Download Nitric oxide cell signaling: S-nitrosation of Ras superfamily GTPases

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Hedgehog signaling pathway wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Cellular differentiation wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Apoptosome wikipedia , lookup

Cytokinesis wikipedia , lookup

Amitosis wikipedia , lookup

Signal transduction wikipedia , lookup

JADE1 wikipedia , lookup

List of types of proteins wikipedia , lookup

Paracrine signalling wikipedia , lookup

Gaseous signaling molecules wikipedia , lookup

Transcript
Cardiovascular Research 75 (2007) 229 – 239
www.elsevier.com/locate/cardiores
Review
Nitric oxide cell signaling: S-nitrosation of Ras superfamily GTPases
Kimberly W. Raines a,b , Marcelo G. Bonini c , Sharon L. Campbell a,b,⁎
a
Department of Biochemistry and Biophysics, University of North Carolina, 530 Mary Ellen Jones Building, Chapel Hill, NC 27599-7260, United States
Lineberger Comprehensive Cancer Center, University of North Carolina, 530 Mary Ellen Jones Building, Chapel Hill, NC 27599-7260, United States
c
Laboratory of Pharmacology and Chemistry, National Institute of Environmental Health Sciences,
National Institutes of Health, Research Triangle Park, NC 27709, United States
b
Received 29 December 2006; received in revised form 13 April 2007; accepted 18 April 2007
Available online 24 April 2007
Time for primary review 33 days
Abstract
The Ras superfamily of small GTPases cycle between inactive GDP-bound and active GTP-bound states to modulate a diverse array of
processes involved in cellular growth control. While the basic mechanisms by which GTPase regulatory proteins regulate GTPase substrates
have been revealed through numerous studies, detailed studies into the mechanism(s) of free radical-mediated GTPase regulation have only
more recently been tackled. This article reviews the mechanism of free radical-mediated GTPase regulation and shows nitric oxide can serve
as important regulator of small GTPase proteins (i.e. Ras and RhoA) through protein modifications such as S-nitrosation.
© 2007 European Society of Cardiology. Published by Elsevier B.V. All rights reserved.
Keywords: Nitric oxide; Redox signaling; Small GTPases
1. Ras superfamily: Introduction
The Ras superfamily consists of a number of small
monomeric GTPases including Ras, Rho and Rab subfamilies. These GTPases cycle between active GTP- and inactive
GDP-bound states to regulate a diverse array of biological
processes including cell proliferation, apoptosis, vesicular
and nuclear transport [1,2]. Given their critical cellular roles,
the GDP- and GTP-bound states of Ras superfamily
GTPases are highly regulated by multiple cellular factors.
Guanine nucleotide exchange factors (GEFs) [3–6], facilitate
exchange of GDP with GTP to produce the active GTPbound form of the GTPase, whereas GTPase-activating
proteins (GAPs) enhance the slow intrinsic rate of GTP
hydrolysis to produce the inactive GDP-bound form of the
GTPase [5,7,8]. In addition, guanine nucleotide inhibitors
⁎ Corresponding author. Lineberger Comprehensive Cancer Center,
University of North Carolina, 530 Mary Ellen Jones Building, Chapel
Hill, NC 27599-7260, United States. Tel.: +1 919 966 7139; fax: +1 919 966
2852.
E-mail address: [email protected] (S.L. Campbell).
(GDIs) down-regulate the activity of a subset of GTPases
(e.g., Rho and Rab subfamilies) by preventing membrane
association as well as inhibiting guanine nucleotide dissociation [3,5,9,10]. In addition to these regulatory proteins,
the free radical nitric oxide (NO), has been shown to react
with Ras and other Ras-related GTPases to regulate their
activity [11]. While the basic mechanisms by which GEFs,
GAPs and GDIs regulate their respective GTPase substrates
have been revealed through numerous studies, detailed
studies to investigate the mechanism(s) of free radicalmediated GTPase regulation have only recently been tackled
since the emergence of protein modifications, such as
nitrosation and glutathionation. Nitrosation is often used
interchangeably with nitrosylation, but differs slightly
because by definition nitrosation is the covalent attachment
of a NO group to an amine, thiol or hydroxyl residue, while
nitrosylation is the addition of NO without a change in
formal charge of the substrate.
A number of GTPases including DexRas [12], Rap1A
[11,13,14], Ran [15], Rac1, Cdc42, RhoA [16], Rab3A [11]
and H-, K- and N-Ras [17] contain redox-active residue(s) that
are sensitive to NO, particularly the NO derivative, nitrogen
0008-6363/$ - see front matter © 2007 European Society of Cardiology. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cardiores.2007.04.013
230
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
U
dioxide ( NO2). We have investigated the mechanism by
which NO modulates the activity for several of these GTPases,
with particular focus on the Ras GTPase, and found that
U
treatment of Ras with NO in the presence of O2 or NO2
produces a Ras-thiyl radical intermediate resulting in GDP
dissociation from Ras [11,17]. This, in turn, can lead to activation or inactivation of Ras depending on the concentration of
NO and oxidizing potential of the medium.
U
Consistent with our proposed mechanism of NO2-mediated
U
Ras guanine nucleotide dissociation [11], reaction of NO2 with
glutathione (GSH) and cysteine (CysSH) has been shown to
U
produce a glutathionyl radical (GS ) and cysteinyl radical
U
U
(Cys–S ), respectively [18–20]. In addition, NO2-mediated
S-nitrosation of GSH and CysSH proceeds through formaU
U
tion of a thiyl radical (i.e., GS and Cys–S ) and this process
was shown to be dominant over N2O3− or NO-mediated Snitrosation [18–20]. As we had previously observed that
S-nitrosation of Ras does not alter its structure or biochemical
U
activity [21], we speculated that NO2-mediated guanine
nucleotide release occurs through a radical propagation mechanism involving Ras thiyl radical conversion to a Ras–GDP
guanine radical. The guanine base is particularly sensitive to
reaction with free radicals [22–27], and formation of a guanine
radical is likely to alter interactions with Ras resulting in release
of Ras bound GDP [11]. Our present data is consistent with this
mechanism of redox mediated GDP dissociation for Ras as well
as other Ras-related proteins [11,14,17,28,29]. Moreover, we
have observed that other radical species (i.e., superoxide
U
(O2 −)) act on redox active sites within select Ras superfamily
GTPases to promote dissociation of GDP via a similar mechanism [11,14,28]. In this context, the mechanism of free radicalmediated GTPase regulation and the target specificity of the
free radical toward the redox-active GTPases are reviewed.
2. Source of NO in the vasculature
NO is produced endogenously by a family of nitric oxide
synthase [30] enzymes in a 2-step oxidation of L-arginine to
NO and L-citrulline [31]. Three distinct isoforms of NOS
have been characterized [32]: neuronal NOS (nNOS) and
endothelial NOS (eNOS), which are constitutively expressed
and activated by influx of intracellular calcium. A third
isoform, inducible NOS (iNOS), is regulated by transcriptional activation in response to inflammation and cytokine
production. Vascular NO is derived primarily from eNOS in
the endothelial cell; however, iNOS can also be induced in
many cell types including endothelial and vascular smooth
muscle cells, by cytokine stimulation [33,34].
3. In vivo nitrosative cell signaling
One of the best characterized targets of NO is the heme of
guanylate cyclase. Nitrosation of guanylate cyclase leads to
activation of the second messenger, cyclic guanosine
monophosphate (cGMP) which regulates a wide variety of
physiological functions, including vascular tone, platelet
aggregation, inflammation, neurotransmission, gastric emptying, and hormone release [35,36].
Cysteine residues in proteins are also well known targets
of NO. Thiol nitrosation has been detected in cell culture
based assays under a myriad of conditions [37–40], and can
alter both protein expression and function [41]. For example,
cellular analyses have demonstrated that S-nitrosation can
inhibit caspases [42,43], c-Jun-N-terminal kinase [44,45],
protein tyrosine phosphatases [46,47], and ornithine decarboxylase [48]. In addition, hemoglobin S-nitrosation increases oxygen affinity and modulates vasodilatation [49,50].
Table 1
Rate constants for UNO, UNO2 and RSU with select physiologically relevant compounds
Reaction
Rate constant
U
2 NO þ O
U NO
2
U
→2 NO
2
U
kf = 1.1 × 10 M s
kr = 8.0 × 104 M− 1s− 1
k ∼ 2.0 × 105
k = 6.6 – 9.4 × 104 M− 1s− 1
N2 O3 þ H2 O→2NO−2 þ 2Hþ
N2 O3 þ GSH→GSNO þ NO−2 þ Hþ
U
[114]
−1 −1
9
þ NO←→N2 O3
2
Reference
k = 2 × 106 M− 1s− 1
U
[120]
[121,122]
3 −1
k = 1.6 × 10 s
[64]
−1 −1
7
GSH þ NO2 →GS þ NO−2 þ Hþ
k = 2 × 10 M s
[123]
GS þ NO→GSNO
k = 1 × 109 M− 1s− 1
[64]
U U
U
U
GS þ GS →GSSG
9
k = 1.5 × 10 M s
U
U
GS þ GS− →GSSG −
U− þ O
GSSG
2 →GSSG
−1 −1
þ O2
U−
U
U
NO þ O − →ONOO−
2
ONOOH=ONOO− þ GSH→GSOH þ NO−2
[124]
−1 −1
kf = 6.6 × 10 – 6.0 × 10 M s
kr = 1.6 × 105 M− 1s− 1
k = 5 × 109 M− 1s− 1
7
8
9
−1 −1
k = 6.7–19 × 10 M s
−1 −1
k = 660 M s
[124]
[124]
[125,126]
[114]
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
Vascular functions regulated by S-nitrosation include ventilation, heart rate, and blood pressure [51].
S-nitrosation can promote post-translational modification
of protein thiols in select proteins to modulate their activity.
However, NO can also regulate the activity of proteins by
mechanisms that are independent of S-nitrosation. Our
studies on the Ras GTPase indicate that it is cysteine–thiol
radical formation rather that S-nitrosation, that alters Ras
activity [14,28]. Although NO can react with protein thiols
through both radical and non-radical based mechanisms, we
will focus herein on the role of the radical pathway in
regulation of Ras superfamily GTPase activity.
4. NO-mediated thiol nitrosation, the radical pathway
glutathione has recently been re-determined and was shown
to be more than one order of magnitude lower than that for
N2O3 hydrolysis [64]. This fact combined with the excess
water available in aqueous solution makes thiol nitrosation
via N2O3 less likely in hydrophilic environments.
Thus the pKa, location and three dimensional environment of the cysteine within the protein, are fundamental in
determining the nitrosothiol yield as well as the prevailing
mechanism [65]. Thiol nitrosation may be summarized by
the following set of reactions:
U
U
2 NO þ O2 →2 NO2
UNO þ NO →N O
2
NO is a stable free radical gas, which is hydrophobic in
nature and extremely small in size. These properties, together
with its high reactivity with other radicals and metal centers,
render a freely diffusible and selective messenger that can
modulate a multitude of physiological processes [30,52–59].
However, NO is a poor nitrosating compound. As physiological levels of molecular oxygen are required for nitrosothiol
formation from NO [18,19,60], increasing attention has been
U
devoted to the autoxidative product of NO, NO2, as a cellular
nitrosating agent. Both NO and oxygen tend to concentrate in
non-polar environments (e.g. plasma membrane) which makes
U
NO2 production (reaction 1) particularly efficient in hydroU
phobic compartments. NO2 is a moderate oxidant [61] with
extremely high reaction rate constants for its reaction with
biothiols (k = 107 M− 1s− 1) (Table 1). The product formed in
this reaction is invariably the biothiol-derived thiyl radical
U
(i.e., RS) which is nitrosated by recombination with NO in a
diffusion controlled reaction (reactions 1–3, Table 1) [62].
2 NO þ O2 Y2 NO2
d
d
d NO
2
þ R SHYRS þ NO
2
d
k ¼ 106 108 M1 s1
RS þ NO YRSNOðnitrosothiolÞ
d d
k ¼ 109 M1 s1
ðReaction1Þ
5. NO-mediated thiol nitrosation, the ionic pathway
U
Once produced, NO2 may also recombine with available
NO to produce N2O3, in an extremely fast diffusion
controlled radical–radical recombination reaction (Table 1).
N2O3 is considered to be a potent electrophilic compound
capable of nitrosating thiol moieties in peptides and proteins
[63]. However N2O3 is also subject to hydrolysis via – OH
attack on the electrophilic nitrogen atoms, resulting in NO2−
U
generation [63]. Similar to NO2, O2 and NO, N2O3 is likely
to be more soluble in hydrophobic environments where it is
U
stable and in equilibrium with its precursors NO and NO2
[64]. Importantly, the rate constant for N2O3 reaction with
231
2
3
N2 O3 ¼þ ON…NO−2 ðIonic pathwayÞ
U
U
RSH þ NO2 →RS þ NO−2 ðRadical pathwayÞ
U U
RS þ NO→RSNOðRadical pathwayÞ
U
U
RS þ RS →RSSRðRadical termination reactionÞ
6. NO and regulation of Ras and Ras-like GTPases in vivo
In the instance of the GTPase protein family, several
GTPases contain a redox-sensitive NKCD motif, and we
have shown that three distinct GTPases in the sub-class of
Ras superfamily GTPases (i.e., H-Ras, Rap1A and Rab3A)
containing this motif, can be regulated by NO and other
redox active species [14]. To date, NO-mediated redox
regulation is best characterized for Ras proteins.
Several cell based assays have shown that NO can react
with Ras, promote guanine nucleotide exchange (GNE) and
Ras activation [66–70]. Although limited studies have been
conducted on the role of endogenously generated NO in
regulation of Ras function, support for such regulation in
cells exists. For example, Yun et al. have demonstrated that
NO derived from nNOS can lead to Ras activation in primary
rat cortical cells [71]. Their results suggested that Ras is
a physiologic target of endogenously produced NO through
a signaling pathway involving the N-methyl-D-aspartate
(NMDA) glutamate receptor, with NO-mediated Ras activation playing an important role in long-lasting neuronal
responses in these cells [71,72]. Intriguingly, it has recently
been shown that NO-induced activation of Ras causes inhibition of NOS [73], suggesting NO feedback regulation.
Lander and coworkers showed that human T cells treated
with exogenously generated NO, contain an increased
232
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
percentage of Ras in the GTP-bound form [69]. Furthermore,
NO-induced activation of Ras led to modulation of signal
transduction pathways downstream of Ras. For example,
NO-induced activation of Ras led to stimulation of the
mitogen-activated protein kinase (MAPK) pathway in rabbit
aortic endothelial cells, resulting in tyrosine phosphorylation
of various cellular proteins (i.e., focal adhesion kinase, Src
kinase, and MAPK) [74]. Moreover, Deora et al. have shown
that NO-producing agents trigger recruitment, Ras activation
of phosphatidylinositol 3′-kinase (PI3K) and Raf in human T
cells, leading to Ras-dependent activation of extracellular
signal regulated kinases (ERKs) [75]. They further identified a
NO-mediated Ras signaling pathway, whereby activation of
Ras by NO, recruits and activates Raf-1, leading to Elk-1
phosphorylation and tumor necrosis factor-α messenger RNA
induction, suggesting that modulation of Ras activity by NO in
T cells may play a critical role in Elk-1-induced activation of T
lymphocytes, host defense and inflammation [76]. Subsequent
work demonstrated that NO released endogenously from
NOS and exogenously from the NO-producing agent NOC18, modulates the expression of Hypoxia-Inducible Factor-1α
(HIF-1α) via the Ras-mediated MAPK pathway [77].
Consistent with these results, expression of a dominant-negative form of Ras significantly suppressed NO-mediated HIF1α activation [78]. Thus, the increased levels of HIF-1α under
hypoxic conditions in prostate cancer cells [77] may result
from NO-mediated Ras activation of the MAPK pathway [79].
Mallis et al. demonstrated activation of the MAPK pathway in NIH-3T3 fibroblasts via S-glutathiolation and S-nitrosation of multiple thiols in H-Ras by exogenous thiol oxidants
such as: hydrogen peroxide, S-nitrosoglutathione, diamide,
glutathione disulphide and cystamine [80]. Although direct
modification of Ras may mediate MAPK activation, it is
also probable that indirect mechanisms of MAPK pathway
activation occur, for example, by inhibition of associated
phosphatases [46].
In contrast to these findings, endogenous production of NO
has been shown to block calcium-induced activation of ERK
through a mechanism involving S-nitrosation of Ras in human
embryonic kidney cells over-expressing nNOS [81]. Moreover, addition of exogenous NO agents led to inhibition of HRas mediated ERK activation in mouse fibroblasts [89]. The
investigators speculated that H-Ras was inactivated under
these conditions by nitrosation of C-terminal cysteines in Ras,
which caused enhanced palmitoyl turnover and reduced association of Ras with membranes [89]. Moreover, it has been
suggested that vascular smooth-muscle cells treated with NO
donors inhibit epidermal growth factor-mediated Ras activation
of Raf-1 [82]. Thus, it appears that Ras can be either activated
or inactivated by NO depending on the cellular conditions.
7. NO-mediated regulation of Ras activity through
cysteine 118 radical formation
It is widely accepted that thiol modifications, such as
nitrosation, oxidation and glutathiolation are recurrent events
leading to key redox-based signals which modulate cellular
growth. All products mentioned above can be converted
back to the reduced thiol (RSH) form at the expense of
reduced cellular thiols such as glutathione (dithiols, sulfenic
acid) or ATP/thiols (sulfinic acid) [83,84].
Although the cysteinyl radical commonly participates in
reactions leading to several chemical modifications, it is rarely
proposed as the actual modulator of enzyme activation. This is
likely to happen, because contrary to the other species which
are relatively long-lived in vivo, thiyl radicals tend to disappear
quickly (milliseconds) and are not detectable in a variety of
systems. Thus, cysteinyl radicals are likely to be formed as
intermediaries of the nitrosation process.
We have recently shown that exposure of Ras to oneelectron oxidizing agents (oxidants that will generate protein
radical intermediates) induce Ras guanine nucleotide
dissociation, a crucial step in the regulation of Ras.
Moreover, we have recently conducted EPR experiments
that provide unequivocal detection of a Ras thiyl radical,
produced under exposure of the protein to NO in aerated
buffers. The thiyl radical production paralleled a marked
increase in the guanine nucleotide dissociation rate.
Importantly, Ras mutated at Cys118 or containing blocked
thiols were not subject to free radical induced GNE nor
produced an EPR detectable thiyl radical signal. Of note, Ras
nitrosated at Cys118 has been shown to exhibit similar
structure and GNE rates as the non-nitrosated form of Ras
[21]. Contrary to the belief that S-nitrosation leads to an
activated state of Ras, recent findings demonstrate that nitrosation is more likely to be a reversible end product of exposure to NO. As previously anticipated [85,86] these data
taken together, constitute compelling evidence for the involvement of a thiyl radical intermediate in modulation of
Ras activity.
Like Ras, other Ras-related proteins (i.e., Ran, DexRas) have
been shown to undergo S-nitrosation in the presence of both
endogenous and exogenous sources of NO [17,66,68,69,87,88].
Our results suggest that Ras thiol nitrosation is likely to be
mediated by the NO autoxidation product, nitrogen dioxide
radical (UNO2), to produce a Ras radical intermediate (Ras–
Cys118U). Further, our data indicate that formation of the Ras
radical intermediate is critical for O2U− or NO2U−mediated Ras
guanine nucleotide dissociation [11,14,17]. The mechanism is
U
fundamentally similar for both NO/O2 and O2 − in that either
UNO or O U− can react with the Ras–Cys118 thiol to produce a
2
2
Ras–Cys 118 -thiyl radical (Ras–Cys 118 U) intermediate
[11,14,17]. The Ras–Cys118U then electronically interacts with
Ras-bound GDP, possibly via the Phe28 side-chain, to initiate
radical-based conversion of Ras-bound GDP into a Ras-bound
U
GDP neutral radical (G –GDP) (Fig. 1) [11]. Ras-bound GU-DP
U
U
can further react with an additional NO2 or O2 − to produce
GDP–NO2 (5-nitro-GDP) or GDP–O2 (5-oxo-GDP) adducts,
U
respectively [11,14]. As formation of GDP+ and sequential
U
transformation to a G –GDP (neutral radical) will disrupt key
hydrogen bond interactions between Ras residues in the NKCD/
U
U
SAK motif and the guanine nucleotide base, the NO2 or O2 −
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
Fig. 1. Proposed radical based mechanism of redox mediated Ras guanine
U
nucleotide dissociation. Reaction of NO in the presence of O2 or NO2 with
118
the Ras Cys thiol, produces a Ras radical intermediate, proposed to be a
U
U
Ras thiyl radical (Ras–Cys118 ). Ras–Cys118 then electronically interacts
28
with Ras-bound GDP, possibly via the Phe side-chain, to initiate radicalbased conversion of Ras-bound GDP into a Ras-bound GDP neutral radical
U
U
U
(G –GDP). Ras-bound G –DP can further react with an additional NO2 to
+U
produce GDP–NO2 (5-nitro-GDP). As formation of GDP − and sequential
U
transformation to a G –GDP will disrupt key hydrogen bond interactions
between Ras residues in the NKCD/SAK motif and the guanine nucleotide
base. The n–π interaction between the neutral Ras Phe28 phenyl side-chain
and the GDP-bound guanine base is shown in wavy lines. Other Ras residues
involved in Ras GDP-binding interactions are omitted for clarity.
233
mediated thiyl–guanine radical propagation process promotes
GDP release from Ras (Fig. 1).
Similar to formation of small molecular weight nitrosothiols (i.e., GSNO), Ras–SNO can be produced by either a
U
N2O3 and/or an NO2-mediated process in the presence of NO.
U
However, our data support NO2-mediated production of SNO
via a radical mediated pathway in which reaction of Ras with
U
NO2 produces Ras–Cys118U, which can then react with NO to
generate Ras–SNO [17]. Moreover, S-nitrosation of Ras at
Cys118 does not promote Ras GNE [11]. Thus, we proposed
and provided evidence that a Ras–Cys118U intermediate, as
opposed to Ras–SNO, electronically couples with Ras-bound
GDP to facilitate Ras guanine nucleotide dissociation [11,14].
As GNE is required to generate the biologically active GTPbound form of Ras, redox mediated GDP dissociation must be
followed by GTP binding to promote Ras activation in situ.
U
U
While NO2 and O2 − can facilitate nucleotide dissociation
from Ras, nucleotide association does not occur in the absence
of a radical quenching agent. Addition of a radical quenching
U
U
agent (i.e., ascorbate) facilitates NO2 or O2 −-mediated Ras
GNE [11,14], most likely due to removal of Ras radical
intermediate(s). If NO serves as a quenching agent, the resultant
reaction product is S-nitrosated Ras (Ras–SNO) [11,14].
Although our studies indicate that Ras–SNO does not directly
promote Ras GNE, Ras nitrosation may function to prevent
further radical reaction processes, since the S-nitrosated Cys118
U
U
thiol moiety can not react with NO2, GS and NO to produce a
Ras radical species. If denitrosation of Ras–SNO occurs by
transfer of NO to other biomolecules, such as glutathione or
ascorbate, denitrosated Ras could then potentially react with
U
NO2 and a second round of GNE may ensue.
U
U
Prolonged incubation of NO2 or O2 − with Ras in the
presence of GTP but in the absence of a radical quenching
agent primarily induces formation of an inactive apo-Ras state,
as only a small fraction of the available Ras (∼10%) is present
as Ras GTP. These results imply that redox agents can serve to
terminate Ras signaling by promoting fast release of GDP or
GTP from Ras to produce an apo-Ras state which cannot
undergo Ras GNE in the absence of a radical scavenger. A
radical scavenger may then be necessary to recover inactive
apo-Ras into a GTP-binding form of Ras. Consistent with this
premise, we have observed efficient in vitro GTP re-association
in the presence of a radical quenching agent to produce an
active GTP-bound Ras (∼50%), which in turn, promotes
binding of Ras to the Ras binding domain of Raf-1 [14,89].
U
U
These results suggest that for NO2 and O2 − to promote
Ras GNE and Ras activation, radical scavenging agents may
be required. Hence, if Ras is exposed to physiological levels
of redox agents and reducing biomolecules are present at
high enough levels to function as radical scavenger agents
[90,91], Ras activation may occur. In contrast, exposure to
higher concentrations of redox agents and lower reducing
potential may result in Ras inactivation. This may occur
through distinct mechanisms (Fig. 2), as highlighted above,
or by modification of additional residues in Ras that
are partially protected, relative to Cys118. For example,
234
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
nitrosation of carboxyl-terminal cysteines in Ras, normally
targeted for palmitoylation, may lead to release of Ras from
the membrane and inhibition of Ras activity [89]. Moreover,
nitrosation of additional residues in Ras (i.e., tyrosine and
other cysteines) may result in Ras destabilization.
8. Redox regulation of NKCD containing Ras
superfamily GTPases
Ras superfamily GTPases encompass several gene
families that regulate a number of events in the eukaryotic
cell [92]. These proteins share amino acid identity with the
classical Ras proteins and have numerous structural features
in common. Intriguingly, the spatial orientation of the Phe28
and Cys118 side chains (Ras numbering) within the NKCD
motif of Rap1A and Rab3A is similar to that observed for
U
Ras [11]. Thus, a similar mechanism of NO2-mediated
guanine nucleotide dissociation is expected for these NKCDmotif containing GTPases. Consistent with this observation,
U
we have shown that NO/O2 and O2 − facilitate guanine
nucleotide dissociation from Rab3A and Rap1 (Fig. 3) [11].
9. Redox regulation of Rho GTPases
Similar to Ras GTPases, Rho family GTPases regulate a
host of cellular processes involved in regulation of cellular
growth control [93–95]. However, Rho proteins differ from
Ras in that they also regulate pathways that modulate cell
morphology and motility through actin cytoskeletal rearrangements as well as oxidant production [93,94,96,97]. The
Rab and Sar1/Arf subclasses are best known for their role in
vesicle trafficking, whereas the Ran sub-class of GTPases
is involved in regulation of nucleocytoplasmic transport
and microtubule organization [9,98–100]. As molecular
switches, these GTPases regulate both temporal and spatial
cellular functions [101].
Although Rho GTPases, like Ras, can regulate pathways
that control cellular growth, Rho GTPases interact with distinct
regulatory factors and cellular targets to mediate pathways that
control these processes [93–95]. We have identified a distinct
class of redox-active GTPases conserved in nearly 50% of
all Rho subfamily GTPases (Fig. 4), that contain a cysteine,
Cys18 (Rac1 numbering), located at the end of the P-loop
(GXXXXGK(S/T), residues 10 to 17, Rac1 numbering). This
cysteine-containing P-loop sequence is designated the
GXXXXGK(S/T)C motif. For comparison, members of the
Ras subfamily of GTPases (i.e., H–, K–, N–Ras and Rap1A)
possess a NKCD motif which forms interactions with the
guanine nucleotide base and contains a redox-active cysteine,
Cys118 (Ras numbering) [11,14,17,68]. Inspection of NMR
and X-ray crystal structures for GXXXXGK(S/T)C motifcontaining Rho GTPases (i.e., Rac1, RhoA and Cdc42) [102–
104], indicates that the Cys18 thiol (Rac1 numbering) in the
GXXXXGK(S/T)C motif is solvent accessible, suggesting that
reactive oxygen species and reactive nitrogen species [105] may
be able to target the Cys18 thiol.
U
U
We have recently shown that both NO2 and O2 − can
enhance the rate of guanine nucleotide dissociation for Rac1,
RhoA and Cdc42 GTPases containing the GXXXXGK(S/T)
C motif in vitro (Fig. 5) [16], and speculated that since the
Rho family GTPase co-localize and modulate NADPH
U
oxidase activity [96,106–108], that O2 − may stimulate
guanine nucleotide dissociation from Rac in vivo (23). The
U
fundamental mechanistic process for O2 −-mediated GNE
associated with GXXXXGK(S/T)C motif-containing
U
GTPases appears similar to that of O2 −-mediated GNE of
NKCD-containing GTPases [16]. The only difference lies in
the initial thiyl radical formation site of the GTPases.
Fig. 2. Differential activity of redox active Ras superfamily GTPases in response to reactive oxygen and nitrogen species under different cellular conditions. In vivo
redox potential mediates the cellular outcome of redox agents on redox active Ras superfamily GTPase activity. In a high cellular reducing environment, redox agents
are likely to activate redox active GTPases, whereas in a strong oxidizing environment redox agents more likely to promote inhibition. Small GTPases (i.e., RhoA)
containing a (GXXXCGK(S/T)C) motif, are more susceptible disulfide formation in the presence of redox agents and thus more prone to inactivation.
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
Fig. 3. Comparison of NO-mediated guanine nucleotide dissociation rates of
NKCD motif-containing small GTPases in the presence of O2. Rates of
apparent mant-GDP dissociation from the small GTPases (0.5 μM), Rab3A,
U
Rap1A, wtRas and Ras C118S, in the presences of NO2 were determined by
fitting the data to a simple exponential decay. The decrease in fluorescence
emission at 460 nm was recorded as a function of time. For comparison,
intrinsic GDP dissociation rates for Rap1A, Rab3A, wtRas, and the Ras
C118S variant were also measured [11]. Reproduced with permission of
Elsevier B.V. Copyright 2005.
This GXXXXGK(S/T)C motif is also found in other Rho
family GTPases (i.e., RhoT and TC10), several Rab GTPases
as well as the Ran family GTPase, TC4. Hence, we speculate
U
U
that O2 − and NO2 may modulate temporal and spatial
cellular functions of the GXXXXGK(S/T)C motif-containing Rab and Ran GTPases via a molecular mechanism
similar to that proposed for redox-active Rho GTPases.
Interestingly, a variant form of the GXXXXGK(S/T)C
motif, which contains an additional cysteine (Cys16, RhoA
numbering) designated as the ‘GXXXCGK(S/T)C motif’, is
found in a subset of redox-active GTPases, including RhoA
and RhoB. In RhoA, the redox-active Cys20 (equivalent to
Rac1 Cys18) is ∼ 3.6 and ∼ 10.3 Å away from Phe30
(equivalent to Phe28 in Ras and Rac1) and Cys16 (not present
Fig. 4. Spatial architecture of the Rac and Cdc42 GXXXX(S/T)C18 motif
and the RhoA GXXXX(S/T)C16 + C20 motif. The distance and spatial
orientation of the Phe28 side-chain, nucleotide ligand, and Cys18 thiol group
in the GXXXX(S/T)C18 motif of Rac and Cdc42 is presented. The distance
and spatial orientation of the Cys16 and Cys20 thiol group in the GXXXX(S/
T)C16 + C20 motif of RhoA is also presented. The scheme was generated
using PDB 1CRQ and RASMOL [109,119].
235
Fig. 5. Effects of redox agents on the rate of guanine nucleotide dissociation
from GXXXXGK(ST)C motif-containing Rho GTPases. To monitor O2mediated Rho GTPase GDP dissociation, xanthine oxidase generated O−2 ,
was accumulated in solution prior to incubation with 1 μM [3H]GDP-loaded
Rho GTPases (Rac1, RhoA, or Cdc42). Aliquots were withdrawn, spotted
onto nitrocellulose filters and radioactivity was determined using a
Beckman–Coulter scintillation counter. The resultant radioactivity values
of Rho GTPase-bound [3H]GDP were converted into the mole fraction of
nucleotide per mole of total Rho GTPase. Data presented represent mean
values from triplicate measurements. As a control, the intrinsic GDP
dissociation rates of Rho GTPases were also measured using identical
experimental conditions in the absence of redox agents. Similar results were
obtained in the presence of NO2U [16]. Reproduced with permission of the
American Society for Biochemistry and Molecular Biology Copyright 2005.
in Ras and Rac1), respectively [102]. Although the redox
architecture of RhoA [102] is nearly identical to that of Rac1
and Cdc42 [103,104], electronic interaction between Cys16
and Cys20 in RhoA may give rise to distinct redox properties
compared to Rho GTPases containing a single cysteine
within the GXXXXGK(S/T)C motif.
Consistent with this premise, we have shown that reaction
of RhoA with redox agents leads to different functional
consequences from that of Rac1 and Cdc42 due to the
presence of an additional cysteine (GXXXCGK(S/T)C) in
the redox-active motif [109]. While reaction of redox agents
with RhoA stimulates guanine nucleotide dissociation,
RhoA is subsequently inactivated through formation of an
intramolecular disulfide that prevents guanine nucleotide
binding thereby causing RhoA inactivation. Thus, redox
agents may function to down-regulate RhoA activity under
conditions that stimulate Rac1 and Cdc42 activity (Fig. 2).
The opposing functions of these GTPases may be due in part
to their differential redox regulation. In addition, we have
shown that the platinated-chemotherapeutic agent, cisplatin,
which is known for targeting nucleic acids, reacts with
RhoA, but not Rac1 or Cdc42, to produce a RhoA thiol–
cisplatin–thiol adduct, leading to inactivation of RhoA [16].
Thus, in addition to redox agents, platinated-chemotherapeutic agents also modulate the cellular activity of GTPases
containing the GXXXCGK(S/T)C motif.
According to the RhoA–GDP crystal structure [102], the
buried RhoA Cys16 thiol is ∼ 10.3 Å away from the solvent
236
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
accessible redox-active RhoA Cys 20 thiol, and the αphosphate of the bound GDP intercalates between these
two RhoA cysteine thiols. Therefore, RhoA Cys16 and Cys20
thiols in the RhoA–GDP complex are not well positioned for
disulfide formation or platin–thiol adduct formation. Thus,
U
U
given our observations that redox agents ( NO2 and O2 −)
promote RhoA guanine nucleotide dissociation, we speculated
that RhoA Cys16–Cys20 disulfide formation may require
release of the RhoA-bound guanine nucleotide to facilitate
interaction between the RhoA Cys16 and Cys20 thiols. Moreover, formation of a RhoA thiol–cisplatin–thiol adduct may
require a conformational rearrangement in RhoA, as reaction
of cisplatin with the Cys20 thiol may destabilize guanine
nucleotide binding interactions leading to release of RhoAbound GDP to produce GDP-deficient RhoA. Consistent with
this premise, we have observed that redox-mediated disulfide
formation as well as cisplatin thiol adduct formation results in a
form of RhoA that does not bind GDP or GTP [109]. These
results are perhaps not unexpected, given recent studies that
phenylarsine oxide inactivates RhoA by forming a bridge
between Cys20 and Cys16, resulting in a guanine nucleotide
deficient form of the protein [110].
In summary, we have identified and characterized a unique
redox-active motif present in Rho family GTPases that is
important for redox-mediated regulation of GNE activity in
vitro. Moreover, the radical-based molecular mechanism of
Rho GTPase GNE appears similar in nature to the mechanism
characterized for Ras GTPases. The redox-active motif is
present in other Ras superfamily GTPases (i.e., Rab GTPases),
suggesting that redox regulation of GTPase signaling is more
widespread than previously envisioned.
10. Future directions
Results from our studies on the Ras GTPase indicate that
NO can promote formation of a cysteine–thiol radical, which
alters Ras activity through a radical propagation mechanism
leading to guanine nucleotide oxidation and release of GDP
from Ras. However, S-nitrosation of this cysteine does not
affect Ras activity. Hence, the mechanism of NO-mediated
regulation of Ras activity is distinct from what has previously
been observed in other systems, in which protein S-nitrosation
alters cellular function. Thus, the studies described herein form
the basis for further investigation of cellular regulation of Ras
superfamily GTPases by redox agents. For example, it would
be helpful to better characterize conditions by which Ras is
activated or inactivated by redox agents in various cellular
conditions. It will also be of interest to understand how Ras
GEFs and GAPs cooperate in the presence of redox agents.
Moreover, we have shown that several Ras related GTPases
are redox active. In particular, redox active solvent accessible
cysteines are highly conserved in Rab GTPases, and we
speculate that Rab GTPases may be sensitive to redox control. Future studies will be required to investigate Rab redox
activity. Additionally, although we have demonstrated that
redox active Rho GTPases respond differently to redox agents
and co-localize with endogenous sources of redox agents,
it will be of interest to investigate the regulation of Rho
GTPases in vivo.
Thiols are major targets for biological oxidants because of
the low oxidizing potential of cysteine and their relatively
low pKa, which render them excellent nucleophiles. Whereas Ras nitrosation and other oxidized forms of the Ras thiol
may not modulate Ras activity directly, we speculate that
oxidation of the redox active cysteine in Rho GTPases may
alter GTPase activity due to its location in the phosphoryl
binding loop. In support of this premise, direct S-nitrosation
of RhoA has recently been shown to inhibit RhoA and
ERK signaling to inhibit vascular smooth muscle cell proliferation [111].
In the case of Ras, SNO modification does not appear to
directly correlate with changes in Ras biochemical activity.
Although other oxidation species of Ras have been observed in
cells (Ras SSG [112]), it is unclear whether these alternate
modifications directly alter Ras activity. Evidence suggesting
that thiyl radicals are recurrent reactive intermediates produced
during oxidative stress conditions is mounting [11,86,113–
115]. Thus, development of methods such as electron paramagnetic resonance coupled freeze-trap or spin-trap technologies for direct detection of thiyl radicals [114,116–118], rather
than SNO or other oxidative thiol end products, should prove
useful for detection of redox active GTPases whose activity is
modulated under a variety of different redox conditions.
Acknowledgements
We thank Jongyun Heo for intellectual contributions and
assistance with figures.
References
[1] Paduch M, Jelen F, Otlewski J. Structure of small G proteins and their
regulators. Acta Biochim Pol 2001;48:829–50.
[2] Takai Y, Sasaki T, Matozaki T. Small GTP-binding proteins. Physiol
Rev 2001;81:153–208.
[3] Karnoub AE, Symons M, Campbell SL, Der CJ. Molecular basis for
Rho GTPase signaling specificity. Breast Cancer Res Treat
2004;84:61–71.
[4] Sprang S. GEFs: master regulators of G-protein activation. Trends
Biochem Sci 2001;26:266–7.
[5] Geyer M, Wittinghofer A. GEFs, GAPs, GDIs and effectors: taking a
closer (3D) look at the regulation of Ras-related GTP-binding
proteins. Curr Opin Struct Biol 1997;7:786–92.
[6] Quilliam LA, Khosravi-Far R, Huff SY, Der CJ. Guanine nucleotide
exchange factors: activators of the Ras superfamily of proteins.
Bioessays 1995;17:395–404.
[7] Moon SY, Zheng Y. Rho GTPase-activating proteins in cell
regulation. Trends Cell Biol 2003;13:13–22.
[8] Sprang SR. G proteins, effectors and GAPs: structure and mechanism.
Curr Opin Struct Biol 1997;7:849–56.
[9] Pfeffer SR. Rab GTPases: specifying and deciphering organelle
identity and function. Trends Cell Biol 2001;11:487–91.
[10] Olofsson B. Rho guanine dissociation inhibitors: pivotal molecules in
cellular signalling. Cell Signal 1999;11:545–54.
[11] Heo J, Prutzman KC, Mocanu V, Campbell SL. Mechanism of free
radical nitric oxide-mediated Ras guanine nucleotide dissociation.
J Mol Biol 2005;346:1423–40.
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
[12] Cheah JH, Kim SF, Hester LD, Clancy KW, Patterson III SE,
Papadopoulos V, et al. NMDA receptor-nitric oxide transmission
mediates neuronal iron homeostasis via the GTPase Dexras1. Neuron
2006;51:431–40.
[13] Mittar D, Sehajpal PK, Lander HM. Nitric oxide activates Rap1 and
Ral in a Ras-independent manner. Biochem Biophys Res Commun
2004;322:203–9.
[14] Heo J, Campbell SL. Superoxide anion radical modulates the activity
of Ras and Ras-related GTPases by a radical-based mechanism
similar to that of nitric oxide. J Biol Chem 2005;280:12438–45.
[15] Ckless K, Reynaert NL, Taatjes DJ, Lounsbury KM, van der
Vliet A, Janssen-Heininger Y. In situ detection and visualization
of S-nitrosylated proteins following chemical derivatization: identification of Ran GTPase as a target for S-nitrosylation. Nitric Oxide 2004;11:
216–27.
[16] Heo J, Campbell SL. Mechanism of redox-mediated guanine
nucleotide exchange on redox-active Rho GTPases. J Biol Chem
2005;280:31003–10.
[17] Heo J, Campbell SL. Mechanism of p21Ras S-nitrosylation and kinetics
of nitric oxide-mediated guanine nucleotide exchange. Biochemistry
2004;43:2314–22.
[18] Schrammel A, Gorren A, Schmidt K, Pfeiffer S, Mayer B. S-nitrosation
U
U
of glutathione by nitric oxide, peroxynitrite, and NO/O2 −. Free Radic
Biol Med 2003;34:1078–88.
[19] Jourd'heuil D, Jourd'heuil F, Feelisch M. Oxidation and nitrosation of
thiols at low micromolar exposure to nitric oxide. Evidence for a free
radical mechanism. J Biol Chem 2003;278:15720–6.
[20] Ford E, Hughes M, Wardman P. Kinetics of the reactions of nitrogen
dioxide with glutathione, cysteine, and uric acid at physiological pH.
Free Radic Biol Med 2002;32:1314–23.
[21] Williams JG, Pappu K, Campbell SL. Structural and biochemical studies of p21Ras S-nitrosylation and nitric oxide-mediated
guanine nucleotide exchange. Proc Natl Acad Sci U S A 2003;100:
6376–81.
[22] Arkin MR, Stemp ED, Pulver SC, Barton JK. Long-range oxidation
of guanine by Ru(III) in duplex DNA. Chem Biol 1997;4:389–400.
[23] Niles JC, Wishnok JS, Tannenbaum SR. A novel nitroimidazole
compound formed during the reaction of peroxynitrite with 2′,3′,5′tri-O-acetyl-guanosine. J Am Chem Soc 2001;123:12147–51.
[24] Wagenknecht HA, Rajski SR, Pascaly M, Stemp ED, Barton JK.
Direct observation of radical intermediates in protein-dependent
DNA charge transport. J Am Chem Soc 2001;123:4400–7.
[25] Gu F, Stillwell WG, Wishnok JS, Shallop AJ, Jones RA, Tannenbaum
SR. Peroxynitrite-induced reactions of synthetic oligo 2′-deoxynucleotides and DNA containing guanine: formation and stability
of a 5-guanidino-4-nitroimidazole lesion. Biochemistry 2002;41:
7508–18.
[26] Joffe A, Mock S, Yun BH, Kolbanovskiy A, Geacintov NE,
Shafirovich V. Oxidative generation of guanine radicals by carbonate
radicals and their reactions with nitrogen dioxide to form site specific
5-guanidino-4-nitroimidazole lesions in oligodeoxynucleotides.
Chem Res Toxicol 2003;16:966–73.
[27] Kobayashi K, Tagawa S. Direct observation of guanine radical cation
deprotonation in duplex DNA using pulse radiolysis. J Am Chem Soc
2003;125:10213–8.
[28] Heo J, Campbell SL. Ras regulation by reactive oxygen and nitrogen
species. Biochemistry 2006;45:2200–10.
[29] Heo J, Gao G, Campbell S. pH-dependent perturbation of Ras guanine–
nucleotide interactions and Ras–guanine–nucleotide exchange. Biochemistry 2004;43:10102–11.
[30] Moncada S, Bolanos JP. Nitric oxide, cell bioenergetics and
neurodegeneration. J Neurochem 2006;97:1676–89.
[31] Stuehr DJ, Kwon NS, Nathan CF, Griffith OW, Feldman PL, Wiseman
J. N omega-hydroxy-L-arginine is an intermediate in the biosynthesis of
nitric oxide from L-arginine. J Biol Chem 1991;266:6259–63.
[32] Marletta MA. Nitric oxide synthase: function and mechanism. Adv
Exp Med Biol 1993;338:281–4.
237
[33] Cohen J, Evans TJ, Spink J. Cytokine regulation of inducible nitric
oxide synthase in vascular smooth muscle cells. Prog Clin Biol Res
1998;397:169–77.
[34] Hecker M, Cattaruzza M, Wagner AH. Regulation of inducible nitric
oxide synthase gene expression in vascular smooth muscle cells. Gen
Pharmacol 1999;32:9–16.
[35] Lane P, Gross SS. Cell signaling by nitric oxide. Semin Nephrol
1999;19:215–29.
[36] Pfeilschifter J, Eberhardt W, Beck KF. Regulation of gene expression
by nitric oxide. Pflugers Arch 2001;442:479–86.
[37] Carraro S, Doherty J, Zaman K, Gainov I, Turner R, Vaughan J, et al.
S-nitrosothiols regulate cell-surface pH buffering by airway epithelial
cells during the human immune response to rhinovirus. Am J Physiol
Lung Cell Mol Physiol 2006;290:L827–32.
[38] Burwell LS, Nadtochiy SM, Tompkins AJ, Young S, Brookes PS.
Direct evidence for S-nitrosation of mitochondrial complex I.
Biochem J 2006;394:627–34.
[39] Jourd'heuil D, Jourd'heuil FL, Lowery AM, Hughes J, Grisham MB.
Detection of nitrosothiols and other nitroso species in vitro and in
cells. Methods Enzymol 2005;396:118–31.
[40] Schonhoff CM, Matsuoka M, Tummala H, Johnson MA, Estevez AG,
Wu R, et al. S-nitrosothiol depletion in amyotrophic lateral sclerosis.
Proc Natl Acad Sci U S A 2006;103:2404–9.
[41] Gaston B, Carver J, Doctor A, Palmer LA. S-nitrosylation signaling
in cell biology. Mol Interv 2003;3:253–63.
[42] Li J, Billiar TR, Talanian RV, Kim YM. Nitric oxide reversibly
inhibits seven members of the caspase family via S-nitrosylation.
Biochem Biophys Res Commun 1997;240:419–24.
[43] Mannick JB, Schonhoff C, Papeta N, Ghafourifar P, Szibor M, Fang
K, et al. S-nitrosylation of mitochondrial caspases. J Cell Biol
2001;154:1111–6.
[44] So HS, Park RK, Kim MS, Lee SR, Jung BH, Chung SY, et al. Nitric
oxide inhibits c-Jun N-terminal kinase 2 (JNK2) via S-nitrosylation.
Biochem Biophys Res Commun 1998;247:809–13.
[45] Park HS, Huh SH, Kim MS, Lee SH, Choi EJ. Nitric oxide negatively
regulates c-Jun N-terminal kinase/stress-activated protein kinase by
means of S-nitrosylation. Proc Natl Acad Sci U S A 2000;97:14382–7.
[46] Barrett DM, Black SM, Todor H, Schmidt-Ullrich RK, Dawson KS,
Mikkelsen RB. Inhibition of protein-tyrosine phosphatases by mild
oxidative stresses is dependent on S-nitrosylation. J Biol Chem
2005;280: 14453–61.
[47] Li S, Whorton AR. Regulation of protein tyrosine phosphatase 1B in
intact cells by S-nitrosothiols. Arch Biochem Biophys 2003;410:
269–79.
[48] Bauer PM, Buga GM, Fukuto JM, Pegg AE, Ignarro LJ. Nitric oxide
inhibits ornithine decarboxylase via S-nitrosylation of cysteine 360 in
the active site of the enzyme. J Biol Chem 2001;276:34458–64.
[49] McMahon TJ, Exton Stone A, Bonaventura J, Singel DJ, Solomon
Stamler J. Functional coupling of oxygen binding and vasoactivity in
S-nitrosohemoglobin. J Biol Chem 2000;275:16738–45.
[50] Jia L, Bonaventura C, Bonaventura J, Stamler JS. S-nitrosohaemoglobin: a dynamic activity of blood involved in vascular control.
Nature 1996;380:221–6.
[51] Rafikova O, Rafikov R, Nudler E. Catalysis of S-nitrosothiols
formation by serum albumin: the mechanism and implication in
vascular control. Proc Natl Acad Sci U S A 2002;99:5913–8.
[52] Dugas B, Debre P, Moncada S. Nitric oxide, a vital poison inside
the immune and inflammatory network. Res Immunol 1995;146:
664–70.
[53] Ignarro LJ. Nitric oxide as a unique signaling molecule in the vascular
system: a historical overview. J Physiol Pharmacol 2002;53:503–14.
[54] Ignarro LJ, Buga GM, Wei LH, Bauer PM, Wu G, del Soldato P. Role
of the arginine-NO pathway in the regulation of vascular smooth
muscle cell proliferation. Proc Natl Acad Sci U S A 2001;98:4202–8.
[55] Moncada S, Erusalimsky JD. Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat Rev Mol Cell Biol 2002;3:
214–20.
238
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
[56] Napoli C, Ignarro LJ. Nitric oxide and atherosclerosis. Nitric Oxide
2001;5:88–97.
[57] Waldman SA, Murad F. Biochemical mechanisms underlying
vascular smooth muscle relaxation: the guanylate cyclase–cyclic
GMP system. J Cardiovasc Pharmacol 1988;12(Suppl 5):S115–8.
[58] Xu W, Charles IG, Moncada S. Nitric oxide: orchestrating hypoxia
regulation through mitochondrial respiration and the endoplasmic
reticulum stress response. Cell Res 2005;15:63–5.
[59] Hummel SG, Fischer AJ, Martin SM, Schafer FQ, Buettner GR.
Nitric oxide as a cellular antioxidant: a little goes a long way. Free
Radic Biol Med 2006;40:501–6.
[60] Fernandes DC, Medinas DB, Alves MJ, Augusto O. Tempol diverts
peroxynitrite/carbon dioxide reactivity toward albumin and cells from
protein–tyrosine nitration to protein–cysteine nitrosation. Free Radic
Biol Med 2005;38:189–200.
[61] Koppenol WH, Moreno JJ, Pryor WA, Ischiropoulos H, Beckman JS.
Peroxynitrite, a cloaked oxidant formed by nitric oxide and
superoxide. Chem Res Toxicol 1992;5:834–42.
[62] Prutz WA, Monig H, Butler J, Land EJ. Reactions of nitrogen dioxide
in aqueous model systems: oxidation of tyrosine units in peptides and
proteins. Arch Biochem Biophys 1985;243:125–34.
[63] Huie RE. The reaction kinetics of NO2. Toxicology 1994;89:193–216.
[64] Herold S, Rock G. Mechanistic studies of S-nitrosothiol formation by
NO⁎/O2 and by NO⁎/methemoglobin. Arch Biochem Biophys
2005;436:386–96.
[65] Hao G, Derakhshan B, Shi L, Campagne F, Gross SS. SNOSID, a
proteomic method for identification of cysteine S-nitrosylation sites in
complex protein mixtures. Proc Natl Acad Sci U S A 2006;103:1012–7.
[66] Dawson TM, Sasaki M, Gonzalez-Zulueta M, Dawson VL. Regulation
of neuronal nitric oxide synthase and identification of novel nitric oxide
signaling pathways. Prog Brain Res 1998;118:3–11.
[67] Lander HM, Hajjar DP, Hempstead BL, Mirza UA, Chait BT,
Campbell S, et al. A molecular redox switch on p21(ras). Structural
basis for the nitric oxide-p21(ras) interaction. J Biol Chem 1997;272:
4323–6.
[68] Lander HM, Milbank AJ, Tauras JM, Hajjar DP, Hempstead BL,
Schwartz GD, et al. Redox regulation of cell signalling. Nature
1996;381: 380–1.
[69] Lander HM, Ogiste JS, Pearce SF, Levi R, Novogrodsky A. Nitric
oxide-stimulated guanine nucleotide exchange on p21ras. J Biol
Chem 1995;270:7017–20.
[70] Lander HM, Sehajpal PK, Novogrodsky A. Nitric oxide signaling: a
possible role for G proteins. Am Assoc Immunol 1993;151:7182–7.
[71] Yun HY, Gonzalez-Zulueta M, Dawson VL, Dawson TM. Nitric
oxide mediates N-methyl-D-aspartate receptor-induced activation of
p21ras. Proc Natl Acad Sci U S A 1998;95:5773–8.
[72] Wainwright MS, Brennan LA, Dizon ML, Black SM. p21ras activation
following hypoxia–ischemia in the newborn rat brain is dependent on
nitric oxide synthase activity but p21ras does not contribute to
neurologic injury. Brain Res Dev Brain Res 2003;146:79–85.
[73] Brennan LA, Wedgwood S, Bekker JM, Black SM. NO activates
p21ras and leads to the inhibition of endothelial NO synthase by
protein nitration. DNA Cell Biol 2003;22:317–28.
[74] Monteiro HP, Gruia-Gray J, Peranovich TM, de Oliveira LC, Stern A.
Nitric oxide stimulates tyrosine phosphorylation of focal adhesion
kinase, Src kinase, and mitogen-activated protein kinases in murine
fibroblasts. Free Radic Biol Med 2000;28:174–82.
[75] Deora AA, Win T, Vanhaesebroeck B, Lander HM. A redox-triggered
ras-effector interaction. Recruitment of phosphatidylinositol 3′kinase to Ras by redox stress. J Biol Chem 1998;273:29923–8.
[76] Deora AA, Hajjar DP, Lander HM. Recruitment and activation of
Raf-1 kinase by nitric oxide-activated Ras. Biochemistry 2000;39:
9901–8.
[77] Sheta EA, Trout H, Gildea JJ, Harding MA, Theodorescu D. Cell
density mediated pericellular hypoxia leads to induction of HIF1alpha via nitric oxide and Ras/MAP kinase mediated signaling
pathways. Oncogene 2001;20:7624–34.
[78] Kasuno K, Takabuchi S, Fukuda K, Kizaka-Kondoh S, Yodoi J,
Adachi T, et al. Nitric oxide induces hypoxia-inducible factor 1
activation that is dependent on MAPK and phosphatidylinositol 3kinase signaling. J Biol Chem 2004;279:2550–8.
[79] Oliveira CJ, Schindler F, Ventura AM, Morais MS, Arai RJ, Debbas
V, et al. Nitric oxide and cGMP activate the Ras–MAP kinase
pathway-stimulating protein tyrosine phosphorylation in rabbit aortic
endothelial cells. Free Radic Biol Med 2003;35:381–96.
[80] Mallis RJ, Buss JE, Thomas JA. Oxidative modification of H-ras:
S-thiolation and S-nitrosylation of reactive cysteines. Biochem J
2001;355: 145–53.
[81] Raines KW, Cao GL, Lee EK, Rosen GM, Shapiro P. Neuronal nitric
oxide synthase-induced S-nitrosylation of HRas inhibits calcium
ionophore-mediated extracellular-signal-regulated kinase activity.
Biochem J 2006;397:329–36.
[82] Yu SM, Hung LM, Lin CC. cGMP-elevating agents suppress
proliferation of vascular smooth muscle cells by inhibiting the
activation of epidermal growth factor signaling pathway. Circulation
1997;95:1269–77.
[83] Woo HA, Kang SW, Kim HK, Yang KS, Chae HZ, Rhee SG.
Reversible oxidation of the active site cysteine of peroxiredoxins to
cysteine sulfinic acid. Immunoblot detection with antibodies specific
for the hyperoxidized cysteine-containing sequence. J Biol Chem
2003;278:47361–4.
[84] Chang TS, Jeong W, Woo HA, Lee SM, Park S, Rhee SG.
Characterization of mammalian sulfiredoxin and its reactivation
of hyperoxidized peroxiredoxin through reduction of cysteine
sulfinic acid in the active site to cysteine. J Biol Chem 2004;279:
50994–1001.
[85] Guikema B, Lu Q, Jourd'heuil D. Chemical considerations and
biological selectivity of protein nitrosation: implications for NOmediated signal transduction. Antioxid Redox Signal 2005;7:593–606.
[86] Augusto O, Bonini MG, Trindade D. Spin trapping of glutathiyl and
protein radicals produced from nitric oxide-derived oxidants. Free
Radic Biol Med 2004;36:1224–32.
[87] Lander HM, Ogiste JS, Teng KK, Novogrodsky A. p21ras as a
common signaling target of reactive free radicals and cellular redox
stress. J Biol Chem 1995;270:21195–8.
[88] Gonzalez-Zulueta M, Feldman AB, Klesse LJ, Kalb RG, Dillman JF,
Parada LF, et al. Requirement for nitric oxide activation of p21(Ras)/
extracellular regulated kinase in neuronal ischemic preconditioning.
Proc Natl Acad Sci U S A 2000;97:436–41.
[89] Baker T, Booden M, Buss J. S-Nitrosocysteine increases palmitate
turnover on Ha-Ras in NIH 3T3 cells. J Biol Chem 2000;275:22037–47.
[90] Heo J, Holbrook GP. Regulation of 2-carboxy-D-arabinitol 1-phosphate
phosphatase: activation by glutathione and interaction with thiol
reagents. Biochem J 1999;338(Pt 2):409–16.
[91] Stamler JS, Jaraki O, Osborne J, Simon DI, Keaney J, Vita J, et al.
Nitric oxide circulates in mammalian plasma primarily as an S-nitroso
adduct of serum albumin. Proc Natl Acad Sci U S A 1992;89:7674–7.
[92] Campbell SL, Khosravi-Far R, Rossman KL, Clark GJ, Der CJ.
Increasing complexity of Ras signaling. Oncogene 1998;17:1395–413.
[93] Etienne-Manneville S, Hall A. Rho GTPases in cell biology. Nature
2002;420:629–35.
[94] Jaffe AB, Hall A. Rho GTPases: biochemistry and biology. Annu Rev
Cell Dev Biol 2005;21:247–69.
[95] Wennerberg K, Rossman KL, Der CJ. The Ras superfamily at a
glance. J Cell Sci 2005;118:843–6.
[96] Takeya R, Sumimoto H. Molecular mechanism for activation of
superoxide-producing NADPH oxidases. Mol Cells 2003;16:271–7.
[97] Bokoch GM, Quilliam LA, Bohl BP, Jesaitis AJ, Quinn MT.
Inhibition of Rap1A binding to cytochrome b558 of NADPH oxidase
by phosphorylation of Rap1A. Science 1991;254:1794–6.
[98] Olkkonen VM, Stenmark H. Role of Rab GTPases in membrane
traffic. Int Rev Cytol 1997;176:1–85.
[99] Stenmark H, Olkkonen VM. The Rab GTPase family. Genome Biol
2001;2:1–7 (reviews3007).
K.W. Raines et al. / Cardiovascular Research 75 (2007) 229–239
[100] Darchen F, Goud B. Multiple aspects of Rab protein action in the
secretory pathway: focus on Rab3 and Rab6. Biochimie 2000;82:
375–84.
[101] Raftopoulou M, Hall A. Cell migration: Rho GTPases lead the way.
Dev Biol 2004;265:23–32.
[102] Ihara K, Muraguchi S, Kato M, Shimizu T, Shirakawa M, Kuroda S,
et al. Crystal structure of human RhoA in a dominantly active form
complexed with a GTP analogue. J Biol Chem 1998;273:9656–66.
[103] Hirshberg M, Stockley RW, Dodson G, Webb MR. The crystal
structure of human rac1, a member of the rho-family complexed with
a GTP analogue. Nat Struct Biol 1997;4:147–52.
[104] Feltham JL, Dotsch V, Raza S, Manor D, Cerione RA, Sutcliffe MJ,
et al. Definition of the switch surface in the solution structure of
Cdc42Hs. Biochemistry 1997;36:8755–66.
[105] Cairns RA, Hill RP. Acute hypoxia enhances spontaneous lymph
node metastasis in an orthotopic murine model of human cervical
carcinoma. Cancer Res 2004;64:2054–61.
[106] Bokoch GM. Regulation of cell function by Rho family GTPases.
Immunol Res 2000;21:139–48.
[107] Bokoch GM, Knaus UG. NADPH oxidases: not just for leukocytes
anymore! Trends Biochem Sci 2003;28:502–8.
[108] Kaibuchi K, Kuroda S, Amano M. Regulation of the cytoskeleton and
cell adhesion by the Rho family GTPases in mammalian cells. Annu
Rev Biochem 1999;68:459–86.
[109] Heo J, Raines KW, Mocanu V, Campbell SL. Redox regulation of
RhoA. Biochemistry 2006;45:14481–9.
[110] Gerhard R, John H, Aktories K, Just I. Thiol-modifying phenylarsine
oxide inhibits guanine nucleotide binding of Rho but not of Rac
GTPases. Mol Pharmacol 2003;63:1349–55.
[111] Zuckerbraun BS, Stoyanovsky DA, Sengupta R, Shapiro RA,
Ozanich BA, Rao J, et al. Nitric oxide-induced inhibition of smooth
muscle cell proliferation involves S-nitrosation and inactivation of
RhoA. Am J Physiol Cell Physiol 2006.
[112] Clavreul N, Adachi T, Pimental DR, Ido Y, Schoneich C, Cohen RA.
S-glutathiolation by peroxynitrite of p21ras at cysteine-118 mediates
its direct activation and downstream signaling in endothelial cells.
FASEB J 2006;20:518–20.
[113] Carballal S, Radi R, Kirk MC, Barnes S, Freeman BA, Alvarez B.
Sulfenic acid formation in human serum albumin by hydrogen
peroxide and peroxynitrite. Biochemistry 2003;42:9906–14.
239
[114] Bonini MG, Augusto O. Carbon dioxide stimulates the production of
thiyl, sulfinyl, and disulfide radical anion from thiol oxidation by
peroxynitrite. J Biol Chem 2001;276:9749–54.
[115] Starke DW, Chock PB, Mieyal JJ. Glutathione-thiyl radical
scavenging and transferase properties of human glutaredoxin
(thioltransferase). Potential role in redox signal transduction. J Biol
Chem 2003;278:14607–13.
[116] Borisenko GG, Martin I, Zhao Q, Amoscato AA, Kagan VE.
Nitroxides scavenge myeloperoxidase-catalyzed thiyl radicals in
model systems and in cells. J Am Chem Soc 2004;126:9221–32.
[117] Kolberg M, Bleifuss G, Graslund A, Sjoberg BM, Lubitz W,
Lendzian F, et al. Protein thiyl radicals directly observed by EPR
spectroscopy. Arch Biochem Biophys 2002;403:141–4.
[118] Lopes de Menezes S, Augusto O. EPR detection of glutathionyl and
protein-tyrosyl radicals during the interaction of peroxynitrite with
macrophages (J774). J Biol Chem 2001;276:39879–84.
[119] Sayle RA, Milner-White EJ. RASMOL: biomolecular graphics for
all. Trends Biochem Sci 1995;20:374.
[120] Graetzel M, Henglein A, Taniguchi S. Pulsradiolytische Untersuchung
der NO-Oxydation und des Gleichgewichts N2O3 b−N NO + NO2 in
Waessriger Loesung. Ber Bunsenges Phys Chem 1970:74.
[121] Goldstein S, Czapski G. Mechanism of nitration of thiols and amines
by NO/O2 in aqueous solution — the nature of the nitrosating
intermediate. J Am Chem Soc 1996;118:3419–25.
[122] Lewis RS, Tannenbaum SR, Deen WM. Kinetics of N-nitrosation in
oxygenated nitric oxide solutions at physiological pH: role of nitrous
anhydride and effects of phosphate and chloride. J Am Chem Soc
1995;117:3933–9.
[123] Ford E, Hughes MN, Wardman P. Kinetics of the reactions of nitrogen
dioxide with glutathione, cysteine, and uric acid at physiological pH.
Free Radic Biol Med 2002;32:1314–23.
[124] Wardman P, von Sonntag C. Kinetic factors that control the fate of
thiyl radicals in cells. Methods Enzymol 1995;251:31–45.
[125] Huie RE, Padmaja S. The reaction of no with superoxide. Free Radic
Res Commun 1993;18:195–9.
[126] Kissner R, Nauser T, Bugnon P, Lye PG, Koppenol WH. Formation
and properties of peroxynitrite as studied by laser flash photolysis,
high-pressure stopped-flow technique, and pulse radiolysis. Chem
Res Toxicol 1997;10:1285–92.