ArticlePDF Available

Alternative Splicing of Rac1 Generates Rac1b, a Self-activating GTPase

Authors:
  • Charles River

Abstract and Figures

Rac1b was recently identified in malignant colorectal tumors as an alternative splice variant of Rac1 containing a 19-amino acid insertion next to the switch II region. The structures of Rac1b in the GDP- and the GppNHp-bound forms, determined at a resolution of 1.75 Å, reveal that the insertion induces an open switch I conformation and a highly mobile switch II. As a consequence, Rac1b has an accelerated GEF-independent GDP/GTP exchange and an impaired GTP hydrolysis, which is restored partially by GTPase-activating proteins. Interestingly, Rac1b is able to bind the GTPase-binding domain of PAK but not full-length PAK in a GTP-dependent manner, suggesting that the insertion does not completely abolish effector interaction. The presented study provides insights into the structural and biochemical mechanism of a self-activating GTPase.
Content may be subject to copyright.
Alternative Splicing of Rac1 Generates Rac1b, a
Self-activating GTPase*
Received for publication, September 16, 2003, and in revised form, October 24, 2003
Published, JBC Papers in Press, November 18, 2003, DOI 10.1074/jbc.M310281200
Dennis Fiegen‡, Lars-Christian Haeusler‡, Lars Blumenstein, Ulrike Herbrand,
Radovan Dvorsky, Ingrid R. Vetter, and Mohammad R. Ahmadian§
From the Max-Planck-Institut fu¨ r molekulare Physiologie, Abteilung Strukturelle Biologie, Otto-Hahn-Strasse 11,
44227 Dortmund, Germany
Rac1b was recently identified in malignant colorectal
tumors as an alternative splice variant of Rac1 contain-
ing a 19-amino acid insertion next to the switch II
region. The structures of Rac1b in the GDP- and the
GppNHp-bound forms, determined at a resolution of
1.75 Å, reveal that the insertion induces an open switch
I conformation and a highly mobile switch II. As a con-
sequence, Rac1b has an accelerated GEF-independent
GDP/GTP exchange and an impaired GTP hydrolysis,
which is restored partially by GTPase-activating pro-
teins. Interestingly, Rac1b is able to bind the GTPase-
binding domain of PAK but not full-length PAK in a
GTP-dependent manner, suggesting that the insertion
does not completely abolish effector interaction. The
presented study provides insights into the structural
and biochemical mechanism of a self-activating GTPase.
The small GTPase Rac acts as a binary molecular switch that
cycles between an inactive GDP-bound state and an active
GTP-bound state in response to a variety of extracellular stim-
uli. The interconversion between both states is controlled by
nucleotide exchange and GTP hydrolysis. The structures of
several GTPases in either state revealed that the switching
mechanism depends on the conformational change of two re-
gions, termed switch I and switch II (1, 2). The switch regions
consequently provide a surface that is in the GTP-bound state
specifically recognized by downstream effectors (1–3) and
GTPase-activating proteins (GAPs),
1
accelerating the slow in-
trinsic GTP hydrolysis reaction (1, 46). After GTP hydrolysis,
release of the cleaved
-phosphate allows the switch regions to
relax into the GDP conformation. Guanine nucleotide exchange
factors (GEFs), stimulating the GDP/GTP exchange, bind in-
dependently of the nucleotide-bound state (1, 7), whereas gua-
nine nucleotide dissociation inhibitors (GDIs), which sequester
the GTPase from the membrane into the cytoplasm, interact
only with the GDP-bound state (8).
Rac1b was discovered in human tumors as an alternative
splice variant of Rac1 containing a 19-amino acid insertion
(between codons 75 and 76) at the end of the switch II region (9,
10). It has been suggested that the insertion may create a novel
effector-binding site in Rac1b and thus participate in signaling
pathways related to the neoplastic growth of the intestinal
mucosa (9). Most recently, it has been shown that Rac1b does
not interact with Rho-GDI and PAK1 and is not involved in
lamellipodia formation but able to activate the transcription
factor NF-
B (11). We tried to address what influence this
insertion might have on the structure and the biochemical
properties of Rac1b in comparison with Rac1. Therefore,
we solved the crystal structures of Rac1b in the GDP- and
GppNHp-bound conformations at 1.75 Å resolution. Further-
more we investigated nucleotide binding and hydrolysis of
Rac1b and studied its regulation by the RacGEF Tiam1 and
p50
GAP
and its interaction with the downstream effector PAK.
The Rac1b structures explain the drastic changes of the bio-
chemical properties of Rac1b, namely a dramatic decrease in
nucleotide affinity and GTP hydrolysis. The presented data
identify Rac1b as a predominantly GTP-bound form of Rac1.
EXPERIMENTAL PROCEDURES
Plasmids—The pcDNA3-FLAG constructs of human Rac1b, Rac1,
and the respective constitutive active Rac1(G12V) mutant were gener-
ated by PCR and cloned via BamHI and EcoRI restriction sites. Rac1,
Rac1C (1–184), Rac1b, and Rac1bC (1–201) were cloned in
pGEX4T1, using BamHI and EcoRI restriction sites. The DH-PH do-
main of Tiam1 (1033–1404) was cloned into pGEX4T1 using BamHI
and XhoI restriction sites. The coding region of Tiam1 contains an
internal BamHI site that was removed for the cloning procedure. pGEX-
PAK1-GBD was kindly provided by J. Collard (12). Full-length pGEX-
PAK was kindly provided by A. Wittinghofer (13).
Preparation of Recombinant Proteins—Rac1, Rac1C, Rac1b, and
Rac1bC, the catalytic domain of p50
GAP
(amino acids 198–439), the
Cdc42/Rac-interacting binding domain of PAK (amino acids 57–141),
full-length PAK, and the DH-PH domain of Tiam1 were produced as
glutathione S-transferase (GST) fusion proteins in Escherichia coli. All
of the proteins were purified as described previously for Rnd3 (14).
Nucleotide-free GTPases as well as fluorescent nucleotide-bound
GTPases were prepared, and concentration and quality were deter-
mined as described (15).
Crystallization and Data Collection—Crystals of truncated Rac1bC
(1–184) in complex with GDP and GppNHp (nonhydrolyzable GTP
analog) were grown at 20 °C using the hanging drop method by mixing
2
lofa0.5mMsolution of the Rac1b G domain in 20 mMTris/HCl, pH
7.5, 2 mMMgCl
2
,2mMdithioerythritol, 100
MGDP or GppNHp with
2
l of reservoir solution consisting of 100 mMHepes buffer, pH 7.5,
18–30% polyethylene glycol 3350, and 2–6% isopropanol. The crystals
of both complexes belonged to space group P2
1
2
1
2
1
(a51.55 Å, b
78.67 Å, c96.88 Å). For data collection at 100 K, the crystals were
* This work was supported in part by a European Community Marie
Curie Fellowship (to R. D.), by funds from the Verband der Chemischen
Industrie and the Bundesministerium fu¨ r Bildung und Forschung (to
D. F.), and by funds from the Deutsche Forschungsgemeinschaft (to
L.-C. H. and L. B.). The costs of publication of this article were defrayed
in part by the payment of page charges. This article must therefore be
hereby marked “advertisement” in accordance with 18 U.S.C. Section
1734 solely to indicate this fact.
The atomic coordinates and structure factors (codes 1RYF and 1RYH)
have been deposited in the Protein Data Bank, Research Collaboratory
for Structural Bioinformatics, Rutgers University, New Brunswick, NJ
(http://www.rcsb.org/).
‡ These authors contributed equally to this work.
§ To whom correspondence should be addressed. E-mail: reza.
ahmadian@mpi-dortmund.mpg.de.
1
The abbreviations used are: GAP, GTPase-activating protein; GEF,
guanine nucleotide exchange factor; GDI, guanine nucleotide dissocia-
tion inhibitor; DH-PH, double homology-pleckstrin homology; GST, glu-
tathione S-transferase; HPLC, high pressure liquid chromatography;
GppNHp, quanosine 5-
,
-imidotriphosphate; GBD, GTPase-binding
domain.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 279, No. 6, Issue of February 6, pp. 4743–4749, 2004
© 2004 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org 4743
by guest on November 18, 2016http://www.jbc.org/Downloaded from
transferred to a solution containing reservoir solution and 10% glycerol.
A cryo-protected crystal was then suspended in a rayon loop (Hampton
Research) and flash frozen in liquid nitrogen. X-ray diffraction data
were collected on an ADSC Q4 CCD detector at the beam line ID141
at the European Synchrotron Radiation Facility and were processed
using XDS (16). Details on data collection, structure determination, and
crystallographic refinement are summarized in Table I.
Structure Determination and Crystallographic RefinementThe ini-
tial phases were calculated by molecular replacement, using the pro-
gram AMoRe (17) and a Rac1 search model based on the Rac1GppNHp
structure lacking the bound nucleotide and the switch regions (18)
(Protein Data Bank code 1mh1). After initial rigid body refinement, 20
cycles of simulated annealing and model building were performed. For
the last refinement steps REFMAC5 was used (19). For the detection of
crystallographic water molecules, ARP/wARP (20) and REFMAC5 were
used. The residue ranges that were included in the final model as well
as the corresponding R-factors are listed in Table I. For all four mole-
cules the two additional N-terminal Gly-Ser residues caused by the
thrombin cleavage site could be observed and were included in the
model.
Fluorescence MeasurementsLong time fluorescence measurements
were monitored on a LS50B PerkinElmer Life Sciences spectrofluorom-
eter, and rapid kinetics were measured with a stopped flow apparatus
(Applied Photophysics SX16MV) as described (15). Nucleotide associa-
tion was performed with 0.1
Mfluorescent nucleotide and varying
concentrations of nucleotide-free Rac proteins at 20 °C as described
(15). The dissociation of the fluorescent nucleotide from Rac proteins
(0.1
M) was measured by the addition of 200-fold excess of nonfluores-
cent nucleotide in the absence and the presence of 5
MTiam1 DH-PH
at 20 °C. The equilibrium dissociation constant (K
d
) for the PAK-GBD
interaction with Rac1b was determined as previously described for the
Ras-Raf kinase interaction (21). The measurements were carried out
using 0.2
MmantGppNHp-bound GTPase, 40
MGppNHp, and in-
creasing concentrations of PAK-GBD at 25 °C for Rac1 and at 10 °Cin
the case of Rac1b because of its fast nucleotide dissociation rate. All of
the measurements were carried out in 30 mMTris/HCl, pH 7.5, 5 mM
MgCl
2
,10mMNa
2
HPO
4
/NaH
2
PO
4
pH 7.5, 5 mMdithioerythritol. The
observed rate constants were evaluated using Grafit (Erithacus
software).
GTPase AssayThe intrinsic and GAP-stimulated GTP hydrolysis
reactions were measured by HPLC on a C
18
reversed phase column
using a mixture of 80
Mnucleotide-free GTPase and 70
MGTP in the
presence and the absence of 8
MGAP at 25 °Cin30mMTris/HCl, pH
7.5, 5 mMdithioerythritol, 10 mMNa
2
HPO
4
/NaH
2
PO
4
,5mMMgCl
2
as
described (15). The relative GTP content was calculated by the ratio of
[GTP]/([GTP][GDP]). For exponential fitting of the data, the program
Grafit (Erithacus software) was used.
Transfection and GTPase Pull-down AssayCOS-7 cells were trans-
fected using DEAE-dextran as described (12). Pull-down assay for the
active GTP-bound Rac proteins was carried out using GST-PAK-GBD
(glutathione S-transferase-fused Rac-binding domain of PAK) conju-
gated with glutathione beads as described (12). The interaction of
full-length GST-PAK with the Rac proteins was examined under the
same conditions using purified proteins. The beads were washed four
times and subjected to SDS-PAGE (15% polyacrylamide). Bound Rac
proteins were detected by Western blot using a monoclonal antibody
against Rac (Upstate Biotechnologies, Inc.).
RESULTS AND DISCUSSION
Rapid GEF-independent Nucleotide Dissociation Reaction of
Rac1bTo investigate the influence of the 19-amino acid in-
sertion on the nucleotide binding affinity, we first determined
kinetic constants for the association of fluorescently (methyl-
anthraniloyl- or mant-) labeled nucleotides to nucleotide-free
Rac1b protein. This allowed us to monitor nucleotide associa-
tion kinetics at increasing protein concentrations. As shown in
Fig. 1A, the formation of the binary complex is not affected by
the 19-amino acid insertion. The association rate constants
(k
on
) for the binding of mantGDP and mantGTP to Rac1b were
obtained by linear fitting of the observed rate constants at the
given protein concentrations. They are only marginally slower
than those of Rac1 (Table II).
To determine the intrinsic and GEF-accelerated nucleotide
dissociation rates, Rac1b and Rac1 were loaded with the re-
spective fluorescently labeled guanine nucleotides. The dis-
placement of the fluorescent nucleotides was initiated by the
addition of excess amounts of nonfluorescent nucleotides in the
presence and absence of the DH-PH domain of Tiam1 (a Rac-
specific GEF). Drastic increases in the intrinsic dissociation of
mantGDP (26-fold), mantGTP (27-fold), and mantGppNHp
(250-fold) from Rac1b compared with the very slow dissociation
rates of the respective nucleotides from Rac1 were observed
(Fig. 1Band Table II). Accordingly, the calculated K
d
for nu-
cleotide binding revealed that the 19-residue insertion dramat-
TABLE I
Data collection and refinement statistics of Rac1b
GDP GppNHp
Intensity data processing
Resolution (Å) 24.921.75 Å26.261.75 Å
Number of reflections 130,220 138,306
Number of independent reflections 38,560 38,875
R
sym
(%)
a
5.4 (40.9)
b
5.1 (40.6)
b
Completeness of data (%) 95.1 (88.3)
b
96.0 (90.1)
b
Mean I/
(I)15.21 (3.37)
b
16.62 (3.75)
b
Molecular replacement statistics
Resolution range rotation translation 10.04.0 A/8.04.0 Å10.04.0 A/8.04.0 Å
Rotation (°)
c
152.02, 70.81, 163.02 (151.49, 70.92, 162.12)
d
152.21, 70.87, 163.31 (151.46, 70.96, 161.47)
d
Translation (Å)
c
4.99, 7.34, 8.63 (1.80, 32.33, 9.06)
d
4.54, 7.50, 8.47 (1.89, 32.30, 9.40)
d
Correlation coefficient 61.5 (22.6)
e
62.6 (25.3)
e
R
cryst
(%)
f
47.9 (64.1)
e
45.3 (60.7)
e
Refinement Statistics
R
cryst
(%)
f
18.8 18.1
R
free
(%)
f
22.3 21.9
Root mean square bond lengths (Å) 0.021 0.021
Root mean square bond angles (°) 1.87 1.88
Total number of residues 614 (273)
g
638 (296)
g
Residue ranges 159 and 93201 (158 and 93199)
h
160 and 93201 (158 and 93199)
h
a
R
sym
100I⫺具I/I.
b
Brackets are quantities calculated in the highest resolution bin at 1.851.75 Å.
c
Eulerian angles (
,
,
) are as defined in AMoRe, and translation is given in the orthogonal system.
d
Brackets show the solution for the second molecule in the asymmetric unit.
e
Brackets are quantities of the second best solution.
f
R
cryst
100F
o
F
c
/F
o
.R
free
is R
cryst
that was calculated using 5% of the data, chosen randomly, and omitted from the subsequent
structure refinement.
g
Brackets show the number of included water molecules.
h
Brackets show the residue range for the second molecule in the asymmetric unit.
Consequences of Alternative Rac1 Splicing4744
by guest on November 18, 2016http://www.jbc.org/Downloaded from
ically affects the overall affinity for GDP, GTP, and particularly
the GTP analog GppNHp (Table II). The reason for the reduced
affinity of GppNHp compared with that of GTP is the disrupted
hydrogen bond of the GTP-
,
-bridging imino group to the
main chain NH group of the P-loop residue Ala
13
. A similar
observation has been reported for RasmantGppNHp (22).
Moreover, in contrast to the slow nucleotide dissociation of
Rac1, which could be 50-fold accelerated in the presence of the
DH-PH domain of Tiam1, the fast intrinsic mantGDP dissoci-
ation of Rac1b could not be further increased by the DH-PH
domain of Tiam1 (Fig. 1Cand Table II).
These in vitro results show that Rac1b does not require any
GEF to get activated. It rather activates itself by a very fast
nucleotide dissociation and by the subsequent binding of the
cellular abundant GTP. Despite the drastically increased nu-
cleotide dissociation, Rac1b exhibits a nucleotide binding affin-
ity in the low nanomolar range, which is still high enough to act
as a GTP-binding protein in cells. However, because the
DH-PH domain of Tiam1 is a weak exchange factor in vitro (51)
and displays a 10 times higher activity on prenylated Rac1
bound to liposomes than on soluble unprenylated Rac1 (23), we
cannot exclude the possibility that Rac1b can in principle in-
teract with Tiam1 under cellular conditions as shown with
constitutive active Tiam1, overexpressed in NIH3T3 cells (11).
Impaired Intrinsic GTP Hydrolysis Reaction of Rac1bA
second crucial function of small GTPases is their slow intrin-
sic GTP hydrolysis reaction, which needs to be stimulated by
GAPs to switch off downstream signaling. Therefore, we
measured the GTP hydrolysis reaction of Rac1b in direct
comparison with that of Rac1 using a HPLC-based technique.
Interestingly, we found that the intrinsic GTP hydrolysis
reaction of Rac1b (0.0035 min
1
) was about 30-fold reduced
compared with that of Rac1 (0.11 min
1
) (Fig. 1Dand Table
II). In contrast to our results it has been previously published
that Rac1 and Rac1b show the same GTPase activity (10).
This discrepancy can be explained by the method this group
employed. The filter binding assay seems to be inappropriate
for a protein with a fast nucleotide dissociation such as
Rac1b. Unlike the constitutive active mutants of Rac1 (G12V
in the P-loop and Q61L in the switch II region) that also have
an impaired intrinsic GTP hydrolysis (24), the defective
GTPase reaction of Rac1b can be restored by GAP proteins.
As shown in Fig. 1D, the catalytic domain of p50
GAP
stimu-
lated the GTPase reaction of Rac1b up to 55-fold (21-fold
for Rac1), showing that GAP is able to stabilize the catalytic
elements of Rac1b and thus accelerate the GTPase
reaction.
High Level of Rac1bGTP in COS-7 CellsConsidering the
FIG.1.Biochemical properties of Rac1b. A, nucleotide association. The binding of increasing concentrations of nucleotide-free Rac1 (filled
symbols) and Rac1b (open symbols) to mantGDP (circles), mantGTP (squares), and mantGppNHp (triangles) was measured, and the dependence
of the observed rate constant on the protein concentration was linearly fitted to determine association constants (k
on
). B, nucleotide dissociation.
The release of mantGDP (open circles), mantGTP (open squares) and mantGppNHp (open triangles) from Rac1b and mantGDP from Rac1 (filled
circles) was measured after addition of excess of the respective nonfluorescent nucleotides and exponentially fitted to determine dissociation
constants (k
off
). C, Tiam1-catalyzed nucleotide dissociation. Dissociation of bound mantGDP from 0.1
MRac1 (circles) and Rac1b (squares) was
monitored in the presence (filled symbols) and absence (open symbols)of5
MTiam1 DH/PH. D, GTP hydrolysis reaction. HPLC measurements
of the GTP hydrolysis of Rac1 (filled symbols) and Rac1b (open symbols) were performed in the presence (squares) and absence (circles)of
p50RhoGAP.
Consequences of Alternative Rac1 Splicing 4745
by guest on November 18, 2016http://www.jbc.org/Downloaded from
increased nucleotide dissociation and the decreased GTP hy-
drolysis, it was tempting to assume that Rac1b is GTP-bound in
cells. To prove this assumption, Rac1, its constitutive active
mutant Rac1(G12V) and Rac1b were overexpressed in COS-7
cells under serum-starved conditions for 48 h. The fact that
wild-type Rac1b could be pulled down with GST-PAK-GBD
verifies our hypothesis that Rac1b exists in an active confor-
mation in serum-starved cells. Thereby it resembles the con-
stitutive active Rac1(G12V) mutant (Fig. 2A). As expected wild-
type Rac1 could not be detected under these conditions and
obviously needs GEF proteins to be activated. A GTP-depend-
ent Rac1b-PAK interaction was demonstrated by performing
the pull-down assay with purified GDP- and GppNHp-bound
Rac1b. As shown in Fig. 2B, GST-PAK-GBD selectively binds
Rac1bGppNHp but not Rac1bGDP, similar to the Rac1 control
experiment. These results reveal that Rac1b is, independent of
external stimuli, GTP-bound in cells and can selectively inter-
act with Rac effector proteins.
Rac1 involvement in transcription and growth control and its
requirement for Ras-induced malignant transformation is
widely known (2529). This knowledge is based on experiments
using the expression of a constitutively active Rac1(G12V) mu-
tant. A fast cycling mutant of Cdc42, Cdc42(F28L), has been
shown to undergo spontaneous nucleotide exchange in the ab-
sence of a GEF while maintaining full GTPase activity (30).
This mutant has an even greater cell-transforming potential in
fibroblasts compared with the constitutively active Cdc42(G12V)
mutant. However, our biochemical data clearly shows that
Rac1b behaves as a self-activating GTPase that is predomi-
nantly GTP-bound in cells.
Low Affinity Binding of Rac1b to PAK-GBDTo characterize
the effect of the insertion on effector interaction, we determined
equilibrium dissociation constants (K
d
) of PAK-GBD binding to
GppNHp-bound Rac1b and Rac1 using the GDI assay (31). As
shown in Fig. 3, increasing concentrations of PAK-GBD re-
sulted in incremental inhibition of the mantGppNHp dissocia-
tion from Rac1b and Rac1. We obtained a K
d
value of 0.49
M
for the PAK-GBD interaction with Rac1, which nicely corre-
sponds to previous reports (32). For Rac1b we observed a 7-fold
reduced binding affinity of PAK-GBD. The lower K
d
of 3.55
M
can be most likely attributed to the 19-amino acid insertion.
MantGDP dissociation from Rac1b was not inhibited under
these conditions (data not shown), confirming the GTP-depend-
ent interaction of Rac1b with PAK-GBD.
Furthermore, we examined the interaction of Rac1 and
Rac1b with full-length PAK using a GST pull-down assay. Fig.
2Bshows that we could not detect binding of Rac1b to full-
length PAK, and hence Rac1b stands in contrast to Rac1.
Compared with the high affinity binding of the isolated PAK-
GBD domains to Cdc42, it has been recently shown that full-
length PAK has a much lower binding affinity for Cdc42 (13,
32, 33). Assuming that this is also true for Rac1, our biochem-
ical data suggest an extremely low affinity of full-length PAK
for Rac1b.
Conserved Overall Structure of Rac1bGDP and Rac1b
GppNHpTo gain insight into the structural impact of the
19-amino acid insertion of Rac1b, we determined the crystal
structures of Rac1b in the GDP- and GppNHp-bound states.
The crystals diffracted to 1.75 Åresolution (Table I). Size
exclusion chromatography showed that Rac1b is monomeric in
solution (data not shown), but it crystallized as a dimer with
two molecules/asymmetric unit in a head to head fashion. The
contact surface has a size of 1347 Å
2
and is build up by
1to
3,
1, and the switch I of both molecules. Both molecules in the
asymmetric unit are very similar as can be derived from the
low root mean square deviation of 0.45 Åfor 165 common C
atoms.
The ribbon representation in Fig. 4 shows the secondary
structure elements of Rac1bGDP and Rac1bGppNHp super-
imposed on the GppNHp-bound Rac1 (18). The overall struc-
tures of both nucleotide-bound forms of Rac1b are remarkably
similar to each other and conserved as compared with the
structures of Rac1GppNHp (18), RhoAGTP
S (34), and
Cdc42GDP (35) with root mean square deviations of 0.67
(156 common C
atoms), 0.80 (155 common C
atoms), and
TABLE II
Biochemical properties of Rac1 and Rac1b
The dissociation constant (K
d
) for the nucleotide binding values has
been calculated from the association and the dissociation rate constants
of the respective nucleotides (K
d
k
off
/k
on
).
Rac1 Rac1b
mantGDP
k
on
(s
1
M
1
)2.47 10
6
1.12 10
6
k
off
(s
1
)7.0 10
5
1.8 10
3
Tiam1-catalyzed k
off
(s
1
)
a
3.6 10
3
1.9 10
3
K
d
(nM)0.028 1.6
mantGppNHp
k
on
(s
1
M
1
)1.05 10
6
k
off
(s
1
)1.1 10
4
2.9 10
2
K
d
(nM)27.6
mantGTP
k
on
(s
1
M
1
)1.49 10
6
1.05 10
6
k
off
(s
1
)1.0 10
4
2.7 10
3
K
d
(nM)0.067 2.6
GTP hydrolysis
Intrinsic (min
1
)0.11 0.0035
GAP-stimulated (min
1
)
b
2.36 0.194
PAK binding
K
d
(
M)0.49 3.55
a
5
MTiam1 was applied in the reaction of 0.1
MGTPase.
b
The GAP concentration (8
M) was 10-fold below the GTPase con-
centration (80
M). FIG.2.GTP-dependent binding to PAK. A, high Rac1bGTP level
in serum-starved COS-7 cells. FLAG-tagged wild-type (wt) Rac1, con-
stitutive active Rac1(G12V), and wild-type Rac1b were pulled down
with GST-PAK-GBD and the respective Rac-proteins were detected by
Western blot. The upper panel shows the expression control (total Rac),
and the lower panel shows the GTP-bound active Rac. The arrows
indicate the endogenous (open) and transfected (filled) Rac proteins. B,
Rac1b binds PAK-GBD but not full-length PAK. GDP- and GppNHp-
bound Rac1 and Rac1b (C-terminal truncated) were employed in pull-
down (PD) assay using GST-PAK-GBD and full-length GST-PAK, re-
spectively. PAK-GBD selectively binds the GTPase in a GTP-dependent
manner, whereas full-length GST-PAK only binds GppNHp-bound
Rac1 but not Rac1b.
Consequences of Alternative Rac1 Splicing4746
by guest on November 18, 2016http://www.jbc.org/Downloaded from
0.81 (151 common C
atoms), respectively, except for the loop
regions as described below.
Nucleotide binding requires five conserved sequence ele-
ments (36), of which three are well ordered in the Rac1bGDP
and Rac1bGppNHp structures. The P-loop (
13
AVGKT motif)
plays a crucial role in phosphate binding and strongly con-
tributes to magnesium (Mg
2
) coordination by the side chain
of the Thr
17
residue. The
115
TKLD and
157
CSAL motifs (Rac
numbering), on the other hand, directly interact with the
guanine base and thereby ensure specificity of guanine
nucleotide binding.
The most critical motifs, the central region of switch I
(
32
YIPT) and the N-terminal part of switch II (
57
DTAGQ),
that govern the Mg
2
and
-phosphate binding display the
major structural difference between Rac1b and Rac1. The
switch I region (amino acids 3038) is drastically displaced
from the nucleotide-binding site in the GDP- and GppNHp-
bound Rac1b structures compared with the Rac1 structure
(see below). We did not observe any electron density for the
switch II region and the proximate 19-amino acid insertion
(amino acids 7694), indicating that this insert is highly
mobile and also leads to a higher mobility of the switch II
region (see below).
Highly Mobile Switch II and the 19-Amino Acid Insertion
The switch II region and the adjacent 19-amino acid insertion
(amino acids 5992 in the GDP-bound and 6193 in the Gp-
pNHp-bound structures) are not resolved and therefore not
included in the crystal structures (Fig. 4). In most structures of
GppNHp-bound GTPases, the switch II region is well ordered
(37), indicating that the insertion in Rac1b contributes to a
higher mobility of the switch II region and thus leads to an
impaired GTPase reaction of Rac1b (Fig. 1D). The catalytic
residue Gln61 in switch II (
57
DTAGQ motif) is crucial in this
context, and its high flexibility is most likely the reason for the
impaired GTPase reaction of Rac1b (Fig. 1D). This is due to the
missing stabilization of the nucleophilic water leading to a
decreased GTP hydrolysis rate of Rac1b. It has been shown
before that the mutation of this key residue (Gln
61
) signifi-
cantly affects the intrinsic and the GAP-stimulated GTP hy-
drolysis reactions (5, 24, 38, 39). Interestingly, the fact that
GAP is able to stimulate the GTP hydrolysis of Rac1b strongly
indicates that the switch regions, naturally providing the GAP-
binding site, can be stabilized in a GTPase-competent confor-
mation. It has been proposed previously that the 19-amino acid
insertion could form a new functional domain built of two
-strands connected by a turn (9), but our results make it
unlikely that this region, because of its flexibility, has any
secondary structure.
Displacement of the Switch I Region in Rac1bThe most
obvious structural difference of Rac1b in comparison with Rac1
is observed in the region between Ala
28
and Asn
39
encompass-
ing the switch I of Rac1b. The drastic displacement of the
switch I region from the nucleotide-binding site is similar in
the Rac1bGDP and Rac1bGppNHp structures (Fig. 5, Aand
B). Switch I exhibits a maximum distance of 6.5 Åto the
nucleotide-binding site compared with 3.2 Åin Rac1, where it
completely covers the nucleotide (Fig. 5C).
The reason for this open switch I conformation is most likely
the disruption of the interaction between switches I and II. In
the Rac1GppNHp structure Phe
37
of switch I lies in a hydro-
phobic cleft that is composed by the side chains of Thr
58
, Tyr
64
,
Leu
67
, and Arg
68
of switch II. Additionally, Val
36
makes a
hydrophobic interaction with Tyr
64
. Because of the high mobil-
ity of the switch II in Rac1b, this interaction cannot take place,
resulting in a destabilization of the switch I region. In the
crystal, the open conformation of switch I is stabilized by a
hydrophobic interaction of the Phe
37
side chains of the adjacent
molecules. Although switch I is stabilized by the crystal pack-
ing, it still exhibits an increased B-factor of about 45 compared
with 30 for residues in the core of the protein. This suggests
FIG.3.Quantitative measurement of the PAK-GBD interaction
with Rac1 (A) and Rac1b (B). Dissociation of mantGppNHp from
Rac1b was inhibited by increasing concentration of the PAK-GBD (250
M). The observed rate constants were plotted against the concentra-
tion of the PAK-GBD to obtain equilibrium dissociation constants of
0.49
Mfor Rac1 and 3.55
Mfor Rac1b.
FIG.4.Comparison of the Rac1b structures with Rac1. Ribbon
representation of Rac1bGDP (purple) and Rac1bGppNHp (red) were
superimposed on the structure of Rac1GppNHp (brown) (17). The
switch II region of Rac1 is highlighted in orange. The nucleotide (Gp-
pNHp of Rac1b) and the Mg
2
ion are shown as ball-and-stick.
Consequences of Alternative Rac1 Splicing 4747
by guest on November 18, 2016http://www.jbc.org/Downloaded from
that switch I is probably very flexible in solution.
High affinity nucleotide binding and GTPase activity of
small GTPases are crucially dependent on the presence of an
Mg
2
ion. The octahedral Mg
2
coordination is highly con-
served throughout small GTPases (4042) but shows signifi-
cant differences in Rac1b. In the GDP-bound state, the magne-
sium is directly coordinated by an oxygen of the
-phosphate,
the side chain of T17 (P-loop) and three water molecules (Wat2,
116 and 117), but lacks the coordination of Thr
35
. This key
interaction is replaced by a fourth water molecule (Wat1; Fig.
5A). A similar arrangement is observed for Rac1bGppNHp,
except for the replacement of one water molecule (Wat2) by the
-phosphate oxygen of GppNHp (Fig. 5B).
As a consequence of the open switch I, contacts of the invari-
ant Thr
35
with the
-phosphate (main chain NH-group) and the
Mg
2
ion (side chain OH group) are disrupted and provide an
explanation for the rapid nucleotide dissociation of Rac1b. This
situation is comparable with the T35A mutation in Ras, which
drastically reduces the nucleotide affinity, because of the loss
in Mg
2
coordination (41, 43) and in particular to the mecha-
nism of Tiam1-catalyzed nucleotide exchange (44). The crystal
structure of the nucleotide-free Rac1Tiam1 complex has shown
that the interaction of the DH domain of Tiam1 with Rac1 has
shifted switch I and
2 (amino acids 2745) up to 2.7 Åalong
the nucleotide-binding cleft (44). Thereby, Thr
35
in Rac1 is
displaced, and the Thr
35
-Mg
2
interaction is disrupted (Fig.
5C). Thus, Tiam1 binding to Rac1 prevents Thr
35
from binding
to the Mg
2
ion and allows GDP release from Rac1 (1, 44).
Because the P-loop contacts with either GDP or GppNHp are
well conserved and the increased dissociation rates of GDP and
GTP are rather similar, we propose that the 19-amino acid
insertion in Rac1b induces similar effects on the switch I region
as Tiam1.
Interaction with Regulators and EffectorsFor comparative
structural analyses of Rac1b interaction with regulators and
effectors, we used the following structures: the nucleotide-free
Rac1Tiam1 complex (44); RhoGDI in complex with Rac1, Rac2,
and Cdc42 (4547); the transition state complex of RhoAGAP
(48); Rac1GppNHpPAK (49); and Rac1(Q61L)GTPp67
PHOX
(50). In addition to
2/
3/
4 (Tiam1),
4/
3 (RhoGDI), or
1/
5
(PAK-GBD), these proteins basically interact with the switch
regions of the GTPase, except for p67
PHOX
, which contacts
1,
2, and
5.
A superposition of GDI on Rac1b revealed that GDI binding
to switch II and the
3 helix of Rac1b could accommodate the
insertion. However, it has been recently shown that GDI does
not bind Rac1b (11). The proper contact of GDI to the
3 helix
of Rac1 has been shown to be essential because a H103A
mutation abolishes its interaction with Rho-GDI (51). Because
the structure of the
3 helix is not changed comparing Rac1b to
Rac1, we suggest that the insertion may sterically interfere
with GDI binding.
The association of Tiam1 with Rac1b has recently been
shown in immunoprecipitation experiments (11), indicating
that the structural requirements for the Rac1b-Tiam1 interac-
tion are not affected by the 19-amino acid insertion. However,
in contrast to the Rac1b-GAP interaction, which resulted in
stimulation of the GTP hydrolysis reaction of Rac1b (Fig. 1D),
no further increase in the nucleotide dissociation rate in the
presence of a 50-fold molar excess of the Tiam1 DH-PH domain
was observed (Fig. 1C). It is generally accepted that both GEFs
and GAPs require activation and membrane recruitment to
FIG.6.Self-activating mechanism of Rac1b. Rac1bGTP accumu-
lates at the membrane because of the missing GDI regulation, the rapid
intrinsic nucleotide exchange, and the impaired GTPase reaction. A
GEF (e.g. Tiam1) may further accelerate the activation process,
whereas a GAP (e.g. p50
GAP
) can effectively catalyze the inactivation,
presuming both regulators are activated and recruited to the mem-
brane. Downstream signaling is PAK independent but possibly involves
p67
PHOX
and other yet unknown effectors (?). The arrows highlight
effective (solid,thick), ineffective (solid,thin), possible (dashed), or no
interaction (crossed).
FIG.5.The nucleotide-binding site. Shown are stereo views of the
final 2F
o
F
c
electron density map around the nucleotide-binding site
of Rac1bGDP (A) and Rac1bGppNHp (B) contoured at 1
. The nucle-
otide is shown as a ball-and-stick representation. Water molecules
inside the nucleotide binding cleft are colored in orange, and the Mg
2
ion is in turquoise.C, structural comparison of the P-loop and switch I
regions of Rac1bGppNHp (red), Rac1GppNHp (brown) (17), and the
nucleotide-free Rac1 in complex with the DH-PH domain of Tiam1
(green) (42). The side chains of Tyr
32
, Ile
33
, and Pro
34
are shown in stick
representation. Thr
35
, GppNHp, and the Mg
2
ion (turquoise) are
shown as ball-and-stick.
Consequences of Alternative Rac1 Splicing4748
by guest on November 18, 2016http://www.jbc.org/Downloaded from
fulfill their regulatory function in the cell. This and the fact
that Rac1b is, without external stimuli, GTP-bound in COS7
cells (Fig. 2A) strongly indicate that, independent of its regu-
lators, Rac1b intrinsically persists in the activated state.
Our biochemical data lead us to the assumption that PAK-
GBD has to stabilize the highly mobile switch regions of Rac1b
by the expense of a 7-fold lower affinity (Table II). Because of
an even more reduced affinity, we could not observe binding of
Rac1b to full-length PAK. This suggests that Rac1b is not able
to interact with PAK in the cellular context. An interaction of
Rac1b with the other major Rac effector, p67
PHOX
has not been
tested and cannot be excluded, because the p67
PHOX
-binding
site on Rac1b is largely conserved.
The fact that Rac1b does interact with GAP and PAK-GBD
(Figs. 1Dand 3) indicates (i) that the highly mobile switch
regions of Rac1b can in principle be recognized by different
binding domains and (ii) that the insertion does not neces-
sarily abolish binding. However, a proper interaction re-
quires a stabilization of the Rac1b switch regions that obvi-
ously causes an overall reduction of binding affinity, as shown
here for PAK-GBD.
ConclusionsThe current study clearly demonstrates how
the 19-amino acid insertion affects two fundamental biochem-
ical properties of small GTP-binding proteins. An impaired
GTP hydrolysis coupled with an accelerated nucleotide disso-
ciation lead to a predominantly GTP-bound protein in the cel-
lular context. The insertion seems to resemble an intrinsic GEF
function by modifying the structure and dynamics of the switch
regions. The fact that Rac1b binds GDP and GTP with compa-
rable affinities and that the P-loop contacts with the phosphate
groups of both nucleotides are unaffected strongly supports the
notion that the missing stabilization of Mg
2
and
-phosphate
binding as well as the highly mobile switch II are the reasons
for the changed properties.
Our data together with the recent report of Matos et al. (11)
enabled us to suggest a concept of Rac1b regulation and its
interaction with downstream targets (Fig. 6). The missing reg-
ulation by GDI most likely keeps Rac1b constitutively mem-
brane-bound. In consideration of the aberrant intrinsic activi-
ties of Rac1b presented in this study, it is tempting to assume
that in contrast to Rac1, Rac1b regulation by GEFs and GAPs
seems to be redundant. Activated GAPs can effectively down-
regulate Rac1b, but it enters immediately a new activation
cycle by a self-activating mechanism. GEFs may contribute to
an even faster exchange of the bound nucleotide. We conclude
that the self-activation, the impaired GTPase reaction, and the
GDI insensitivity are the reasons for the accumulation of GTP-
bound Rac1b in cells. Although the presented overall struc-
tures of both nucleotide-bound forms of Rac1b are remarkably
similar, we cannot exclude the possibility that the missing
switch II region exhibits conformational differences and there-
fore may be responsible for the exclusive recognition and bind-
ing of GppNHp-bound Rac1b by PAK-GBD. These observations
strongly suggest that the very flexible switch regions of Rac1b
may adopt a GTP-bound conformation upon effector binding. In
this state Rac1b may interact with p67
PHOX
and certainly other
yet unknown effectors. However, important questions concern-
ing the involvement of Rac1b in signal transduction and in
tumor progression remain to be elucidated.
AcknowledgmentsWe thank A. Wittinghofer for continuous sup-
port and E. Lengyel for providing the cDNA of human Rac1b. We
thank I. Schlichting, W. Blankenfeld and the staff of the European
Synchrotron Radiation Facility for data collection at the ID14 beam-
line, as well as O. Daumke and A. Ghosh for critical reading of the
manuscript.
REFERENCES
1. Vetter, I. R., and Wittinghofer, A. (2001) Science 294, 12991304
2. Corbett, K. D., and Alber, T. (2001) Trends Biochem. Sci. 26, 706710
3. Bishop, A. L., and Hall, A. (2000) Biochem. J. 348, 241255
4. Gamblin, S. J., and Smerdon, S. J. (1998) Curr. Opin. Struct. Biol. 8, 195201
5. Scheffzek, K., Ahmadian, M. R., and Wittinghofer, A. (1998) Trends Biochem.
Sci. 23, 257262
6. Moon, S. Y., and Zheng, Y. (2003) Trends Cell Biol. 13, 1322
7. Hoffman, G. R., and Cerione, R. A. (2002) FEBS Lett. 513, 8591
8. Olofsson, B. (1999) Cell Signal. 11, 545554
9. Jordan, P., Brazao, R., Boavida, M. G., Gespach, C., and Chastre, E. (1999)
Oncogene 18, 68356839
10. Schnelzer, A., Prechtel, D., Knaus, U., Dehne, K., Gerhard, M., Graeff, H.,
Harbeck, N., Schmitt, M., and Lengyel, E. (2000) Oncogene 19, 30133020
11. Matos, P., Collard, J. G., Jordan, P. (2003) J. Biol. Chem. 278, 5044250448
12. Reid, T., Bathoorn, A., Ahmadian, M. R., and Collard, J. G. (1999) J. Biol.
Chem. 274, 3358733593
13. Buchwald, G., Hostinova, E., Rudolph, M. G., Kraemer, A., Sickmann, A.,
Meyer, H. E., Scheffzek, K., and Wittinghofer, A. (2001) Mol. Cell. Biol. 21,
51795189
14. Fiegen, D., Blumenstein, L., Stege, P., Vetter, I. R., and Ahmadian, M. R.
(2002) FEBS Lett. 525, 100104
15. Ahmadian, M. R., Wittinghofer, A., and Herrmann, C. (2002) Methods Mol.
Biol. 189, 4563
16. Kabsch, W. (1993) J. Appl. Cryst. 26, 795800
17. Navaza, J. (2001) Acta Crystallogr. Sect. D Biol. Crystallogr. 57, 13671372
18. Hirshberg, M., Stockley, R. W., Dodson, G., and Webb, M. R. (1997) Nat.
Struct. Biol. 4, 147152
19. Murshudov, G. N., Vagin, A. A., and Dodson, E. J. (1997) Acta Crystallogr.
Sect. D Biol. Crystallogr. 53, 240255
20. Perrakis, A., Sixma, T. K., Wilson, K. S., and Lamzin, V. S. (1997) Acta
Crystallogr. Sect. D Biol. Crystallogr. 53, 448455
21. Herrmann, C., Martin, G. A., and Wittinghofer, A. (1995) J. Biol. Chem. 270,
29012905
22. Schmidt, G., Lenzen, C., Simon, I., Deuter, R., Cool, R. H., Goody, R. S., and
Wittinghofer, A. (1996) Oncogene 12, 8796
23. Robbe, K., Otto-Bruc, A., Chardin, P., Antonny, B. (2003) J. Biol. Chem. 278,
47564762
24. Xu, X., Wang, Y., Barry, D. C., Chanock, S. J., and Bokoch, G. M. (1997)
Biochemistry 36, 626632
25. Qiu, R. G., Chen, J., Kirn, D., McCormick, F., and Symons, M. (1995) Nature
374, 457459
26. Olson, M. F., Ashworth, A., and Hall, A. (1995) Science 269, 12701272
27. Zohn, I. E., Symons, M., Chrzanowska-Wodnicka, M., Westwick, J. K., and
Der, C. J. (1998) Mol. Cell. Biol. 18, 12251235
28. Mira, J. P., Benard, V., Groffen, J., Sanders, L. C., and Knaus, U. G. (2000)
Proc. Natl. Acad. Sci. U. S. A. 97, 185189
29. Malliri, A., van der Kammen, R. A., Clark, K., van der Valk, M., Michiels, F.,
and Collard, J. G. (2002) Nature 417, 867871
30. Lin, R., Bagrodia, S., Cerione, R., and Manor, D. (1997) Curr. Biol. 7, 794797
31. Herrmann, C., Horn, G., Spaargaren, M., and Wittinghofer, A. (1996) J. Biol.
Chem. 271, 67946800
32. Thompson, G., Owen, D., Chalk, P. A., and Lowe, P. N. (1998) Biochemistry 37,
78857891
33. Owen, D., Mott, H. R., Laue, E. D., and Lowe, P. N. (2000) Biochemistry 39,
12431250
34. Ihara, K., Muraguchi, S., Kato, M., Shimizu, T., Shirakawa, M., Kuroda, S.,
Kaibuchi, K., and Hakoshima, T. (1998) J. Biol. Chem. 273, 96569666
35. Rudolph, M. G., Wittinghofer, A., and Vetter, I. R. (1999) Protein Sci. 8,
778787
36. Bourne, H. R., Sanders, D. A., and McCormick, F. (1991) Nature 349, 117127
37. Menetrey, J., and Cherfils, J. (1999) Proteins 37, 465473
38. Hwang, M. C., Sung, Y. J., and Hwang, Y. W. (1996) J. Biol. Chem. 271,
81968202
39. Ahmadian, M. R., Zor, T., Vogt, D., Kabsch, W., Selinger, Z., Wittinghofer, A.,
and Scheffzek, K. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 70657070
40. Milburn, M. V., Tong, L., deVos, A. M., Brunger, A., Yamaizumi, Z., Nish-
imura, S., and Kim, S. H. (1990) Science 247, 939945
41. John, J., Rensland, H., Schlichting, I., Vetter, I., Borasio, G. D., Goody, R. S.,
and Wittinghofer, A. (1993) J. Biol. Chem. 268, 923929
42. Pan, J. Y., and Wessling-Resnick, M. (1998) Bioessays 20, 516521
43. Spoerner, M., Herrmann, C., Vetter, I. R., Kalbitzer, H. R., and Wittinghofer A.
(2001) Proc. Natl. Acad. Sci. U. S. A. 98, 49444949
44. Worthylake, D. K., Rossman, K. L., and Sondek, J. (2000) Nature 408, 682688
45. Grizot, S., Faure, J., Fieschi, F., Vignais, P. V., Dagher, M. C., and Pebay-
Peyroula, E. (2001) Biochemistry 40, 1000710013
46. Scheffzek, K., Ahmadian, M. R., Kabsch, W., Wiesmuller, L., Lautwein, A.,
Schmitz, F., and Wittinghofer, A. (1997) Science 277, 333338
47. Hoffman, G. R., Nassar, N., and Cerione, R. A. (2000) Cell 100, 345356
48. Rittinger, K., Walker, P. A., Eccleston, J. F., Smerdon, S. J., and Gamblin, S. J.
(1997) Nature 389, 758762
49. Morreale, A., Venkatesan, M., Mott, H. R., Owen, D., Nietlispach, D., Lowe,
P. N., and Laue, E. D. (2000) Nat. Struct. Biol. 7, 384388
50. Lapouge, K., Smith, S. J., Walker, P. A., Gamblin, S. J., Smerdon, S. J., and
Rittinger, K. (2000) Mol. Cell. 6, 899907
51. Haeusler, L. C., Blumenstein, L., Stege, P., Dvorsky, R., and Ahmadian, M. R.
(2003) FEBS Lett. 555, 556560
Consequences of Alternative Rac1 Splicing 4749
by guest on November 18, 2016http://www.jbc.org/Downloaded from
Dvorsky, Ingrid R. Vetter and Mohammad R. Ahmadian
Dennis Fiegen, Lars-Christian Haeusler, Lars Blumenstein, Ulrike Herbrand, Radovan
Alternative Splicing of Rac1 Generates Rac1b, a Self-activating GTPase
doi: 10.1074/jbc.M310281200 originally published online November 18, 2003
2004, 279:4743-4749.J. Biol. Chem.
10.1074/jbc.M310281200Access the most updated version of this article at doi:
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/279/6/4743.full.html#ref-list-1
This article cites 49 references, 17 of which can be accessed free at
by guest on November 18, 2016http://www.jbc.org/Downloaded from
... RAC1B has an additional exon (i.e., exon3b) encoding 19 amino acids with an in-frame insertion just after its Switch-II domain. This leads to a structural change favoring the active GTP-bound state independent of GEF-mediated activation [27]. ...
Article
Full-text available
Breast cancer stem cells (BCSC) are presumed to be responsible for treatment resistance, tumor recurrence and metastasis of breast tumors. However, development of BCSC-targeting therapies has been held back by their heterogeneity and the lack of BCSC-selective molecular targets. Here, we demonstrate that RAC1B, the only known alternatively spliced variant of the small GTPase RAC1, is expressed in a subset of BCSCs in vivo and its function is required for the maintenance of BCSCs and their chemoresistance to doxorubicin. In human breast cancer cell line MCF7, RAC1B is required for BCSC plasticity and chemoresistance to doxorubicin in vitro and for tumor-initiating abilities in vivo. Unlike Rac1, Rac1b function is dispensable for normal mammary gland development and mammary epithelial stem cell (MaSC) activity. In contrast, loss of Rac1b function in a mouse model of breast cancer hampers the BCSC activity and increases their chemosensitivity to doxorubicin treatment. Collectively, our data suggest that RAC1B is a clinically relevant molecular target for the development of BCSC-targeting therapies that may improve the effectiveness of doxorubicin-mediated chemotherapy.
... The splice variant Rac1b, produced from the inclusion of an alternative exon 3b, encodes a constitutively active GTPase protein. 10,11 Rac1b is considered a pro-tumorigenic GTPase that promotes cellular transformation. 12 Likewise, alternative splicing of Cdc42, resulting in the switching of the CDC42-v2 variant to the CDC42-v1 variant, is associated with malignant transformation. ...
Article
Full-text available
Numerous recent studies suggest that cancer‐specific splicing alteration is a critical contributor to the pathogenesis of cancer. RNA‐binding protein with serine‐rich domain 1, RNPS1, is an essential regulator of the splicing process. However, the defined role of RNPS1 in tumorigenesis still remains elusive. We report here that the expression of RNPS1 is higher in cervical carcinoma samples (TCGA‐CESC) compared to the normal tissues. Consistently, the expression of RNPS1 was high in cervical cancer cells compared to a normal cell line. This study shows for the first time that RNPS1 promotes cell proliferation and colony‐forming ability of cervical cancer cells. Importantly, RNPS1 positively regulates migration‐invasion of cervical cancer cells. Intriguingly, depletion of RNPS1 increases the chemosensitivity against the chemotherapeutic drug doxorubicin in cervical cancer cells. Further, we characterized the genome‐wide isoform switching stimulated by RNPS1 in cervical cancer cell. Mechanistically, RNA‐sequencing analysis showed that RNPS1 regulates the generation of tumor‐associated isoforms of key genes, particularly Rac1b, RhoA, MDM4, and WDR1, through alternative splicing. RNPS1 regulates the splicing of Rac1 pre‐mRNA via a specific alternative splicing switch and promotes the formation of its tumorigenic splice variant, Rac1b. While the transcriptional regulation of RhoA has been well studied, the role of alternative splicing in RhoA upregulation in cancer cells is largely unknown. Here, we have shown that the knockdown of RNPS1 in cervical cancer cells leads to the skipping of exons encoding the RAS domain of RhoA, consequently causing decreased expression of RhoA. Collectively, we conclude that the gain of RNPS1 expression may be associated with tumor progression in cervical carcinoma. RNPS1‐mediated alternative splicing favors an active Rac1b/RhoA signaling axis that could contribute to cervical cancer cell invasion and metastasis. Thus, our work unveils a novel role of RNPS1 in the development of cervical cancer.
... exon3b) encoding 19 amino acids with an in-frame insertion just after its Switch-II domain. This leads to a structural change favouring the active GTP-bound state independent of GEFmediated activation [27]. Previous studies have suggested that RAC1B is a tumor-speci c splice variant of RAC1, as its expression during organogenesis ceases in adulthood and only gets upregulated in some solid tumors [24]. ...
Preprint
Full-text available
Breast cancer stem cells (BCSC) are presumed to be responsible for treatment resistance, tumor recurrence and metastasis of breast tumors. However, development of BCSC-targeting therapies has been held back by their heterogeneity and the lack of BCSC-selective molecular targets. Here, we demonstrate that Rac1b, the only known alternatively spliced variant of the small GTPase Rac1, is expressed in a subset of BCSCs in vivo and its function is required for the BCSC maintenance and the chemoresistance of breast tumor cells. In human breast cancer cell line MCF7, RAC1B is required for BCSC plasticity and chemoresistance in vitro and for tumor-initiating abilities in vivo. Unlike Rac1, Rac1b function is dispensable for normal mammary gland development and mammary epithelial stem cell (MaSC) activity. In contrast, loss of Rac1b function in a mouse model of breast cancer hampers BCSC activity in vivo and increases the chemosensitivity of primary tumor cells to doxorubicin. Collectively, our data suggest that RAC1B is a clinically relevant molecular target for the development of BCSC-targeting therapies that will improve the effectiveness of currently available chemotherapy modalities.
... RAC1B results from an alternative splicing event that adds 57 extra nucleotides between codons 75 and 76, which results in 19 extra amino-acid residues immediately behind the switch II motif [51,52]. Importantly, studies of its kinetics revealed that RAC1B has fast-cycling properties [75]. RAC1B was shown not to bind RHOGDI, and in contrast to RAC1/P29S, not to interact with PAK1, or at least, the interaction with full-length PAK1 appears to be abolished [76,77]. ...
Article
Full-text available
Simple Summary For many years, cancer-associated mutations in RHO GTPases were not identified and observations suggesting roles for RHO GTPases in cancer were sparse. Instead, RHO GTPases were considered primarily to regulate cell morphology and cell migration, processes that rely on the dynamic behavior of the cytoskeleton. This notion is in contrast to the RAS proteins, which are famous oncogenes and found to be mutated at high incidence in human cancers. Recent advancements in the tools for large-scale genome analysis have resulted in a paradigm shift and RHO GTPases are today found altered in many cancer types. This review article deals with the recent views on the roles of RHO GTPases in cancer, with a focus on the so-called fast-cycling RHO GTPases. Abstract The RHO GTPases comprise a subfamily within the RAS superfamily of small GTP-hydrolyzing enzymes and have primarily been ascribed roles in regulation of cytoskeletal dynamics in eukaryotic cells. An oncogenic role for the RHO GTPases has been disregarded, as no activating point mutations were found for genes encoding RHO GTPases. Instead, dysregulated expression of RHO GTPases and their regulators have been identified in cancer, often in the context of increased tumor cell migration and invasion. In the new landscape of cancer genomics, activating point mutations in members of the RHO GTPases have been identified, in particular in RAC1, RHOA, and CDC42, which has suggested that RHO GTPases can indeed serve as oncogenes in certain cancer types. This review describes the current knowledge of these cancer-associated mutant RHO GTPases, with a focus on how their altered kinetics can contribute to cancer progression.
... We have previously shown that, in normal colon mucosa about 95% of the pre-mRNA transcribed from the RAC1 gene yields the RAC1 mRNA, but in colon tumours, the alternative transcript encoding RAC1B can increase to up to 20% of the total RAC1 gene-derived mRNA [26,93]. This increase is significant because we found that the resulting RAC1B protein predominantly adopts the active and signallingcompetent conformation, in contrast to RAC1, which is maintained, mostly inactive, in cells through interaction with Rho-GDI [26,28,94]. The process of Caco-2 cell polarization somehow appears to repress RAC1B levels and prime these cells to become responsive to pro-inflammatory stimuli, including IL-6. ...
Article
Full-text available
Simple Summary Tumours are now known to develop more quickly when the tumour cell mass is located in a tissue that shows signs of chronic inflammation. Under such conditions, inflammatory cells from the surrounding tumour microenvironment provide survival signals to which cancer cells respond. We have previously found that some colorectal tumours overexpress the protein RAC1B that sustains tumour cell survival. Here we used a colon mucosa-like in vitro cell model and found that the presence of cancer-associated fibroblasts and pro-inflammatory macrophages stimulated colorectal cells to increase their RAC1B levels. Under these conditions, the secreted survival signals were analysed, and interleukin-6 identified as the main trigger for the increase in RAC1B levels. The results contribute to understand the tumour-promoting effect of inflammation at the molecular level. Abstract An inflammatory microenvironment is a tumour-promoting condition that provides survival signals to which cancer cells respond with gene expression changes. One example is the alternative splicing variant Rat Sarcoma Viral Oncogene Homolog (Ras)-Related C3 Botulinum Toxin Substrate 1 (RAC1)B, which we previously identified in a subset of V-Raf Murine Sarcoma Viral Oncogene Homolog B (BRAF)-mutated colorectal tumours. RAC1B was also increased in samples from inflammatory bowel disease patients or in an acute colitis mouse model. Here, we used an epithelial-like layer of polarized Caco-2 or T84 colorectal cancer (CRC) cells in co-culture with fibroblasts, monocytes or macrophages and analysed the effect on RAC1B expression in the CRC cells by RT-PCR, Western blot and confocal fluorescence microscopy. We found that the presence of cancer-associated fibroblasts and M1 macrophages induced the most significant increase in RAC1B levels in the polarized CRC cells, accompanied by a progressive loss of epithelial organization. Under these conditions, we identified interleukin (IL)-6 as the main trigger for the increase in RAC1B levels, associated with Signal Transducer and Activator of Transcription (STAT)3 activation. IL-6 neutralization by a specific antibody abrogated both RAC1B overexpression and STAT3 phosphorylation in polarized CRC cells. Our data identify that pro-inflammatory extracellular signals from stromal cells can trigger the overexpression of tumour-related RAC1B in polarized CRC cells. The results will help to understand the tumour-promoting effect of inflammation and identify novel therapeutic strategies.
... Competition experiments were carried out by measuring the association of IQGAP1 C794 with RAC1•mGppNHp in the presence and absence of excess amounts of RAC-binding proteins: TIAM1 DH-PH, TRIO DH-PH, DOCK2 DHR2, p50 GAP , PAK1 GBD, plexin B1 RBD, and p67 Phox TRP. Previous studies have shown that the association of these proteins with RAC1•mGppNHp, except for IQGAP, does not lead to a change in fluorescence [50,58,65,75], which is a crucial prerequisite for stopped-flow fluorometric competition experiments. The experiment was based on the concept that an increase in fluorescence upon IQGAP association with RAC1•mGppNHp is attenuated by any one of the RAC-binding proteins that compete with IQGAP for the same binding sites. ...
Article
Full-text available
IQ motif-containing GTPase-activating proteins (IQGAPs) modulate a wide range of cellular processes by acting as scaffolds and driving protein components into distinct signaling networks. Their functional states have been proposed to be controlled by members of the RHO family of GTPases, among other regulators. In this study, we show that IQGAP1 and IQGAP2 can associate with CDC42 and RAC1-like proteins but not with RIF, RHOD, or RHO-like proteins, including RHOA. This seems to be based on the distribution of charged surface residues, which varies significantly among RHO GTPases despite their high sequence homology. Although effector proteins bind first to the highly flexible switch regions of RHO GTPases, additional contacts outside are required for effector activation. Sequence alignment and structural, mutational, and competitive biochemical analyses revealed that RHO GTPases possess paralog-specific residues outside the two highly conserved switch regions that essentially determine the selectivity of RHO GTPase binding to IQGAPs. Amino acid substitution of these specific residues in RHOA to the corresponding residues in RAC1 resulted in RHOA association with IQGAP1. Thus, electrostatics most likely plays a decisive role in these interactions.
Chapter
The Rho family GTPases play an important role in mediating signal transduction in multiple fundamental cellular processes. Most of the Rho family members act as molecular switches and cycle between an inactive GDP-bound state and an active GTP-bound state. Upon GTP binding, Rho GTPases undergo conformational changes to interact with a wide variety of downstream effectors. The dynamic cycling of GTP loading/GTP hydrolysis is key for Rho GTPase functions and is tightly regulated by guanine nucleotide exchange factors, GTPase-activating proteins, GDP-dissociation inhibitors, as well as additional posttranslational modifications. Dysregulation of Rho GTPase signaling pathways is involved in multiple human pathological conditions including cancer, inflammation, and cardiovascular diseases. Among the Rho family members, Rac1 and Rac2 are crucial for the assembly and activation of NADPH oxidases NOX1 and NOX2. This chapter provides an overview of the structural mechanisms of the regulation and function of Rho GTPases, with an emphasis on Rac1 and Rac2 where they apply.KeywordsRho GTPasesRacSignaling effectorsNOX2Rho GEFsRho GAPsRho GDIs
Article
RAC1 is a member of the Rac/Rho GTPase subfamily within the RAS superfamily of small GTP-binding proteins, comprising 3 paralogs playing a critical role in actin cytoskeleton remodeling, cell migration, proliferation and differentiation. De novo missense variants in RAC1 are associated with a rare neurodevelopmental disorder (MRD48) characterized by DD/ID and brain abnormalities coupled with a wide range of additional features. Structural and functional studies have documented either a dominant negative or constitutively active behavior for a subset of mutations. Here, we describe two individuals with previously unreported de novo missense RAC1 variants. We functionally demonstrate their pathogenicity proving a gain-of-function (GoF) effect for both. By reviewing the clinical features of these two individuals and the previously published MRD48 subjects, we further delineate the clinical profile of the disorder, confirming its phenotypic variability. Moreover, we compare the main features of MRD48 with the neurodevelopmental disease caused by GoF variants in the paralog RAC3, highlighting similarities and differences. Finally, we review all previously reported variants in RAC proteins and in the closely related CDC42, providing an updated overview of the spectrum and hotspots of pathogenic variants affecting these functionally related GTPases.
Article
Identifying events that regulate the prenylation and localization of small GTPases will help define new strategies for therapeutic targeting of these proteins in disorders such as cancer, cardiovascular disease, and neurological deficits. Splice variants of the chaperone protein SmgGDS (encoded by RAP1GDS1) are known to regulate prenylation and trafficking of small GTPases. The SmgGDS-607 splice variant regulates prenylation by binding pre-prenylated small GTPases, but the effects of SmgGDS binding to the small GTPase RAC1 versus the splice variant RAC1B are not well defined. Here we report unexpected differences in the prenylation and localization of RAC1 and RAC1B, and their binding to SmgGDS. Compared to RAC1, RAC1B more stably associates with SmgGDS-607, is less prenylated, and accumulates more in the nucleus. We show that the small GTPase DIRAS1 inhibits binding of RAC1 and RAC1B to SmgGDS and reduces their prenylation. These results suggest that prenylation of RAC1 and RAC1B is facilitated by binding to SmgGDS-607, but the greater retention of RAC1B by SmgGDS-607 slows RAC1B prenylation. We show that inhibiting RAC1 prenylation by mutating the CAAX motif promotes RAC1 nuclear accumulation, suggesting that differences in prenylation contribute to the different nuclear localization of RAC1 versus RAC1B. Finally, we demonstrate RAC1 and RAC1B that cannot be prenylated bind GTP in cells, indicating that prenylation is not a prerequisite for activation. We report differential expression of RAC1 and RAC1B transcripts in tissues, consistent with these two splice variants having unique functions that might arise in part from their differences in prenylation and localization.
Preprint
Full-text available
Background RAC1 is involved in many physiological processes such as cytoskeleton regulation, and is highly related to the occurrence and development of various tumors. The global characterization of RAC1 in different tumor backgrounds is necessary to understand the mechanism of RAC1 involved in tumor progression Methods Based on the data of The Cancer Genome Atlas (TCGA), Gene Expression Omnibu (GEO), Human Protein Atlas (HPA), Genotype-Tissue Expression (GTEx) and so on, this paper comprehensively expounded the expression and prognosis of RAC1 in different tumors, and further analyzed the effects of genetic changes (gene mutation, epigenetic changes) of RAC1 on expression and immune infiltration on prognosis. Results RAC1 was highly expressed in most tumors and was associated with poor prognosis. Gene mutation, alternative splicing events and methylation of RAC1 may affect the abnormal expression of RAC1 in tumors. Interestingly, we found that RAC1-related immune cell infiltration was consistent between endothelial cells and fibroblasts, and RAC1-related macrophage infiltration in LIHC and LUAD was consistent with poor patient outcomes. Conclusions We have enriched the tumor-promoting role of RAC1 in different tumor profiles as a whole and speculated the intrinsic link, providing a basis for the search for potential biomarker.
Article
Full-text available
The Ras-binding domain (RBD) of human Raf-1 was purified from Escherichia coli, and its interaction with Ras was investigated. Its dissociation constant with p21•guanyl-5′-yl imidodiphosphate was found to be 18 nM, with a slight preference for H-ras over K- and N-ras. Oncogenic forms bind with slightly lower affinity. The affinity of RBD for effector region mutants or the GDP-bound form of p21 is in the micromolar range, which means that 100-fold lower affinity is not sufficient for signal transduction. The rate of the GTPase of p21 is not modified by RBD. Since Pi release is found not to be rate limiting, the Ras-Raf signal of the cell may be terminated by the intrinsic GTPase of p21.
Article
Full-text available
Uncontrolled cell proliferation is a major feature of cancer. Experimental cellular models have implicated some members of the Rho GTPase family in this process. However, direct evidence for active Rho GTPases in tumors or cancer cell lines has never been provided. In this paper, we show that endogenous, hyperactive Rac3 is present in highly proliferative human breast cancer-derived cell lines and tumor tissues. Rac3 activity results from both its distinct subcellular localization at the membrane and altered regulatory factors affecting the guanine nucleotide state of Rac3. Associated with active Rac3 was deregulated, persistent kinase activity of two isoforms of the Rac effector p21-activated kinase (Pak) and of c-Jun N-terminal kinase (JNK). Introducing dominant-negative Rac3 and Pak1 fragments into a breast cancer cell line revealed that active Rac3 drives Pak and JNK kinase activities by two separate pathways. Only the Rac3-Pak pathway was critical for DNA synthesis, independently of JNK. These findings identify Rac3 as a consistently active Rho GTPase in human cancer cells and suggest an important role for Rac3 and Pak in tumor growth.
Article
Full-text available
The 2.4-A resolution crystal structure of a dominantly active form of the small guanosine triphosphatase (GTPase) RhoA, RhoAV14, complexed with the nonhydrolyzable GTP analogue, guanosine 5'-3-O-(thio)triphosphate (GTPgammaS), reveals a fold similar to RhoA-GDP, which has been recently reported (Wei, Y., Zhang, Y., Derewenda, U., Liu, X., Minor, W., Nakamoto, R. K., Somlyo, A. V., Somlyo, A. P., and Derewenda, Z. S. (1997) Nat. Struct. Biol. 4, 699-703), but shows large conformational differences localized in switch I and switch II. These changes produce hydrophobic patches on the molecular surface of switch I, which has been suggested to be involved in its effector binding. Compared with H-Ras and other GTPases bound to GTP or GTP analogues, the significant conformational differences are located in regions involving switches I and II and part of the antiparallel beta-sheet between switches I and II. Key residues that produce these conformational differences were identified. In addition to these differences, RhoA contains four insertion or deletion sites with an extra helical subdomain that seems to be characteristic of members of the Rho family, including Rac1, but with several variations in details. These sites also display large displacements from those of H-Ras. The ADP-ribosylation residue, Asn41, by C3-like exoenzymes stacks on the indole ring of Trp58 with a hydrogen bond to the main chain of Glu40. The recognition of the guanosine moiety of GTPgammaS by the GTPase contains water-mediated hydrogen bonds, which seem to be common in the Rho family. These structural differences provide an insight into specific interaction sites with the effectors, as well as with modulators such as guanine nucleotide exchange factor (GEF) and guanine nucleotide dissociation inhibitor (GDI).
Article
Full-text available
The Mas oncogene encodes a novel G-protein-coupled receptor that was identified originally as a transforming protein when overexpressed in NIH 3T3 cells. The mechanism and signaling pathways that mediate Mas transformation have not been determined. We observed that the foci of transformed NIH 3T3 cells caused by Mas were similar to those caused by activated Rho and Rac proteins. Therefore, we determined if Mas signaling and transformation are mediated through activation of a specific Rho family protein. First, we observed that, like activated Rac1, Mas cooperated with activated Raf and caused synergistic transformation of NIH 3T3 cells. Second, both Mas- and Rac1-transformed NIH 3T3 cells retained actin stress fibers and showed enhanced membrane ruffling. Third, like Rac, Mas induced lamellipodium formation in porcine aortic endothelial cells. Fourth, Mas and Rac1 strongly activated the JNK and p38, but not ERK, mitogen-activated protein kinases. Fifth, Mas and Rac1 stimulated transcription from common DNA promoter elements: NF-κB, serum response factor (SRF), Jun/ATF-2, and the cyclin D1 promoter. Finally, Mas transformation and some of Mas signaling (SRF and cyclin D1 but not NF-κB activation) were blocked by dominant negative Rac1. Taken together, these observations suggest that Mas transformation is mediated in part by activation of Rac-dependent signaling pathways. Thus, Rho family proteins are common mediators of transformation by a diverse variety of oncogene proteins that include Ras, Dbl family, and G-protein-coupled oncogene proteins.
Article
Full-text available
Ras proteins participate as a molecular switch in the early steps of the signal transduction pathway that is associated with cell growth and differentiation. When the protein is in its GTP complexed form it is active in signal transduction, whereas it is inactive in its GDP complexed form. A comparison of eight three-dimensional structures of ras proteins in four different crystal lattices, five with a nonhydrolyzable GTP analog and three with GDP, reveals that the "on" and "off" states of the switch are distinguished by conformational differences that span a length of more than 40 A, and are induced by the gamma-phosphate. The most significant differences are localized in two regions: residues 30 to 38 (the switch I region) in the second loop and residues 60 to 76 (the switch II region) consisting of the fourth loop and the short alpha-helix that follows the loop. Both regions are highly exposed and form a continuous strip on the molecular surface most likely to be the recognition sites for the effector and receptor molecule(or molecules). The conformational differences also provide a structural basis for understanding the biological and biochemical changes of the proteins due to oncogenic mutations, autophosphorylation, and GTP hydrolysis, and for understanding the interactions with other proteins.
Article
We report a novel crystal form of the small G protein Rap2A in complex with GTP which has no GTPase activity in the crystal. The asymmetric unit contains two complexes which show that a conserved switch I residue, Tyr 32, contributes an extra hydrogen bond to the γ-phosphate of GTP as compared to related structures with GTP analogs. Since GTP is not hydrolyzed in the crystal, this interaction is unlikely to contribute to the intrinsic GTPase activity. The comparison of other G protein structures to the Rap2-GTP complex suggests that an equivalent interaction is likely to exist in their GTP form, whether unbound or bound to an effector. This interaction has to be released to allow the GAP-activated GTPase, and presumably the intrinsic GTPase activity as well. We also discuss the definition of the flexible regions and their hinges in the light of this structure and the expanding database of G protein structures. We propose that the switch I and switch II undergo either partial or complete disorder-to-order transitions according to their cellular status, thus defining a complex energy landscape comprising more than two conformational states. We observe in addition that the region connecting the switch I and switch II is flexible in Rap2 and other G proteins. This region may be important for protein-protein interactions and possibly behave as a conformational lever arm, as characterized for Arf. Taken together, these observations suggest that the structural mechanisms of small G proteins are significantly driven by entropy-based free energy changes. Proteins 1999;37:465–473. ©1999 Wiley-Liss, Inc.
Article
p67phox is an essential part of the NADPH oxidase, a multiprotein enzyme complex that produces superoxide ions in response to microbial infection. Binding of the small GTPase Rac to p67phox is a key step in the assembly of the active enzyme complex. The structure of Rac·GTP bound to the N-terminal TPR (tetratrico-peptide repeat) domain of p67phox reveals a novel mode of Rho family/effector interaction and explains the basis of GTPase specificity. Complex formation is largely mediated by an insertion between two TPR motifs, suggesting an unsuspected versatility of TPR domains in target recognition and in their more general role as scaffolds for the assembly of multiprotein complexes.
Article
The small G proteins of the Ras family act as bimodal relays in the transfer of intracellular signals. This is a dynamic phenomenon involving a cascade of protein–protein interactions modulated by chemical modifications, structural rearrangements and intracellular relocalisations. Most of the small G proteins could be operationally defined as proteins having two conformational states, each of which interacts with different cellular partners. These two states are determined by the nature of the bound nucleotide, GDP or GTP. This capacity to cycle between a GDP-bound conformation and a GTP-bound conformation enables them to filter, to amplify or to temporise the upstream signals that they receive. Thus the control of this cycle is crucial. Membrane anchoring of the proteins in the Ras family is a prerequisite for their activity. Most of the proteins in the Rho/Rac and Rab subfamilies of Ras proteins cycle between cytosol and membranes. Then the control of membrane association/dissociation is an other important regulation level. This review will describe one family of crucial regulators acting on proteins in the Rho/Rac family—the Rho guanine nucleotide dissociation inhibitors, or RhoGDIs. As yet, only three RhoGDIs have been described: RhoGDI-1, RhoGDI-2 (or D4/Ly-GDI) and RhoGDI-3. RhoGDI 1 and 2 are cytosolic and participate in the regulation of both the GDP/GTP cycle and the membrane association/dissociation cycle of Rho/Rac proteins. The non-cytosolic RhoGDI-3 seems to act in a slightly different way.
Article
GTPases are conserved molecular switches, built according to a common structural design. Rapidly accruing knowledge of individual GTPases--crystal structures, biochemical properties, or results of molecular genetic experiments--support and generate hypotheses relating structure to function in other members of the diverse family of GTPases.
Article
Members of the Rho family of small guanosine triphosphatases (GTPases) regulate the organization of the actin cytoskeleton; Rho controls the assembly of actin stress fibers and focal adhesion complexes, Rac regulates actin filament accumulation at the plasma membrane to produce lamellipodia and membrane ruffles, and Cdc42 stimulates the formation of filopodia. When microinjected into quiescent fibroblasts, Rho, Rac, and Cdc42 stimulated cell cycle progression through G1 and subsequent DNA synthesis. Furthermore, microinjection of dominant negative forms of Rac and Cdc42 or of the Rho inhibitor C3 transferase blocked serum-induced DNA synthesis. Unlike Ras, none of the Rho GTPases activated the mitogen-activated protein kinase (MAPK) cascade that contains the protein kinases c-Raf1, MEK (MAPK or ERK kinase), and ERK (extracellular signal-regulated kinase). Instead, Rac and Cdc42, but not Rho, stimulated a distinct MAP kinase, the c-Jun kinase JNK/SAPK (Jun NH2-terminal kinase or stress-activated protein kinase). Rho, Rac, and Cdc42 control signal transduction pathways that are essential for cell growth.