushakova, M. Heinig, T. Clarke, M. Campbell- Thompson, Y. Yuzawa, M.A. Atkinson, R.J. Johnson, B. Croker, Diabetic endothelial nitric oxide synthase knockout mice develop advanced diabetic nephropathy. J. Am. Soc. Nephrol 18 (2007) 539e550. [161] J. Molnar, S. Yu, N. Mzhavia, C. Pau, I. Chereshnev, H.M. Dansky, Diabetes induces endothelial dysfunction but does not increase neointimal formation in high-fat diet fed C57BL/6J mice. Circ. Res. 96 (2005) 1178e1184. [162] S. Pennathur, J.W. Heinecke, Oxidative stress and endothelial dysfunction in vascular disease. Curr. Diab. Rep. 7 (2007) 257e264. [163] U. Hink, H. Li, H. Mollnau, M. Oelze, E. Matheis, M. Hartmann, M. Skatchkov, F. Thaiss, R.A. Stahl, A. Warnholtz, T. Meinertz, K. Griendling, D.G. Harrison, U. Forstermann, T. Munzel, Mechanisms underlying endothelial dysfunction in diabetes mellitus. Circ. Res. 88 (2001) 14e22. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e11981196
  • 12. [164] K. Chalupsky, H. Cai, Endothelial dihydrofolate reductase: critical for nitric oxide bioavailability and role in angiotensin II uncoupling of endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 102 (2005) 9056e9061. [165] N.J. Alp, S. Mussa, J. Khoo, S. Cai, T. Guzik, A. Jefferson, N. Goh, K.A. Rockett, K.M. Channon, Tetrahydrobiopterin-dependent preservation of nitric oxidemediated endothelial function in diabetes by targeted transgenic GTPcyclohydrolase I overexpression. J. Clin. Invest. 112 (2003) 725e735. [166] J.B. Laursen, M. Somers, S. Kurz, L. McCann, A. Warnholtz, B.A. Freeman, M. Tarpey, T. Fukai, D.G. Harrison, Endothelial regulation of vasomotion in apoE-deficient mice: implications for interactions between peroxynitrite and tetrahydrobiopterin. Circulation 103 (2001) 1282e1288. [167] U. Landmesser, S. Dikalov, S.R. Price, L. McCann, T. Fukai, S.M. Holland, W.E. Mitch, D.G. Harrison, Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J. Clin. Invest. 111 (2003) 1201e1209. [168] T. Munzel, A. Daiber, V. Ullrich, A. Mulsch, Vascular consequences of endo- thelial nitric oxide synthase uncoupling for the activity and expression of the soluble guanylyl cyclase and the cGMP-dependent protein kinase. Arte- rioscler. Thromb. Vasc. Biol. 25 (2005) 1551e1557. [169] M.B. Harris, H. Ju, V.J. Venema, H. Liang, R. Zou, B.J. Michell, et al., Reciprocal phosphorylation and regulation of endothelial nitric-oxide synthase in response to bradykinin stimulation. J. Biol. Chem. 276 (2001) 16587e16591. [170] C.A. Chen, L.J. Druhan, S. Varadharaj, Y.R. Chen, J.L. Zweier, Phosphorylation of endothelial nitric-oxide synthase regulates superoxide generation from the enzyme. J. Biol. Chem. 283 (2008) 27038e27047. [171] Y. O’Hara, T.E. Peterson, D.G. Harrison, Hypercholesterolemia increases endothelial superoxide anion production. J. Clin. Invest. 91 (1993) 2546e2551. [172] T.E. Peterson, V. Poppa, H. Ueba, A. Wu, C. Yan, B.C. Berk, Opposing effects of reactive oxygen species and cholesterol on endothelial nitric oxide synthase and endothelial cell caveolae. Circ. Res. 85 (1999) 29e37. [173] U. Landmesser, H. Cai, S. Dikalov, L. McCann, J. Hwang, et al., Role of p47 (phox) in vascular oxidative stress and hypertension caused by angiotensin II. Hypertension 40 (2002) 511e515. [174] S.R. Thomas, K. Chen, J.F. Keaney Jr., Hydrogen peroxide activates en- dothelial nitric-oxide synthase through coordinated phosphorylation and dephosphorylation via a phosphoinositide 3- kinase-dependent signaling pathway. J. Biol. Chem. 277 (2002) 6017e6024. [175] H. Zhang, J. Zhang, Z. Ungvari, C. Zhang, Resveratrol improves endothelial function: role of TNF {alpha} and vascular oxidative stress. Arterioscler. Thromb. Vasc. Biol. 29 (2009) 1164e1171. [176] W. Insull, The pathology of atherosclerosis: plaque development and plaque responses to medical treatment. Am. J. Med. 122 (2009) 3e14. [177] C.D. Liapis, E.D. Avgerinos, N.P. Kadoglou, J.D. Kakisis, What a vascular surgeon should know and do about atherosclerotic risk factors. J. Vasc. Surg. 49 (2009) 1348e1354. [178] M.E. Safar, P. Jankowski, Central blood pressure and hypertension: role in cardiovascular risk assessment. Clin. Sci. 116 (2009) 273e282. [179] C.M. Werner, M. Böhm, The therapeutic role of RAS blockade in chronic heart failure. Ther. Adv. Cardiovasc. Dis. 2 (2008) 167e177. [180] K.K. Gaddam, A. Verma, M. Thompson, R. Amin, H. Ventura, Hypertension and cardiac failure in its various forms. Med. Clin. North Am. 93 (2009) 665e680. [181] E. Agabiti-Rosei, From macro- to microcirculation: benefits in hypertension and diabetes. J. Hypertens. 26 (2008) 15e21. [182] B.P. Murphy, T. Stanton, F.G. Dunn, Hypertension and myocardial ischemia. Med. Clin. North Am. 93 (2009) 681e695. [183] L. Tylicki, B. Rutkowski, Hypertensive nephropathy: pathogenesis, diagnosis and treatment. Pol. Merkuriusz. Lek. 14 (2003) 168e173. [184] L.D. Truong, S.S. Shen, M.H. Park, B. Krishnan, Diagnosing nonneoplastic lesions in nephrectomy specimens. Archiv. Pathol. Lab Med. 133 (2009) 189e200. [185] R.E. Tracy, S. White, A method for quantifying adrenocortical nodular hyperplasia at autopsy: some use of the method in illuminating hyperten- sion and atherosclerosis. Ann. Diagn. Pathol. 6 (2002) 20e29. [186] D.R. Singer, A. Kite, Management of hypertension in peripheral arterial disease: does the choice of drugs matter? Eur. J. Vasc. Endovasc. Surg. 35 (2008) 701e708. [187] W.S. Aronow, Hypertension and the older diabetic. Clin. Geriatr. Med. 24 (2008) 489e501. [188] A.W. Gardner, A. Afaq, Management of lower extremity peripheral arterial disease. J. Cardiopulm. Rehab. Prev. 28 (6) (2008) 349e357. [189] P.A. Cahill, C. Foster, E.M. Redmond, C. Gingalewski, Y. Wu, J.V. Sitzmann, Enhanced nitric oxide synthase activity in portal hypertensive rabbits. Hepatology 22 (1995) 598e606. [190] P.Y. Martin, D.L. Xu, M. Niederberger, A. Weigert, P. Tsai, J. St John, P. Gines, R.W. Schrier, Upregulation of endothelial constitutive NOS: a major role in the increased NO production in cirrhotic rats. Am. J. Physiol. Renal Fluid Electrolyte Physiol. 270 (1996) 494e499. [191] M. Niederberger, P. Gines, P.Y. Martin, P. Tsai, K. Morris, I. McMurtry, R.W. Schrier, Comparison of vascular nitric oxide production and systemic hemodynamics in cirrhosis versus prehepatic portal hypertension in rats. Hepatology 24 (1996) 947e951. [192] M. Niederberger, P.Y. Martin, P. Gines, K. Morris, P. Tsai, D.L. Xu, I. McMurtry, R.W. Schrier, Normalization of nitric oxide production corrects arterial vasodilation and hyperdynamic circulation in cirrhotic rats. Gastroenter- ology 109 (1995) 1624e1630. [193] Y. Iwakiri, M. Tsai, J. Timothy, T.J. McCabe, J.P. Gratton, D. Fulton, J.R. Groszmann, W.C. Sessa, Phosphorylation of eNOS initiates excessive NO production in early phases of portal hypertension. Am. J. Physiol. Heart Circ. Physiol. 282 (2002) 2084e2090. [194] V. Shah, Molecular mechanisms in the pathogenesis of cirrhotic portal hypertension: focus on nitric oxide. J. Gastroenterol. Hepatol. 19 (2004) 145e149. [195] T.J. Bivalacqua, F.Mustafa Usta, C.H. Champion, J.P. Kadowitz, J.Wayne G. Hellstrom, Endothelial dysfunction in erectile dysfunction: role of the endothelium in erectile physiology and disease. J. Androl. 24 (2003). [196] B. Musicki, A.l. Burnett, eNOS function and dysfunction in the penis. Exp. Biol. Med. 231 (2006) 154e165. [197] J. Hidetaka, E. Masato, M.R. Kano, S. Ogawa, K. Iijima, M. Akishita, Y. Ouchi, Cilostazol inhibits oxidative stress-induced premature senescence via upregulation of sirt1 in human endothelial cells. Arterioscler. Thromb. Vasc. Biol. 28 (2008) 1634. [198] T. Minamino, H. Miyauci, T. Yoshida, Y. Ishida, H. Yoshida, I. Komuro, Endothelial cell senescence in human atherosclerosis: role of telomere in endothelial dysfunction. Circulation 105 (2002) 1541e1544. [199] K.F. Burrig, The endothelium of advanced arteriosclerotic plaques in humans. Arterioscler. Thromb. 11 (1991) 1678e1689. [200] T. Hayashi, H. Matsui-Hirai, A. Miyazaki-Akita, A. Fukatsu, J. Funami, Q.F. Ding, S. Kamalanathan, Y. Hattori, L.J. Ignarro, A. Iguchi, Endothelial cellular senescence is inhibited by nitric oxide: implications in atherosclerosis asso- ciated with menopause and diabetes. Proc. Natl. Acad. Sci. U.S.A. 103 (2006) 17018e17023. [201] M. Vasa, K. Breitschopf, A.M. Zeiher, S. Dimmeler, Nitric oxide activates telomerase and delays endothelial cell senescence. Circ. Res. 87 (2000) 540e542. [202] J. Kambayashi, Y. Liu, B. Sun, Y. Shakur, M. Yoshitake, F. Czerwiec, Cilostazol as a unique antithrombotic agent. Curr. Pharm. Des. 9 (2003) 2289e2302. [203] T. Hayashi, K. Yano, H. Matsui-Hirai, H. Yokoo, Y. Hattori, A. Iguchi, Nitric oxide and endothelial cellular senescence. Pharmacol. Ther. 120 (2008) 333e339. [204] A. Hashimoto, G. Miyakoda, Y. Hirose, T. Mori, Activation of endothelial nitric oxide synthase by cilostazol via a cAMP/protein kinase A- and phosphati- dylinositol 3-kinase/Akt-dependent mechanism. Atherosclerosis 189 (2006) 350e357. [205] M. Braunstein, A.B. Rose, S.G. Holmes, C.D. Allis, J.R. Broach, Transcriptional silencing in yeast is associated with reduced nucleosome acetylation. Genes Dev. 7 (1993) 592e604. [206] H. Vaziri, S.K. Dessain, N.E. Eaton, S.I. Imai, R.A. Frye, T.K. Pandita, L. Guarente, R.A. Weinberg, hSIR2 (SIRT1) functions as an NAD-dependent p53 deacety- lase. Cell 107 (2001) 149e159. [207] E. Langley, M. Pearson, M. Faretta, U.M. Bauer, R.A. Frye, S. Minucci, P.G. Pelicci, T. Kouzarides, Human SIR2 deacetylates p53 and antagonizes PML/p53- induced cellular senescence. EMBO J. 21 (2002) 2383e2396. [208] F. Picard, M. Kurtev, N. Chung, A. Topark-Ngarm, T. Senawong, R. Machado De Oliveira, M. Leid, M.W. McBurney, L. Guarente, Sirt1 promotes fat mobilization in white adipocytes by repressing PPAR-gamma. Nature 429 (2004) 771e776. [209] E. Nisoli, C. Tonello, A. Cardile, V. Cozzi, R. Bracale, L. Tedesco, S. Falcone, A. Valerio, O. Cantoni, E. Clementi, S. Moncada, M.O. Carruba, Calorie restriction promotes mitochondrial biogenesis by inducing the expression of eNOS. Science 310 (2005) 314e317. [210] M. Potente, S. Dimmeler, NO targets SIRT1: a novel signaling network in endothelial senescence. Arterioscler. Thromb. Vasc. Biol. 28 (2008) 1577e1579. [211] Y. Korkmaz, The basal phosphorylation sites of endothelial nitric oxide synthase at serine (Ser)1177, Ser116, and threonine (Thr)495 in rat molar epithelial rests of malassez. J. Periodontol. 76 (2005) 1513e1519. [212] M.T. Gentile, C. Vecchione, A. Maffei, A. Aretini, G. Marino, R. Poulet, L. Capobianco, G. Selvetella, G. Lembo, Mechanisms of soluble b-amyloid impairment of endothelial function. J. Biol. Chem. 279 (2004) 48135e48142. [213] J.S. Stamler, L. Jia, J.P. Eu, T.J. McMahon, I.T. Demchenko, J. Bonaventura, K. Gernert, C.A. Piantadosi, Blood flow regulation by S-nitro- sohemoglobin in the physiological oxygen gradient. Science 276 (1997) 2034e2037. [214] J. Kleinbongard, R. Schulz, T. Rassaf, et al., Red blood cells express a func- tional endothelial nitric oxide synthase. Blood 107 (2006) 2943e2951. [215] T. Anuk, E.B. Assayag, R. Rotstein, R. Fusman, D. Zeltser, S. Berliner, D. Avitzour, I. Shapira, N. Arber, N.M. Bornstein, Prognostic implications of admission inflammatory profile in acute ischemic neurological events. Acta Neurol. Scand. 106 (2002) 196e199. [216] P. Muda, P. Kampus, M. Zilmer, K. Zilmer, C. Kairane, T. Ristimäe, K. Fischer, R. Teesalu, Homocysteine and red blood cell glutathione as indices for middle-aged untreated essential hypertension patients. J. Hypertens. 21 (2003) 2329e2333. [217] D. Mihov, J. Vogel, M. Gassmann, A. Bogdanova, Erythropoietin activates nitric oxide synthase in murine erythrocytes. Am. J. Physiol. Cell. Physiol. 297 (2009) 378e388. G.K. Kolluru et al. / Biochimie 92 (2010) 1186e1198 1197