ArticlePDF Available

Abstract

Integrin-mediated platelet adhesion and aggregation are essential for sealing injured blood vessels and preventing blood loss, and excessive platelet aggregation can initiate arterial thrombosis, causing heart attacks and stroke1. To ensure that platelets aggregate only at injury sites, integrins on circulating platelets exist in a low-affinity state and shift to a high-affinity state (in a process known as integrin activation or priming) after contacting a wounded vessel2. The shift is mediated through binding of the cytoskeletal protein Talin to the subunit cytoplasmic tail3, 4, 5. Here we show that platelets lacking the adhesion plaque protein Kindlin-3 cannot activate integrins despite normal Talin expression. As a direct consequence, Kindlin-3 deficiency results in severe bleeding and resistance to arterial thrombosis. Mechanistically, Kindlin-3 can directly bind to regions of -integrin tails distinct from those of Talin and trigger integrin activation. We have therefore identified Kindlin-3 as a novel and essential element for platelet integrin activation in hemostasis and thrombosis.
Kindlin-3 is essential for integrin activation and
platelet aggregation
Markus Moser
1
, Bernhard Nieswandt
2,3
, Siegfried Ussar
1
, Miroslava Pozgajova
2
& Reinhard Fa
¨ssler
1
Integrin-mediated platelet adhesion and aggregation are
essential for sealing injured blood vessels and preventing
blood loss, and excessive platelet aggregation can initiate
arterial thrombosis, causing heart attacks and stroke1.To
ensure that platelets aggregate only at injury sites, integrins
on circulating platelets exist in a low-affinity state and shift
to a high-affinity state (in a process known as integrin
activation or priming) after contacting a wounded vessel2.
The shift is mediated through binding of the cytoskeletal
protein Talin to the bsubunit cytoplasmic tail3–5.Herewe
show that platelets lacking the adhesion plaque protein
Kindlin-3 cannot activate integrins despite normal
Talin expression. As a direct consequence, Kindlin-3 deficiency
results in severe bleeding and resistance to arterial thrombosis.
Mechanistically, Kindlin-3 can directly bind to regions of
b-integrin tails distinct from those of Talin and trigger integrin
activation. We have therefore identified Kindlin-3 as a novel
and essential element for platelet integrin activation in
hemostasis and thrombosis.
Cellular control of integrin activation is essential for virtually all cells,
including platelets, which seal injured blood vessels and stop bleeding.
At sites of injury, the platelet receptors GPIb and GPVI bind to von
Willebrand factor (vWF) and collagen, respectively1–3, which together
with locally produced thrombin trigger activation of integrin a
IIb
b
3
and the release of soluble platelet agonists including ADP and
thromboxane A2 (TxA2). Activated a
IIb
b
3
integrins bind fibrinogen,
vWF and fibronectin, thus allowing firm platelet adhesion and platelet
aggregation. The central role of integrin activation in platelet adhesion
and aggregation sparked the search for critical integrin tail-binding
proteins that control integrin affinity for ligands. Irrespective of the
platelet-activating stimulus and signaling pathways, Talin binding to
the b-integrin tails was shown to be the final common step in a
IIb
b
3
integrin activation and ligand binding4,6,7. Talin, a major cytoskeletal
protein at integrin adhesion sites, consists of a large C-terminal rod-
like domain and an N-terminal FERM (protein 4.1, ezrin, radixin,
moesin) domain with three subdomains: F1, F2 and F3 (ref. 8). The
phosphotyrosine-binding (PTB) subdomain in the F3 domain sequen-
tially binds to two distinct regions in the bcytoplasmic tails and is
sufficient for integrin activation in vitro8,9.
In addition to Talin, other FERM domain–containing proteins,
including the Kindlins, interact with integrin btails10. The Kindlin
protein family consist of three members (Kindlin-1, Kindlin-2 and
Kindlin-3), all of which localize to integrin adhesion sites11–13.In
contrast to the widely expressed Kindlin-1 and Kindlin-2, Kindlin-3 is
restricted to hematopoietic cells and is particularly abundant in
megakaryocytes and platelets12. The structural hallmark of Kindlins
is a FERM domain whose F2 subdomain is split by a pleckstrin
homology (PH) domain. In a comparison of FERM-domain proteins,
the F3 subdomains of Kindlins have been found to share highest
homology with the F3 domain of Talin10.
To address the function of Kindlin-3 in vivo, we disrupted the Kind3
(also called Fermt3)geneinmice(Supplementary Fig. 1 online). Mice
heterozygous for the Kindlin-3–null mutation (Kind3
+/–
)werenor-
mal, whereas mice lacking Kindlin-3 (Kind3
/
;Fig. 1a) died within a
week of birth and showed a pronounced osteopetrosis (unpublished
data) and severe hemorrhages in the gastrointestinal tract, skin, brain
and bladder, which were already apparent during development
(Fig. 1b and data not shown). To test whether the severe bleeding
of Kindlin-3–deficient mice was due to impaired platelet production
and/or function, we generated fetal liver cell chimeras by transferring
either Kind3
/
or wild-type fetal liver cells into lethally irradiated
wild-type recipient mice. Tail-bleed assays revealed that Kind3
/
chimeras suffer from a pronounced hemostatic defect like that of
Kindlin-3–deficient mice. After the tail-tip cut, bleeding in control
mice arrested within 10 min (mean of 5.4 ± 4.3 min), whereas
Kind3
/
chimeras bled for longer than 15 min, suggesting severe
platelet dysfunction (Fig. 1c).
Kind3
/
chimeras showed platelet counts similar to those of wild-
type chimeras (Fig. 1d), ruling out an essential role for Kindlin-3 in
platelet formation. Analysis of glycoprotein abundance on platelets
revealed elevated levels of the vWF receptor complex GPIb-IX in
Kindlin-3–deficient as compared to wild-type platelets, whereas levels
of other glycoproteins, including GPVI, CD9, and b
1
and b
3
integrins,
were reduced (Supplementary Table 1 online). Thus Kindlin-3 has an
apparent yet undefined role in the expression of several glycoproteins.
However, the reduced expression of integrin a
IIb
b
3
does not account
for the hemostasis defect, as mice carrying a heterozygous null
mutation in the b
3
integrin express even less a
IIb
b
3
integrin on their
platelets (50% of wild-type) without developing a bleeding defect14.
Received 11 October 2007; accepted 4 January 2008; published online 17 February 2008; doi:10.1038/nm1722
1
Department of Molecular Medicine, Max Planck Institute of Biochemistry, Am Klopferspitz 18, 82152 Martinsried, Germany.
2
University of Wu
¨rzburg, Rudolf Virchow
Center, Deutsche Forschungsgemeinschaft Research Center for Experimental Biomedicine, Zinklesweg 10, 97080 Wu
¨rzburg, Germany.
3
Institute of Clinical
Biochemistry and Pathobiochemisty, Josef-Schneider-Str. 2, 97078 Wu
¨rzburg, Germany. Correspondence should be addressed to R.F. (faessler@biochem.mpg.de).
NATURE MEDICINE VOLUME 14
[
NUMBER 3
[
MARCH 2008 325
LETTERS
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
To determine the mechanism of the platelet defect, we performed
ex vivo platelet aggregation studies. Wild-type platelets aggregated in
response to ADP, the TxA2 analog U46619, thrombin, collagen and
the GPVI-activating collagen-related peptide (CRP), whereas none of
the agonists induced aggregation of Kind3
/
platelets (Fig. 2a).
Notably, all agonists induced a comparable activation-dependent
change from discoid to spherical shape in control and Kind3
/
platelets, which can be seen in aggregometry as a short decrease in
light transmission following the addition of agonists. This suggests a
selective defect in a
IIb
b
3
-dependent aggregation rather than a general
impairment of signaling pathways in Kind3
/
platelets.
To test whether activation of a
IIb
b
3
integrin is indeed abrogated in
Kind3
/
platelets, we determined their ability to bind fibrinogen
using flow cytometry. Wild-type platelets showed robust fibrinogen
binding in response to ADP, ADP plus U46619 and CRP, whereas
Kind3
/
platelets failed to bind fibrinogen upon agonist treatment
(Fig. 2b). The antibody JON/A-PE, which specifically detects the
activated form of mouse a
IIb
b
3
integrin14, likewise did not bind
stimulated Kind3
/
platelets (Fig. 2c). When cellular activation of
a
IIb
b
3
integrin was bypassed by the addition of MnCl
2
, comparable
fibrinogen binding to Kind3
/
and wild-type platelets occurred
(Fig. 2b). Together, these findings indicate that loss of Kindlin-3
expression prevents energy-dependent conformational rearrangements
required for integrin-a
IIb
b
3
activation.
Resting platelets store P-selectin in a-granules, which fuse with the
plasma membrane during agonist-induced platelet activation1. Both
CRP and thrombin induced P-selectin translocation on wild-type and
Kind3
/
platelets, although a significant reduction was consistently
observed at intermediate concentrations of thrombin (Po0.001;
Fig. 2d). As expected, the weak agonist ADP did not induce P-selectin
surface expression in wild-type and Kind3
/
platelets. Thus, although
loss of Kindlin-3 specifically disables integrin activation, it still permits
agonist-induced P-selectin translocation.
Integrin-a
IIb
b
3
is also involved in adhesion to immobilized ligands,
including collagen-bound vWF, where it acts in concert with the
collagen-binding a
2
b
1
integrin2,15. We analyzed the ability of Kind3
/
platelets to interact with fibrous collagen in a whole-blood perfusion
Spleen
Kindlin-3
Kind3+/+
Kind3+/–
Kind3–/–
Kind3+/+
Kind3–/–
GAPDH
Platelets
cd
ab
> 900
900
600
Kind3+/+ Kind3 –/–
Kind3+/+
Kind3–/–
Tail bleeding times (s)
300
0
1,000
Platelets/µl (10–6)
800
600
400
200
0
ADP
5 µM
Light transmission
U46619
1 µM
Thrombin
1 U/ml
CRP
5 µg/ml
Collagen
3 µg/ml
Time
Kind3+/+
Kind3+/+
Kind3–/–
Kind3–/–
Kind3+/+
Kind3+/+
Kind3–/–
Kind3–/–
10
Rest. ADP
(µM)
CRP
(µg/ml)
Thrombin
(U/ml)
9 min 25 min
10 0.1 0.01 0.001
1 min
Kind3+/+
Kind3–/–
Kind3+/+
Kind3–/–
1,000
800
600
400
200
0
MFI
(Alexa-488–fibrinogen)
Rest. ADP ADP
+
U46619
CRP Mn2+
600
400
MFI
200
0
10
Rest. ADP
(µM)
CRP
(µg/ml)
Thrombin
(U/ml)
10 0.1 0.01 0.001
Kind3+/+
Kind3–/–
150
100
MFI
50
0
** 60
40
20
0
Surface coverage (%)
0
10
20
30
40
Occlusion time
(min)
>40
abc
def
Figure 2 Impaired platelet function in Kind3
/
mice. (a) Platelet aggregation assay reveals impaired aggregation of Kind3
/
platelets (gray lines) in
response to ADP, U46619, thrombin, CRP and collagen when compared with control platelets (black lines). Arrows denote addition of agonist. (b) Wild-type
(Kind
+/+
) platelets (black bars), but not Kind3
/
platelets (gray bars), bind fibrinogen in response to ADP (10 mM), ADP (10 mM) plus U46619 (3 mM) or
CRP (10 mg/ml). Treatment with MnCl
2
(Mn
2+
; 3 mM) triggers comparable binding. Resting (rest.) platelets were used as a control. (c,d)Kind3
/
platelets
(gray bars) show a complete block in activation of integrin-a
IIb
b
3
after stimulation with ADP (10 mM), CRP (10 mg/ml) and different concentrations (0.001–0.1
U/ml) of thrombin (c), whereas platelet degranulation measured by the surface expression of P-selectin is largely intact after the same treatments (d). Wild-type
platelets (black bars) were used as a control. At the intermediate thrombin concentration, moderately but significantly reduced degranulation was observed with
mutant platelets (** Po0.01). MFI, mean fluorescence intensity. (e)Kind3
/
platelets in whole blood failed to form thrombi when perfused over a
collagen-coated (0.25 mg/ml) surface at a wall shear rate of 1,000 s
–1
. Scale bar, 30 mm. (f) Mesenteric arterioles were injured with FeCl
3
, and adhesion
and thrombus formation of fluorescently labeled platelets were monitored by in vivo video microscopy. Representative images (left) and time to vessel
occlusion (right) are shown. Each symbol represents one individual. Scale bar, 30 mm.
Figure 1 Kindlin-3–deficient animals show severe hemorrhages. (a) Western
blot analyses from spleen and platelet lysates of wild-type (Kind3
+/+
),
heterozygous (Kind3
+/
) and Kindlin-3–deficient (Kind3
/
)mice.
(b) E15.5 embryos reveal severe bleeding. Postnatally, Kind3
/
mice
show skin (arrowhead) and intestinal (arrows) bleeding. All scale bars,
1mm.(c) Tail-bleeding times in wild-type and Kind3
/
mice.
(d) Peripheral platelet counts in wild-type and Kind3
/
chimeras.
LETTERS
326 VOLUME 14
[
NUMBER 3
[
MARCH 2008 NATURE MEDICINE
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
assay. Under high- (1,000 s
–1
,Fig. 2e) or low-shear (150 s
–1
,datanot
shown) flow conditions, wild-type platelets adhered to the collagen
fibers and rapidly built stable three-dimensional aggregates (Fig. 2e).
In contrast, Kind3
/
platelets never established stable adhesions and
either detached immediately or translocated along the fibers for a few
seconds, resulting in virtually no platelets attaching to the collagen-
coated surface at the end of the 4-min perfusion time (Fig. 2e).
Furthermore, agonist-stimulated Kind3
/
platelets failed to bind the
monoclonal anti-b
1
integrin antibody 9EG7, which specifically recog-
nizes the activated form of b
1
integrins16, and also failed to adhere to
soluble, pepsin-digested collagen type I under static conditions (Sup-
plementary Fig. 2 online), a process known to be mediated exclusively
by a
2
b
1
integrins17. Together these findings indicate that Kind3
/
platelets can establish initial contacts with vWF/collagen via GPIb and
probably GPVI, but are unable to adhere firmly as a result of defective
activation of a
IIb
b
3
and a
2
b
1
integrins.
As platelet aggregation may lead to pathological occlusive thrombus
formation, we examined whether lack of Kindlin-3 is protective
against ischemia and infarction after mesenteric arteriole injury.
This injury was induced by ferric chloride and assessed by in vivo
fluorescence microscopy. Five minutes after injury, numerous platelets
adhered firmly to the denuded vessel wall in control chimeras (5,380 ±
2,465/mm
2
); after approximately 10 min, the first thrombi
were observed; and after 18–39 min, the vessels were occluded
(mean occlusion time: 27.6 ± 8.1 min; Fig. 2f). In contrast, a few
Kind3
/
platelets transiently (o5 s) attached to the injured vessel
wall, but virtually none adhered firmly throughout the 40-min
observation period (50 ± 24 / mm
2
). Furthermore, no thrombi
formed in the injured vessels of Kind3
/
chimeras, and blood flow
was maintained in all vessels tested (Fig. 2f). These results confirm the
ex vivo results and underscore the pivotal function of Kindlin-3 in
integrin-mediated platelet adhesion to injured vessels in vivo.
The requirement for Kindlin-3 to trigger agonist-induced integrin
activation on platelets, integrin-mediated platelet adhesion and
thrombus formation suggests that it may be a downstream target of
cellular signaling pathways that activate integrins. To test whether
Kindlin-3 is able to activate integrins, we overexpressed Kindlin-3 in
integrin-a
IIb
b
3
–overexpressing CHO cells. In these cells Kindlin-3 was
unable to trigger integrin activation, likely because the hematopoietic
Kindlin-3 is not recruited to integrin containing focal adhesions13.
Kindlin-3 overexpression in the mouse macrophage cell line RAW
264.7 (RAW), however, yielded a 2.2-fold increase in binding of the
Cy5-labeled fibronectin repeat 7-10 (FN7-10), which harbors the
integrin-binding RGD motif (Fig. 3a). Enhanced green fluorescent
protein (EGFP)-transfected RAW cells showed virtually no increase in
FN7-10 binding as compared to untransfected cells, whereas Talin
overexpression and treatment with Mn
2+
induced a 2.5-fold increase.
Notably, overexpression of a Kindlin-3 variant with a point mutation
in the PTB-like domain (Q597A), which in Talin (R358) reduces
binding to btails4, did not trigger FN7-10 binding. These data
indicate that Kindlin-3 is capable of activating FN binding integrins
and that this activity requires an intact PTB-like domain.
How does Kindlin-3 activate integrins? As reduced Talin expression
did not account for the defect (Fig. 3b), we investigated whether the
FERM domain of Kindlin-3, which is similar to that of Talin, might
play a direct role in integrin activation. As previously shown, recom-
binant Talin bound wild-type b
1
and b
3
integrin tails, and this binding
was significantly reduced when alanine mutations were introduced
into the membrane-proximal tryptophan residue or NPxY motif
(b
3
W739A, b
3
Y747A, b
1
W780A and b
1
Y788A)18. Kindlin-3 was also
able to interact with the wild-type b
1
and b
3
integrin tails (Fig. 3c), in
thepresenceandabsenceofTalin1(Supplementary Fig. 3 online),
and the F3 subdomain of Kindlin-3 was sufficient for this interaction
and this interaction occurred in a direct manner (Fig. 3d). However,
specific point mutations within the b
1
and b
3
integrin cytoplasmic
tails revealed that the binding properties of Kindlin-3 were different
from those of Talin, as the former still bound to b
3
W739A, b
3
Y747A,
b
1
W780A and b
1
Y788A tails, although less efficiently than to wild-
type tails (Fig. 3e,f). Mutations to the membrane distal NxxY motif of
the b
1
and b
3
tails (b
1
Y800A, b
3
Y759A) and the b
1
TT793/794AA and
3
2
1
0
EGFP
Fold FN
binding
Kindlin-3
Kindlin-3
Q597A
Talin
EGFP
+ Mn2+
P = 0.045
P =
0.019
Kindlin-3
Kindlin-3
Coomassie 10 kDa
GST-K3F3
αllb
β3
Coomassie
Tal i n
5% input
αllb
β1
β3
+/+
–/–
Tal i n
GAPDH
Kindlin-3
Coomassie
5% input
GST
β3β3S752P
a
g
bc d
Tal i n
Kindlin-3
Coomassie
5% input
GST
β1β1Y788A
β1Y800A
β1W780A
β1TT793/794AA
αllb
e
Tal i n
Kindlin-3
Coomassie
5% input
GST
β3β3Y747A
β3Y759A
β3W739A
β3ST752/753AA
αllb
f
Figure 3 Biochemical analyses of Kind3
/
platelets. (a) Binding of Cy-5–
labeled fibronectin III 7-10 fragments (FNIII7-10; FN) to RAW 264.7 cells
transfected with EGFP, Talin, EGFP-Kindlin-3 or EGFP-Kindlin-3(Q597A).
Binding of Cy-5–labeled fibronectin III 7-10 lacking the RGD binding loop
(FNIII7-10DRGD) was used to estimate nonspecific binding, and binding in the
presence of 5 mM Mn
2+
served as a positive control. Data shown are mean ±
s.e.m. of five independent experiments, normalized to FNIII7-10 binding to
EGFP-transfected cells and with the background binding of FNIII7-10DRGD
subtracted. Pvalues were obtained by unpaired t-test. (b) Western blot analysis
of lysates from wild-type (+/+) control and Kind3
/
(–/–) platelets.
(c) Pull-down experiment with His-tagged aIIb, b
1
and b
3
integrin tails incubated with 100 mg platelet lysate. (d) Protein-protein interaction assay of
recombinant GST–Kindlin-3 F3 domain (GST-K3F3) incubated with His-tagged aIIb and b
3
integrin tails reveals direct binding between the F3 domain of
Kindlin-3 and the b
3
tail. GST pull-down experiments were performed with recombinant GST-b
1
A (wild-type), GST-b
1
Y788A, GST-b
1
Y800A, GST-b
1
W780A
and GST-b
1
TT793/794AA (e); with GST-b
3
(wild-type), GST-b
3
Y747A, GST-b
3
Y759A, GST-b
3
W739A and GST-b
3
ST752/753AA (f); and with GST-b
3
(wild-
type) and GST-b
3
S752A (g) integrin cytoplasmic domains after incubation with 100 mg platelet lysate. GST protein and GST-aIIb were used as controls. 5%
of the platelet lysate used for the pull-down experiment is shown as input control. Bound Talin and Kindlin-3 proteins were detected by western blotting.
Coomassie blue staining showed that equal amount of GST fusion proteins were used. Shown are results from pull-down assays representative of a minimum
of seven experiments.
LETTERS
NATURE MEDICINE VOLUME 14
[
NUMBER 3
[
MARCH 2008 327
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
b
3
ST752/753AA, however, abolished Kindlin-3 but not Talin binding
(Fig. 3e,f). Moreover, the b
3
S752P mutation found in a subset of
individuals with Glanzmann’s thrombasthenia also abolished Kindlin-
3binding
19 (Fig. 3g). These data indicate that Talin primarily requires
membrane-proximal residues for binding, whereas Kindlin-3 requires
membrane-distal residues for binding to the b
1
and b
3
tails.
Ligand-occupied integrins transduce signals that lead to the activa-
tion of Src family kinases, resulting in cell spreading (outside-in
signaling). We tested the role of Kindlin-3 in outside-in signaling by
analyzing the adhesion of washed control and Kind3
/
platelets to
fibrinogen under static conditions. As mouse platelets, in contrast to
human platelets, do not spread well on immobilized fibrinogen
without cellular activation20, we performed the experiments in the
presence of 0.01 U/ml thrombin, with and without added Mn
2+
.
Comparable adhesion of control and Kind3
/
platelets to the
fibrinogen matrix occurred, confirming a previous observation that
integrin-a
IIb
b
3
activation is not required for static adhesion of platelets
to fibrinogen21. However, whereas control platelets readily formed
lamellipodia and spread within 10–15 min, Kind3
/
platelets only
formed filopodia, with occasional transient small lamellipodia, and
completely failed to spread for up to 45 min (Fig. 4a,b). We obtained
similar results when we carried out adhesion in the presence of
Mn
2+
(Fig. 4c). Thus, Kindlin-3 is also required for integrin a
IIb
b
3
-
dependent outside-in signaling.
Our study demonstrates that Kindlin-3 is essential for platelet
integrin activation and subsequent integrin outside-in signaling.
Furthermore, we found that Kindlin-3 regulates activation of both
b
3
(a
IIb
b
3
)andb
1
(a
2
b
1
) integrins, suggesting that Kindlin-3, like
Talin, is a general regulator of integrin activation. We propose that this
regulatory mechanism is mediated through a direct interaction
between the PTB site of the F3 domain in Kindlin-3 and the integrin
b
3
and b
1
tails, including their distal NxxY motifs. Because Talin
binding requires an intact proximal NPxY motif, our findings raise
questions regarding the roles of Kindlin-3 and Talin in integrin
activation and the hierarchy of their binding to the integrin btails.
Future studies on the structure of the Kindlin-3–integrin complex are
required to examine the relative roles of Kindlin-3 and Talin inter-
actions with integrin tails so as to fully understand how these receptors
become activated. Finally, we show that elimination of Kindlin-3
prevents the formation of pathological thrombi. As Kindlin-3 is
selectively expressed in cells of hematopoietic origin, it may serve as
a potential target for the design of therapeutics aimed at specifically
disrupting integrin activation in platelets.
METHODS
Inactivation of the Kindlin-3 gene. A BAC clone containing the Kind3
(Fermt3) gene was isolated and used to generate the targeting construct
containing an IRES-lacZ-neo cassette between exons 3 and 6. The targeting
vector was electroporated into R1 ES cells, and targeted ES-cell clones were
identified by southern blotting and injected into host blastocysts to generate
germline chimeras.
Generation of fetal liver cell chimeras. Fetal liver cells from E15 wild-type and
Kind3
/
embryos were obtained by pushing the liver through a cell strainer
(Falcon). 4 10
6
cells were injected into the tail vein of lethally irradiated
(10 Gy) recipient C57BL/6 mice. At 3–4 weeks after transfer, platelets were
isolated from whole blood collected from the retro-orbital plexus.
Western blot analysis. Platelet lysates were subjected to a 5–15% gradient
SDS-PAGE. After blotting, PVDF membranes were probed with anti–Kindlin-3
(ref. 12), anti-Talin (Sigma) and anti-GAPDH (Chemicon).
GST fusion protein pull-down assays. The b
1
Aandb
3
integrin cyto-
plasmic domains and their mutant forms (GST-b
1
Y788A, GST-b
1
Y800A,
GST-b
1
W780A and GST-b
1
TT793/794AA; GST-b
3
Y747A, GST-b
3
Y759A,
GST-b
3
W739A, GST-b
3
ST752/753AA and GST-b
3
S752A) were expressed as
GST-fusion proteins in BL21 cells upon induction with 1 mM IPTG. Bacteria
were washed and lysed in buffer A (150 mM NaCl, 1 mM EDTA, 10 mM
Tris, pH 8) containing 100 mg/ml lysozyme for 15 min on ice and then
sonicated. After dialysis against buffer B (100 mM NaCl, 50 mM Tris
pH 7.5, 1% NP-40, 10% glycerol, 2 mM MgCl
2
), GST-fusion proteins were
bound to glutathione-Sepharose beads (Novagen), eluted in 50 mM Tris,
pH 8, 20 mM glutathione, and dialyzed against buffer C (20 mM Tris
pH 7.5, 20% glycerol, 100 mM KCl, 0.2 mM EDTA, 2 mM DTT,
10 mM b-glycerophosphate).
For GST pull-down experiments, GST fusion proteins were bound to
glutathione-Sepharose beads for 1 h at room temperature (B20 1C) in buffer
A. Fresh platelet lysates were incubated with GST or GST-integrin cytoplasmic-
domain fusion proteins for 4 h or overnight at 4 1C. The glutathione-Sepharose
beads were washed four times with buffer A containing 1% Triton X-100 and
10 mM EDTA. Bound proteins were eluted from the beads by boiling in
Laemmli buffer (2% SDS, 10% glycerol, 5% 2-mercaptoethanol, 0.002%
bromphenol blue, 62.5 mM Tris-HCl, pH 6.8) after separation by a SDS-PAGE
and western blotting.
For direct protein-protein interaction assay, recombinant Kindlin-3 F3
domain, spanning amino acids 550–665 (GST-K3F3), was expressed in BL21
cells as described above. Histidine (His)-tagged integrin cytoplasmic tails
were expressed in BL21 bacteria upon induction with 1 mM IPTG and
purified under denaturing conditions. Ten micrograms of GST-K3F3 was
incubated with His-tagged aIIb, and b
3
integrin cytoplasmic tails bound to
Ni
2+
-coated beads for 2 h in buffer D (50 mM NaCl, 10 mM PIPES, 150 mM
sucrose, 0.1% Triton X-100, pH 6.8) containing phosphatase and protease
inhibitor cocktails (Sigma, Roche). After being washed in buffer D, bound
proteins were analyzed by SDS-PAGE and western blotting. Loading of
Ni
2+
-coated beads with the recombinant integrin tails was assessed by
Coomassie Blue staining.
a
b
c
0 min 5 min 15 min
100
80
60
Percentage
of lamellipodia
forming platelets
40
20
0Kind3+/+ Kind3 –/–
Figure 4 Defective adhesion and spreading of Kind3
/
platelets.
(a) Washed wild-type and Kind3
/
platelets were stimulated with
0.01 U/ml thrombin and then allowed to adhere to immobilized fibrinogen
for 15 min. Scale bar, 5 mm. (b) Scanning electron microscopy of wild-type
and Kind3
/
platelets after thrombin stimulation and adhesion to
fibrinogen for 30 min. Scale bars, 1 mm. (c) Washed platelets were allowed
to adhere to immobilized fibrinogen in the presence of 3 mM Mn
2+
for
30 min. Left, representative DIC images. Scale bar, 5 mm. Right, number
of lamellipodia-forming platelets (% of adherent platelets; mean ± s.d.
of four experiments per group).
LETTERS
328 VOLUME 14
[
NUMBER 3
[
MARCH 2008 NATURE MEDICINE
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
Fibronectin binding assay. RAW 264.7 cells were electroporated with 4 mgof
the indicated DNAs using a Macrophage Kit from the Amaxa system. Twenty-
four hours after transfection, cells were trypsinized, washed in FACS Tris-buffer
(24 mM Tris-HCl, pH 7.4, 137 mM NaCl and 2.7 mM KCl) and incubated for
40 min with 0.3 mM Cy5-labeled recombinant fibronectin III 7-10 fragment or
FNIII 7-10DRGD fragment22. As a positive control, EGFP-transfected cells were
incubated with 5 mM Mn
2+
. After washing, the amount of cell-bound
fibronectin fragment was measured with a FACSCalibur (Becton Dickinson).
Dead cells were excluded from FACS analysis by addition of 2.5 mg/ml
propidium iodide and gating for living (propidium iodide negative) and
EGFP-positive cells.
Chemicals. The anesthetic drugs xylazine (Rompun) and ketamine (Imalgene
1000) were purchased from Bayer and Me
´rial, respectively. High-molecular-
weight heparin and human fibrinogen and ADP were from Sigma and
collagen was from Kollagenreagent Horm, Nycomed. Monoclonal antibodies
conjugated to either fluorescein isothiocyanate (FITC) or phycoerythrin (PE)
were from Emfret Analytics. Alexa-Fluor-488–labeled fibrinogen was from
Molecular Probes.
Aggregometry. To determine platelet aggregation, light transmission was
measured using washed platelets (2 10
8
/ml) in the presence of 70 mg/ml
human fibrinogen. Transmission was recorded on a Fibrintimer 4 channel
aggregometer (APACT Laborgera
¨te und Analysensysteme) over 10 min and was
expressed as arbitrary units with transmission through buffer defined as
100% transmission.
Flow cytometry. Heparinized whole blood was diluted 1:20 with modified
Tyrode’s-HEPES buffer (134 mM NaCl, 0.34 mM Na
2
HPO
4
,2.9mMKCl,
12 mM NaHCO
3
, 20 mM HEPES, pH 7.0) containing 5 mM glucose, 0.35%
bovine serum albumin (BSA) and 1 mM CaCl
2
. For assessment of glycoprotein
expression and platelet count, blood samples were incubated with appropriate
fluorophore-conjugated monoclonal antibodies for 15 min at room tempera-
ture and directly analyzed on a FACSCalibur (Becton Dickinson). Activation
studies were performed with blood samples washed twice with modified
Tyrode’s-HEPES buffer, which then were activated with the indicated agonists
or 3 mM MnCl
2
for 15 min, stained with fluorophore-labeled antibodies for
15 min at room temperature and directly analyzed.
Adhesion under flow conditions. Rectangular coverslips (24 60 mm) were
coated with 0.25 mg/ml fibrillar type I collagen (Nycomed) for 1 h at 37 1Cand
blocked with 1% BSA. Perfusion of heparinized whole blood was performed as
described15. Briefly, transparent flow chambers with a slit depth of 50 mm,
equipped with the collagen-coated coverslips, were rinsed with HEPES buffer,
pH 7.45, and connected to a syringe filled with the anticoagulated blood.
Perfusion was carried out at room temperature using a pulse-free pump at low
(150 s
–1
) and high shear stress (1,000 s
–1
). During perfusion, microscopic
phase-contrast images were recorded in real time. Thereafter, the chambers
were rinsed by a 10-min perfusion with HEPES buffer, pH 7.45, at the same
shear stress, and phase-contrast pictures were recorded from at least five
different microscopic fields (63objectives). Image analysis was performed
off-line using Metamorph software (Visitron). Thrombus formation results are
expressed as the mean percentage of total area covered by thrombi.
Analysis of bleeding time. Mice were anesthetized and a 3-mm segment of the
tail tip was cut off with a scalpel. Tail bleeding was monitored by gently
absorbing the bead of blood with a filter paper without contacting the wound
site. When no blood was observed on the paper after 15-s intervals, bleeding
was determined to have ceased. The experiment was stopped after 15 min.
Intravital microscopy of thrombus formation in FeCl
3
injured mesenteric
arterioles. Mice 4–5 weeks old were anesthetized, and the mesentery was gently
exteriorized through a midline abdominal incision. Arterioles (35–60-mm
diameter) were visualized with a Zeiss Axiovert 200 inverted microscope
(10) equipped with a 100-W HBO fluorescent lamp source and a
CoolSNAP-EZ camera (Visitron). Digital images were recorded and analyzed
off-line using Metavue software (Visitron). Injury was induced by topical
application of a 3-mm
2
filter paper tip saturated with FeCl
3
(20%) for 10 s.
Adhesion and aggregation of fluorescently labeled platelets in arterioles were
monitored for 40 min or until complete occlusion occurred (blood flow
stopped for 41min).
Platelet spreading. Cover slips were coated overnight with 1 mg/ml human
fibrinogen and then blocked for 1 h with 1% BSA in PBS. Washed platelets of
wild-type or Kind3
/
mice were resuspended at a concentration of 0.5 10
6
platelets/ml and then further diluted 1:10 in Tyrode’s-HEPES buffer. Shortly
before platelets were seeded on the fibrinogen-coated coverslip, they were acti-
vatedwith0.01U/mlthrombin.Plateletswereallowedtospreadfor30minand
analyzed by differential interference contrast (DIC) microscopy. In parallel,
platelets were fixed in 2.5% glutaraldehyde in Tyrode’s-HEPES buffer and
processed for scanning electron microscopy as previously described23.Inanother
set of experiments, washed platelets were allowed to adhere to fibrinogen in the
presence of 3 mM Mn
2+
without thrombin stimulation and analyzed as above.
Note: Supplementary information is available on the Nature Medicine website.
ACKNOWLEDGMENTS
We thank D. Calderwood (Yale University) and I. Campbell (Oxford University)
for recombinant integrin tails and integrin tail expression vectors and help with
pull-down assays, G. Wanner for imaging of platelets by scanning electron
microscopy, M. Sixt and M. Boesl for mouse manipulation experiments, and
R. Zent, A. Pozzi, M. Humphries and M. Schwartz for critical reading of the
manuscript. This work was supported by the Deutsche Forschungsgemeinschaft
and the Max Planck Society.
AUTHOR CONTRIBUTIONS
M.M. and R.F. designed and supervised research. M.M., B.N. and R.F. wrote the
manuscript. M.M., B.N., S.U. and M.P. performed experiments. All authors
discussed the results and commented on the manuscript.
Published o nline at http: //www.nature.com/n aturemedicine
Reprints and permissions information is available online at http://npg.nature.com/
reprintsandpermissions
1. Ruggeri, Z.M. Platelets in atherothrombosis. Nat. Med. 8, 1227–1234 (2002).
2. Savage, B., Almus-Jacobs, F. & Ruggeri, Z.M. Specific synergy of multiple substrate-
receptor interactions in platelet thrombus formation under flow. Cell 94, 657–666
(1998).
3. Nieswandt, B. & Watson, S.P. Platelet-collagen interaction: is GPVI the central
receptor? Blood 102, 449–461 (2003).
4. Tadokoro, S. et al. Talin binding to integrin btails: a final common step in integrin
activation. Science 302, 103–106 (2003).
5. Critchley, D.R. Cytoskeletal protein talin and vinculin in integrin-mediated adhesion.
Biochem. Soc. Trans. 32, 831–836 (2004).
6. Nieswandt, B. et al. Loss of talin1 in platelets abrogates integrin activation, platelet
aggregation, and thrombus formation in vitro and in vivo. J. Exp. Med. (in the press).
7. Calderwood, D.A. et al. The phosphotyrosine binding-like domain of talin activates
integrins. J. Biol. Chem. 277, 21749–21758 (2002).
8. Calderwood, D.A. et al. The talin head domain binds to integrin bsubunit cytoplasmic
tails and regulates integrin activation. J. Biol. Chem. 274, 28071–28074 ( 1999).
9. Wegener, K.L. et al. Structural basis of integrin activation by talin. Cell 128, 171–182
(2007).
10. Kloeker, S. et al. The Kindler syndrome protein is regulated by transforming growth
factor-band involved in integrin-mediated adhesion. J. Biol. Chem. 279, 6824–6833
(2004).
11. Tu, Y., Wu, S., Shi, X., Chen, K. & Wu, C. Migfilin and Mig-2 link focal adhesions to
filamin and the actin cytoskeleton and function in cell shape modulation. Cell 113,
37–47 (2003).
12. Ussar, S., Wang, H.V., Linder, S., Fa
¨ssler, R. & Moser, M. The Kindlins: subcellular
localization and expression during murine development. Exp. Cell Res. 312,
3142–3151 (2006).
13. Weinstein, E.J. et al. URP1: a member of a novel family of PH and FERM domain-
containing membrane-associated proteins is significantly over-expressed in lung and
colon carcinomas. Biochim. Biophys. Acta 1637, 207–216 (2003).
14. Hodivala-Dilke, K.M. et al. b3-integrin-deficient mice are a model for Glanzmann
thrombasthenia showing placental defects and reduced survival. J. Clin. Invest. 103,
229–238 (1999).
15. Nieswandt, B. et al. Glycoprotein VI but not a2b1 integrin is essential for platelet
interaction with collagen. EMBO J. 20, 2120–2130 (2001).
16. Lenter, M. et al. A monoclonalantibody against an activation epitope on mouse integrin
chain b1 blocks adhesion of lymphocytes to the endothelial integrin a6beta1.Proc.
Natl. Acad. Sci. USA 90, 9051–9055 (1993).
17. Holtkotter, O. et al. Integrin a2-deficient mice develop normally, are fertile, but
display partially defective platelet interaction with collagen. J. Biol. Chem. 277,
10789–10794 (2002).
LETTERS
NATURE MEDICINE VOLUME 14
[
NUMBER 3
[
MARCH 2008 329
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
18. Pfaff, M., Liu, S., Erle, D.J. & Ginsberg, M.H. Integrin bcytoplasmic domains
differentially bind to cytoskeletal proteins. J. Biol. Chem. 273, 6104–6109
(1998).
19. Chen, Y.P. et al. Ser-752-Pro mutation in the cytoplasmic domain of integrin b3
subunit and defe ctive activation of platelet integrin aIIbb3 (glycoprotein IIb-IIIa) in a
variant of Glanzmann thrombasthenia. Proc. Natl. Acad. Sci. USA 89, 10169–10173
(1992).
20. McCarty, O.J. et al. Rac1 is essential for platelet lamellipodia formation and aggregate
stability under flow. J. Biol. Chem. 280, 39474–39484 (2005).
21. Savage, B., Shattil, S.J. & Ruggeri, Z.M. Modulation of platelet function through
adhesion receptors. A dual role for glycoprotein IIb-IIIa (integrin aIIbb3) mediated by
fibrinogen and glycoprotein Ib-von Willebrand factor. J. Biol. Chem. 267,
11300–11306 (1992).
22. Takahashi, S. et al. The RGD motif in fibronectin is essential for development but
dispensable for fibril assembly. J. Cell Biol. 178, 167–178 (2007).
23. Fa
¨ssler, R. et al. Lack of b1 integrin gene in embryonic stem cells affects morphology,
adhesion, and migration but not integration into the inner cell mass of blastocysts.
J. Cell Biol. 128, 979–988 (1995).
LETTERS
330 VOLUME 14
[
NUMBER 3
[
MARCH 2008 NATURE MEDICINE
©2008 Nature Publishing Group http://www.nature.com/naturemedicine
... It has already been shown that kindlin-3 interacts with and regulates β1-integrin activation through a direct interaction between its F3 domain and the integrin β subunit tail, including their distal NxxY motifs [47]. In addition previous studies demonstrated that DDR1 also regulates β1-integrin interactions with fibrillar collagen promoting β1-integrin mediated cell adhesion [48,49]. ...
Article
Full-text available
The role of the focal adhesion protein kindlin-3 as a tumor suppressor and its interaction mechanisms with extracellular matrix constitute a major field of investigation to better decipher tumor progression. Besides the well-described role of kindlin-3 in integrin activation, evidence regarding modulatory functions between melanoma cells and tumor microenvironment are lacking and data are needed to understand mechanisms driven by kindlin-3 inactivation. Here, we show that kindlin-3 inactivation through knockdown or somatic mutations increases BRAFV600mut melanoma cells oncogenic properties via collagen-related signaling by decreasing cell adhesion and enhancing proliferation and migration in vitro, and by promoting tumor growth in mice. Mechanistic analysis reveals that kindlin-3 interacts with the collagen-activated tyrosine kinase receptor DDR1 (Discoidin domain receptor 1) modulating its expression and its interaction with β1-integrin. Kindlin-3 knockdown or mutational inactivation disrupt DDR1/β1-integrin complex in vitro and in vivo and its loss improves the anti-proliferative effect of DDR1 inhibition. In agreement, kindlin-3 downregulation is associated with DDR1 over-expression in situ and linked to worse melanoma prognosis. Our study reveals a unique mechanism of action of kindlin-3 in the regulation of tumorigenesis mediated by the collagen-activated tyrosine kinase receptor DDR1 thus paving the way for innovative therapeutic targeting approaches in melanoma.
... Integrins serve as transmembrane linkers that facilitate interactions between the cytoskeleton and the extracellular matrix. Stimulation of cells recruits cytoplasmic factors to the cytoplasmic motifs of integrin's beta-chain, leading to integrin activation [47][48][49] . Recent research has demonstrated that the GPR motif on ECM1 binds to integrin αX and β2, as validated by proximity ligation and co-IP assays using an ECM1 mutant with a VAQ sequence substitution 32 . ...
Article
Full-text available
The cargo content in small extracellular vesicles (sEVs) changes under pathological conditions. Our data shows that in obesity, extracellular matrix protein 1 (ECM1) protein levels are significantly increased in circulating sEVs, which is dependent on integrin-β2. Knockdown of integrin-β2 does not affect cellular ECM1 protein levels but significantly reduces ECM1 protein levels in the sEVs released by these cells. In breast cancer (BC), overexpressing ECM1 increases matrix metalloproteinase 3 (MMP3) and S100A/B protein levels. Interestingly, sEVs purified from high-fat diet-induced obesity mice (D-sEVs) deliver more ECM1 protein to BC cells compared to sEVs from control diet-fed mice. Consequently, BC cells secrete more ECM1 protein, which promotes cancer cell invasion and migration. D-sEVs treatment also significantly enhances ECM1-mediated BC metastasis and growth in mouse models, as evidenced by the elevated tumor levels of MMP3 and S100A/B. Our study reveals a mechanism and suggests sEV-based strategies for treating obesity-associated BC.
... A large body of studies have revealed that the mechanism of platelet activation depends on the functional changes in integrin α IIb β 3 . 1 The structural characteristics of both the extracellular domain that mediates biological function [5][6][7] and the intracellular domain that induce functional changes [8][9][10][11] were deeply investigated. Recently, Tong et al revealed the importance of an intermediate structure between active and nonactive conformation for platelet adhesion by the use of molecular dynamic (MD) calculations. ...
Article
Full-text available
Background The structure and functions of the extracellular domain of platelet integrin αIIbβ3 (platelet membrane glycoprotein: GPIIb–IIIa) change substantially upon platelet activation. However, the stability of the integrated model of extracellular/transmembrane/intracellular domains of integrin αIIbβ3 with the inactive state of the extracellular domain has not been clarified. Methods The integrated model of integrin αIIbβ3 was developed by combining the extracellular domain adopted from the crystal structure and the transmembrane and intracellular domain obtained by Nuclear Magnetic Resonace (NMR). The transmembrane domain was settled into the phosphatidylcholine (2-oleoyl-1-palmitoyl-sn-glycerol-3-phosphocholine (POPC)) lipid bilayer model. The position coordinates and velocity vectors of all atoms and water molecules around them were calculated by molecular dynamic (MD) simulation with the use of Chemistry at Harvard Macromolecular Mechanics force field in every 2 × 10−15 seconds. Results The root-mean-square deviations (RMSDs) of atoms constructing the integrated αIIbβ3 model apparently stabilized at approximately 23 Å after 200 ns of calculation. However, minor fluctuation persisted during the entire calculation period of 650 ns. The RMSDs of both αIIb and β3 showed similar trends before 200 ns. The RMSD of β3 apparently stabilized approximately at 15 Å at 400 ns with persisting minor fluctuation afterward, while the structural fluctuation in αIIb persisted throughout the 650 ns calculation period. Conclusion In conclusion, the integrated model of the intracellular, transmembrane, and extracellular domain of integrin αIIbβ3 suggested persisting fluctuation even after convergence of MD calculation.
... Activation of αIIbβ3 integrin on the platelet surface strongly depends on the talin-linked protein complex, which is coupled to the cytoplasmic tail of β3 integrin. 28,29 This complex includes vinculin, kindlin-3, and ADAP and plays a crucial role in regulating the movement of the actomyosin complex in response to shear stress. The interaction between the αIIbβ3 integrin and talin complex is dynamic and determines the contractility of the actomyosin complex, thereby regulating cell adhesion, spreading, and migration on the surface of the ECM. ...
Article
Blood flow-induced hemodynamic changes result in mechanical stress on blood cells and vessel walls. Increased shear stress can activate platelets and other circulating cells, triggering the rapid activation of receptors, calcium channels, and related signaling mechanisms. Shear stress can also modify the folding of extracellular molecules and directly activate mechanosensitive receptors and calcium channels. The mechanical movement of the ECM (extracellular matrix) and the intracellular cortical actin cytoskeleton can change the conformation of platelet receptors and gate channel pores in the plasma membrane. Mechanosensitive platelet receptors and their downstream signaling events and metabolic products can also indirectly activate calcium channels. While the molecular composite of mechanotransduction pathways has been described in mammals, shear stress-induced platelet receptors and their cross talk with calcium channels have been incompletely characterized. In this review, we discuss current knowledge about the role of mechanosensitive platelet receptors and calcium channels in shear-dependent platelet activation and arterial thrombus formation.
Article
Full-text available
Extracellular vesicles (EVs) are nano-sized, membranous structures secreted into the extracellular space. They exhibit diverse sizes, contents, and surface markers and are ubiquitously released from cells under normal and pathological conditions. Human serum is a rich source of these EVs, though their isolation from serum proteins and non-EV lipid particles poses challenges. These vesicles transport various cellular components such as proteins, mRNAs, miRNAs, DNA, and lipids across distances, influencing numerous physiological and pathological events, including those within the tumor microenvironment (TME). Their pivotal roles in cellular communication make EVs promising candidates for therapeutic agents, drug delivery systems, and disease biomarkers. Especially in cancer diagnostics, EV detection can pave the way for early identification and offers potential as diagnostic biomarkers. Moreover, various EV subtypes are emerging as targeted drug delivery tools, highlighting their potential clinical significance. The need for non-invasive biomarkers to monitor biological processes for diagnostic and therapeutic purposes remains unfulfilled. Tapping into the unique composition of EVs could unlock advanced diagnostic and therapeutic avenues in the future. In this review, we discuss in detail the roles of EVs across various conditions, including cancers (encompassing head and neck, lung, gastric, breast, and hepatocellular carcinoma), neurodegenerative disorders, diabetes, viral infections, autoimmune and renal diseases, emphasizing the potential advancements in molecular diagnostics and drug delivery.
Article
Full-text available
Integrin affinity regulation, also termed integrin activation, is essential for metazoan life. Although talin and kindlin binding to the β-integrin cytoplasmic tail is indispensable for integrin activation, it is unknown how they achieve this function. By combining NMR, biochemistry and cell biology techniques, we found that talin and kindlin binding to the β-tail can induce a conformational change that increases talin affinity and decreases kindlin affinity toward it. We also discovered that this asymmetric affinity regulation is accompanied by a direct interaction between talin and kindlin, which promotes simultaneous binding of talin and kindlin to β-tails. Disrupting allosteric communication between the β-tail-binding sites of talin and kindlin or their direct interaction in cells severely compromised integrin functions. These data show how talin and kindlin cooperate to generate a small but critical population of ternary talin–β-integrin–kindlin complexes with high talin–integrin affinity and high dynamics.
Article
Full-text available
Cell adhesion and migration depend on the assembly and disassembly of adhesive structures known as focal adhesions. Cells adhere to the extracellular matrix (ECM) and form these structures via receptors, such as integrins and syndecans, which initiate signal transduction pathways that bridge the ECM to the cytoskeleton, thus governing adhesion and migration processes. Integrins bind to the ECM and soluble or cell surface ligands to form integrin adhesion complexes (IAC), whose composition depends on the cellular context and cell type. Proteomic analyses of these IACs led to the curation of the term adhesome, which is a complex molecular network containing hundreds of proteins involved in signaling, adhesion, and cell movement. One of the hallmarks of these IACs is to sense mechanical cues that arise due to ECM rigidity, as well as the tension exerted by cell-cell interactions, and transduce this force by modifying the actin cytoskeleton to regulate cell migration. Among the integrin/syndecan cell surface ligands, we have described Thy-1 (CD90), a GPI-anchored protein that possesses binding domains for each of these receptors and, upon engaging them, stimulates cell adhesion and migration. In this review, we examine what is currently known about adhesomes, revise how mechanical forces have changed our view on the regulation of cell migration, and, in this context, discuss how we have contributed to the understanding of signaling mechanisms that control cell adhesion and migration.
Article
Full-text available
The integrins and G protein-coupled receptors are both fundamental in cell biology. The cross talk between these two, however, is unclear. Here we show that β3 integrins negatively regulate G protein-coupled signaling by directly inhibiting the Gα13-p115RhoGEF interaction. Furthermore, whereas β3 deficiency or integrin antagonists inhibit integrin-dependent platelet aggregation and exocytosis (granule secretion), they enhance G protein-coupled RhoA activation and integrin-independent secretion. In contrast, a β3-derived Gα13-binding peptide or Gα13 knockout inhibits G protein-coupled RhoA activation and both integrin-independent and dependent platelet secretion without affecting primary platelet aggregation. In a mouse model of myocardial ischemia/reperfusion injury in vivo, the β3-derived Gα13-binding peptide inhibits platelet secretion of granule constituents, which exacerbates inflammation and ischemia/reperfusion injury. These data establish crucial integrin-G protein crosstalk, providing a rationale for therapeutic approaches that inhibit exocytosis in platelets and possibly other cells without adverse effects associated with loss of cell adhesion.
Article
Full-text available
We have found that the form of glycoprotein (GP) IIb-IIIa (integrin-alpha(IIb)beta-3) expressed on nonstimulated platelets is a functional receptor that mediates selective and irreversible adhesion to immobilized fibrinogen. This occurs even in the presence of the elevated intracellular cAMP levels induced by prostaglandin E1 or after inhibition of protein kinase C activity by sphingosine. In the absence of inhibitors, platelets adhering to fibrinogen through GP IIb-IIIa become fully activated and aggregate with one another. Immobilized von Willebrand factor (vWF), in contrast, is recognized by nonstimulated platelets through another receptor, GP Ib. This interaction leads to a change in the ligand recognition specificity of GP IIb-IIIa that can then bind to immobilized vWF and mediate irreversible platelet adhesion and aggregation; this process, however, is inhibited by elevated intracellular cAMP levels or blockade of protein kinase C activity. Therefore, GP Ib and GP IIb-IIIa induce platelet activation through the selective recognition of immobilized vWF and fibrinogen, respectively, in the absence of exogenous agonists. Moreover, "nonactivated" and "activated" GP IIb-IIIa exhibits distinctly different reactivity toward surface-bound vWF, and the functional switch can be induced by the binding of vWF to GP Ib. These findings demonstrate the modulation of platelet function by two different adhesion receptors, GP Ib and GP IIb-IIIa, as well as the distinct dual role of the latter as the necessary common mediator of irreversible adhesion and aggregation on both fibrinogen and vWF.
Article
Full-text available
C.Brakebusch and W.Bergmeier contributed equally to this work Platelet adhesion on and activation by components of the extracellular matrix are crucial to arrest post-traumatic bleeding, but can also harm tissue by occluding diseased vessels. Integrin a2b1 is thought to be essential for platelet adhesion to subendothelial collagens, facilitating subsequent interactions with the activating platelet collagen receptor, glycoprotein VI (GPVI). Here we show that Cre/loxP-mediated loss of b1 integrin on platelets has no signi®cant effect on the bleeding time in mice. Aggregation of b1-null platelets to native ®brillar collagen is delayed, but not reduced, whereas aggregation to enzymatically digested soluble collagen is abolished. Furthermore, b1-null platelets adhere to ®brillar, but not soluble collagen under static as well as low (150 s ±1) and high (1000 s ±1) shear ¯ow conditions, probably through binding of aIIbb3 to von Willebrand factor. On the other hand, we show that platelets lacking GPVI can not activate integrins and consequently fail to adhere to and aggregate on ®brillar as well as soluble collagen. These data show that GPVI plays the central role in platelet±collagen interactions by activating different adhesive receptors, including a2b1 integrin, which strengthens adhesion without being essential.
Article
Full-text available
Integrins are membrane receptors which mediate cell-cell or cell-matrix adhesion. Integrin αIIbβ3 (glycoprotein Ilb-IIIa) acts as a fibrinogen receptor of platelets and mediates platelet aggregation. Platelet activation is required for α IIbβ3 to shift from noncompetent to competent for binding soluble fibrinogen. The steps involved in this transition are poorly understood. We have studied a variant of Glanzmann thrombasthenia, a congenital bleeding disorder characterized by absence of platelet aggregation and fibrinogen binding. The patient's platelets did not bind fibrinogen after platelet activation by ADP or thrombin, though his platelets contained α IIbβ3. However, isolated α IIbβ3 was able to bind to an Arg-Gly-Asp-Ser affinity column, and binding of soluble fibrinogen to the patient's platelets could be triggered by modulators of α IIbβ3 conformation such as the Arg-Gly-Asp-Ser peptide and α-chymotrypsin. These data suggested that a functional Arg-Gly-Asp binding site was present within α IIbβ3 and that the patient's defect was not secondary to a blockade of α IIbβ3 in a noncompetent conformational state. This was evocative of a defect in the coupling between platelet activation and α IIbβ3 up-regulation. We therefore sequenced the cytoplasmic domain of β3, following polymerase chain reaction (PCR) on platelet RNA, and found a T → C mutation at nucleotide 2259, corresponding to a Ser-752 → Pro substitution. This mutation is likely to be responsible for the uncoupling of α IIbβ3 from cellular activation because (i) it is not a polymorphism, (ii) it is the only mutation in the entire α IIbβ3 sequence, and (iii) genetic analysis of the family showed that absence of the Pro-752 β3 allele was associated with the normal phenotype. Our data thus identify the C-terminal portion of the cytoplasmic domain of β3 as an intrinsic element in the coupling between α IIbβ3 and platelet activation.
Article
Full-text available
A gene trap-type targeting vector was designed to inactivate the beta 1 integrin gene in embryonic stem (ES) cells. Using this vector more than 50% of the ES cell clones acquired a disruption in the beta 1 integrin gene and a single clone was mutated in both alleles. The homozygous mutant did not produce beta 1 integrin mRNA or protein, while alpha 3, alpha 5, and alpha 6 integrin subunits were transcribed but not detectable on the cell surface. Heterozygous mutants showed reduced beta 1 expression and surface localization of alpha/beta 1 heterodimers. The alpha V subunit expression was not impaired on any of the mutants. Homozygous ES cell mutants lacked adhesiveness for laminin and fibronectin but not for vitronectin and showed a reduced association with a fibroblast feeder layer. Furthermore, they did not migrate towards chemoattractants in fibroblast medium. None of these functions were impaired in heterozygous mutants. Scanning electron microscopy revealed that homozygous cells showed fewer cell-cell junctions and had many microvilli not usually found on wild type and heterozygous cells. This profound change in cell shape is not associated with gross alterations in the expression and distribution of cytoskeletal components. Unexpectedly, microinjection into blastocysts demonstrated full integration of homozygous and heterozygous mutants into the inner cell mass. This will allow studies of the consequences of beta 1 integrin deficiency in several in vivo situations.
Article
Full-text available
We have generated a monoclonal antibody (mAb), 9EG7, against mouse endothelial cells that blocks adhesion of lymphocytes to endothelial cells. Sequencing of four tryptic peptides of the purified antigen revealed its identity with the integrin chain beta 1. The only beta 1 integrin that is known to mediate cell-cell adhesion is alpha 4 beta 1 (VLA-4). This is not the integrin that is functionally defined by the mAb 9EG7 on endothelial cells. First, alpha 4 is not present on the analyzed endothelial cells. Second, mAb 9EG7 does not block the cell-adhesion function of alpha 4 beta 1 on the nonactivated mouse lymphoma L1-2. Thus, the mAb 9EG7 can functionally distinguish between different beta 1 integrins and defines a beta 1 integrin other than alpha 4 beta 1 as a newly discovered cell-cell adhesion molecule. This integrin is most likely alpha 6 beta 1, since an antibody against the alpha 6 chain blocks lymphocyte adhesion to the same degree as the mAb 9EG7, the effect of both antibodies is not additive, and the alpha 6 chain is coprecipitated with beta 1 in 9EG7 immunoprecipitations. Surprisingly, activation of alpha 4 beta 1 on L1-2 cells with phorbol ester or Mn2+ allows blocking of alpha 4 beta 1-mediated adhesion of L1-2 cells to endothelial cells with mAb 9EG7. Furthermore, only the activated alpha 4 beta 1 heterodimer, but not the unactivated complex, is detectable with 9EG7 in immunoprecipitations and by flow cytometry. Thus, mAb 9EG7 defines an epitope on integrin chain beta 1, which is accessible on the alpha 4 beta 1 heterodimer only after activation of this integrin.
Article
Full-text available
beta3 integrins have been implicated in a wide variety of functions, including platelet aggregation and thrombosis (alphaIIbbeta3) and implantation, placentation, angiogenesis, bone remodeling, and tumor progression (alphavbeta3). The human bleeding disorder Glanzmann thrombasthenia (GT) can result from defects in the genes for either the alphaIIb or the beta3 subunit. In order to develop a mouse model of this disease and to further studies of hemostasis, thrombosis, and other suggested roles of beta3 integrins, we have generated a strain of beta3-null mice. The mice are viable and fertile, and show all the cardinal features of GT (defects in platelet aggregation and clot retraction, prolonged bleeding times, and cutaneous and gastrointestinal bleeding). Implantation appears to be unaffected, but placental defects do occur and lead to fetal mortality. Postnatal hemorrhage leads to anemia and reduced survival. These mice will allow analyses of the other suggested functions of beta3 integrins and we report that postnatal neovascularization of the retina appears to be beta3-integrin-independent, contrary to expectations from inhibition experiments.
Article
A gene trap-type targeting vector was designed to inactivate the beta 1 integrin gene in embryonic stem (ES) cells. Using this vector more than 50% of the ES cell clones acquired a disruption in the beta 1 integrin gene and a single clone was mutated in both alleles. The homozygous mutant did not produce beta 1 integrin mRNA or protein, while alpha 3, alpha 5, and alpha 6 integrin subunits were transcribed but not detectable on the cell surface. Heterozygous mutants showed reduced beta 1 expression and surface localization of alpha/beta 1 heterodimers. The alpha V subunit expression was not impaired on any of the mutants. Homozygous ES cell mutants lacked adhesiveness for laminin and fibronectin but not for vitronectin and showed a reduced association with a fibroblast feeder layer. Furthermore, they did not migrate towards chemoattractants in fibroblast medium. None of these functions were impaired in heterozygous mutants. Scanning electron microscopy revealed that homozygous cells showed fewer cell-cell junctions and had many microvilli not usually found on wild type and heterozygous cells. This profound change in cell shape is not associated with gross alterations in the expression and distribution of cytoskeletal components. Unexpectedly, microinjection into blastocysts demonstrated full integration of homozygous and heterozygous mutants into the inner cell mass. This will allow studies of the consequences of beta 1 integrin deficiency in several in vivo situations.
Article
We have found that the form of glycoprotein (GP) IIb-IIIa (integrin alpha IIb beta 3) expressed on nonstimulated platelets is a functional receptor that mediates selective and irreversible adhesion to immobilized fibrinogen. This occurs even in the presence of the elevated intracellular cAMP levels induced by prostaglandin E1 or after inhibition of protein kinase C activity by sphingosine. In the absence of inhibitors, platelets adhering to fibrinogen through GP IIb-IIIa become fully activated and aggregate with one another. Immobilized von Willebrand factor (vWF), in contrast, is recognized by nonstimulated platelets through another receptor, GP Ib. This interaction leads to a change in the ligand recognition specificity of GP IIb-IIIa that can then bind to immobilized vWF and mediate irreversible platelet adhesion and aggregation; this process, however, is inhibited by elevated intracellular cAMP levels or blockade of protein kinase C activity. Therefore, GP Ib and GP IIb-IIIa induce platelet activation through the selective recognition of immobilized vWF and fibrinogen, respectively, in the absence of exogenous agonists. Moreover, "nonactivated" and "activated" GP IIb-IIIa exhibits distinctly different reactivity toward surface-bound vWF, and the functional switch can be induced by the binding of vWF to GP Ib. These findings demonstrate the modulation of platelet function by two different adhesion receptors, GP Ib and GP IIb-IIIa, as well as the distinct dual role of the latter as the necessary common mediator of irreversible adhesion and aggregation on both fibrinogen and vWF.
Article
Integrin cytoplasmic domains connect these receptors to the cytoskeleton. Furthermore, integrin-cytoskeletal interactions involve ligand binding (occupancy) to the integrin extracellular domain and clustering of the integrin. To construct mimics of the cytoplasmic face of an occupied and clustered integrin, we fused the cytoplasmic domains of integrin beta subunits to an N-terminal sequence containing four heptad repeat sequences. The heptad repeats form coiled coil dimers in which the cytoplasmic domains are parallel dimerized and held in an appropriate vertical stagger. In these mimics we found 1) that both conformation and protein binding properties are altered by insertion of Gly spacers C-terminal to the heptad repeat sequences; 2) that the cytoskeletal proteins talin and filamin are among the polypeptides that bind to the integrin beta1A tail. Filamin, but not talin binding, is enhanced by the insertion of Gly spacers; 3) binding of both cytoskeletal proteins to beta1A is direct and specific, since it occurs with purified talin and filamin and is inhibited in a point mutant (beta1A(Y788A)) or in splice variants (beta1B, beta1C) known to disrupt cytoskeletal associations of beta1 integrins; 4) that the muscle-specific splice variant, beta1D, binds talin more tightly than beta1A and is therefore predicted to form more stable cytoskeletal associations; and 5) that the beta7 cytoplasmic domain binds filamin better than beta1A. The structural specificity of these associations suggests that these mimics offer a useful approach for the analysis of the interactions and structure of the integrin cytoplasmic face.
Article
We have used confocal videomicroscopy in real time to delineate the adhesive interactions supporting platelet thrombus formation on biologically relevant surfaces. Type I collagen fibrils exposed to flowing blood adsorb von Willebrand factor (vWF), to which platelets become initially tethered with continuous surface translocation mediated by the membrane glycoprotein Ib alpha. This step is essential at high wall shear rates to allow subsequent irreversible adhesion and thrombus growth mediated by the integrins alpha2beta1 and alpha(IIb)beta3. On subendothelial matrix, endogenous vWF and adsorbed plasma vWF synergistically initiate platelet recruitment, and alpha2beta1 remains key along with alpha(IIb)beta3 for normal thrombus development at all but low shear rates. Thus, hemodynamic forces and substrate characteristics define the platelet adhesion pathways leading to thrombogenesis.