Download 1 - Utrecht University Repository

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

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

Document related concepts

Cytosol wikipedia , lookup

Apoptosis wikipedia , lookup

Cell encapsulation wikipedia , lookup

Biochemical switches in the cell cycle wikipedia , lookup

Cytoplasmic streaming wikipedia , lookup

Cell cycle wikipedia , lookup

Cell membrane wikipedia , lookup

Cell culture wikipedia , lookup

Apoptosome wikipedia , lookup

Cell growth wikipedia , lookup

Mitosis wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Endomembrane system wikipedia , lookup

Cellular differentiation wikipedia , lookup

Amitosis wikipedia , lookup

Cytokinesis wikipedia , lookup

Paracrine signalling wikipedia , lookup

Extracellular matrix wikipedia , lookup

List of types of proteins wikipedia , lookup

Signal transduction wikipedia , lookup

Transcript
Regulation of integrin
activation and trafficking
by the β-cytoplasmic tail
Hanneke N Monsuur
November 2010 - March 2011
Department of Cell Biology, Netherlands Cancer Institute (NKI), Amsterdam
Student number: 3052664
Master:
Biology of Disease, University of Utrecht
Supervisors:
Coert Margadant (NKI)
Dr. Arnoud Sonnenberg (NKI)
Dr. Peter van der Sluijs (UMCU)
Index
Regulation of integrin activation and trafficking by the β-cytoplasmic tail.......................... 1
Index ................................................................................................................................................................... 2
1 Introduction ................................................................................................................................................. 3
1.1 The integrin family ........................................................................................................................ 3
1.2 Integrin activation and signaling ............................................................................................. 4
1.3 Integrin domains and β-tail sequences ............................................................................... 6
2 Integrin activation .................................................................................................................................. 7
2.1 Inside-out activation ..................................................................................................................... 7
2.1.1 Regulation of inside-out activation by talin and kindlins ...................................... 9
2.1.2 Regulation of inside-out activation by the NPxY motifs and the
serine/threonine-rich region...........................................................................................10
2.2 Outside-in signaling.....................................................................................................................12
2.2.1 Focal adhesions ....................................................................................................................13
2.2.2 Fibrillar adhesions...............................................................................................................14
2.2.3 Regulation of outside-in signaling by the NPxY motifs .........................................15
3 Integrin trafficking ...............................................................................................................................17
3.1 Integrin endocytosis ...................................................................................................................17
3.2 Anterograde integrin transport ..............................................................................................20
3.3 Regulation of integrin trafficking by the β-subunit cytoplasmic tails .....................20
3.4 Functional consequences of integrin trafficking ..............................................................21
4 Discussion ................................................................................................................................................23
5 Abbreviations .........................................................................................................................................27
6 Reference list ..........................................................................................................................................28
2
1 Introduction
Integrins are heterodimeric transmembrane receptors, which integrate the interior of
the cell with the cell environment. Integrins mediate cell adhesion to extracellular
matrix (ECM) proteins and to other cells. Integrins are connected to the cytoskeleton,
and regulate several signal transduction pathways. In this way, they control many
cellular processes, including adhesion, migration, proliferation, differentiation, survival,
and gene expression, which is important for development, hemostasis, and immune
surveillance, but also in pathological situations as cancer, thrombosis, and
inflammation.
1.1 The integrin family
Each integrin is formed by dimerization of an α- and a β-subunit. Currently 18 α- and
eight β-subunits are known in mammals, which form 24 specific integrins (reviewed by
Hynes, 2002). The integrin family comprises collagen receptors (β1+α1/2/10/11),
laminin receptors (β1+α3/6/7 and α6β4), and RGD receptors (αIIbβ3, α5β1, α8β1, and
the αv-containing integrins) (Fig 1). RGD receptors recognize Arg-Gly-Asp (RGD)containing ligands including vitronectin, fibronectin (Fn), the latency-associatedpeptide of transforming growth factor-β1/3 (LAP-TGF-β1/3) and osteopontin. Ligand
specificity is defined by the combination of an α- and a β-subunit. Integrin expression
and activation, together with the availability of ligand ultimately determines which
interactions with ligands occur in vivo (reviewed by van der Flier and Sonnenberg,
2001; Humphries et al., 2006). The integrin family members and their ligands are
depicted in table I.
Figure 1. The integrin family. 18 α- and 8
β-subunits form 24 integrins, which can be
distinguished
according
to
ligand
specificity or cell-type-specific expression
(from Hynes, 2002).
3
Integrin
α1β1
α2β1
α3β1
α4
α5β1
α6
β1
β7
β1
β4
α7β1
α8β1
α9β1
α10β1
α11β1
αV
β1
β3
β5
β6
β8
αIIbβ3
αLβ2
αMβ2
αXβ2
αDβ2
αEβ7
Ligand
ECM
Soluble
Col;
Col;
Ln;
Fn; Op
Fn; Op
Fn; Op
Ln
Ln
Ln
Fn; Op; Tn; Vn
(Nn; LAP-TGF-β1/3)
Op; Tn
Col; Ln
Col
Fn; Op; LAP-TGF-β1/3
Fn; Op; Tsp; Vn; Tn;
vWF; Fg; Fb
LAP-TGF- β1/3
Fn; Op; Vn; LAP-TGFβ-1/3
Fn; Op; LAP-TGF-β1/3
Fn; LAP-TGF-β1/3
Fn; TSP; Vn
vWF; Fg
Col
(Fn; Vn)
Fg, FX; iC3b
Fg; iC3b
Cell-cell
(E-Cadherin)
VCAM-1; MAdCAM-1
VCAM-1; MAdCAM-1
VCAM-1
PECAM-1
ICAM
ICAM
ICAM
VCAM-1; ICAM
E-Cadherin
Table I. The 24 mammalian integrins and their ligands. Col = collagen, E-Cadherin = Epithelial
Cadherin, Fb = fibrillin, Fg = fibrinogen, Fn = fibronectin, FX = factor X, ICAM = Inter-Cellular Adhesion
Molecule, iC3b = inactive complement factor 3b, LAP-TGF-β = Latency-Associated Peptide of
Transforming Growth Factor-β, Ln = laminin, MAdCAM-1 = Mucosal Addressin Cell Adhesion Molecule-1,
Nn = nephronectin, Op = osteopontin, PECAM-1 = Platelet Endothelial Cell Adhesion Molecule-1, TSP =
thrombospondin, Tn = tenascin, VCAM-1 = Vascular Cell Adhesion Molecule-1, Vn = vitronectin, vWF =
von Willebrand factor (from Humphries et al., 2006)
1.2 Integrin activation and signaling
Integrin binding to ligand occurs upon integrin activation, which is defined as the
conformational change that causes a shift from a low to a high affinity for ligands.
Integrin activation can be induced by cytoplasmic events; a process designated ‘insideout’ activation or ‘inside-out’ signaling (reviewed by Ginsberg et al., 1992). After ligand
binding, integrins mediate cytoskeletal reorganization and cell spreading over the
substratum, and several cellular signal transduction pathways are activated that affect
cell motility, proliferation, survival, and gene expression, which is called ‘outside-in’
signaling (Hynes, 1992; Ginsberg et al., 1992). In addition to high affinity, tight adhesion
also requires integrin clustering (high avidity).
Regulation of integrin activation by ‘inside-out’ signals is particularly important in nonadherent cells such as leukocytes and platelets. Whereas these cells have inactive
integrins by default, inside-out activation occurs in response to specific cues, e.g. during
hemorrhaging or inflammation. Upon endothelial injury, platelet integrins are activated,
after which the platelets adhere to the damaged vascular endothelium and stop the
hemorrhage. However, inappropriate integrin activation on platelets and subsequent
uncontrolled platelet adhesion results in a pathological situation called thrombosis.
Integrins on leukocytes and platelets adhere, in addition to immobilized ECM proteins,
to cellular receptors including ICAM (Inter-Cellular Adhesion Molecule), VCAM-1
(Vascular Cell Adhesion Molecule-1) and E-Cadherin, or to soluble ligands like
fibrinogen and von Willebrand Factor (vWF).
In contrast to non-adherent cells, most adherent cells need to be attached to the ECM at
all times in order to maintain tissue integrity, requiring continuously active integrins.
Integrins in adherent cells reside in focal adhesions (FAs) or hemidesmosomes (HDs),
which are large macro-molecular complexes containing clusters of active integrins
connected to the cytoskeleton (reviewed by Zamir and Geiger, 2001; van der Flier and
Sonnenberg, 2001; Legate and Fassler, 2009). The interaction with the cytoskeleton
increases integrin clustering, and thus adhesion strengthening (Fig. 2; Legate and
Fässler, 2009). Integrin connection to the cytoskeleton occurs via proteins that bind the
cytoplasmic tail of the -subunit (the -tail). Interestingly, the -subunit targets the
integrin into FAs, while the α-subunit seems to inhibit integrin localization into FAs
(reviewed by Burridge et al., 1996). HDs mediate stable adhesion of basal epithelial cells
to the underlying basement membrane. HDs are formed by binding of the integrin α6β4
and BP180 to laminin-332, which is intracellularly stabilized by connection to the
intermediate filament system through BP230 and plectin (reviewed by Margadant et al.,
2008). In FAs, integrins connect to actin filaments, and FAs contain many additional
proteins that either provide a structural link to the cytoskeleton (structural adaptors),
or binding sites for other proteins (scaffold adaptors). In contrast to growth factor
receptors, integrins lack intrinsic enzymatic activity. Therefore, they require the binding
of signaling proteins to initiate signal transduction pathways (catalytic adaptors), which
are also abundant in FAs (Legate and Fässler, 2009). Structural adaptors in FAs include
talin and tensin, scaffold adaptors include paxillin and vinculin, and catalytic adaptors
are focal adhesion kinase (FAK) and Src.
Figure 2. Talin connects integrins to the actin
cytoskeleton in FAs. Talin is both a structural- and
a scaffold adaptor between the integrin -tail and
the actin cytoskeleton. Talin interacts with F-actin
either directly or indirectly via vinculin (from
Legate and Fässler, 2009).
5
1.3 Integrin domains and β-tail sequences
Both the α- and the β-subunit contain an extracellular domain of ~800 amino acids, a
transmembrane domain of ~20 amino acids, and a cytoplasmic tail of 13-70 amino
acids, with the exception of β4 which has a cytoplasmic tail of 1088 amino acids
(reviewed by Moser et al., 2009). The αextracellular domains comprise a βpropeller or head-domain, a thigh-domain,
and two calf-domains. About half of the αsubunits contain an additional I-domain.
The β-subunits contain a β-I-domain (or βA-domain after von Willebrand A domain),
which is analogous to the α-I-domain, a
hybrid domain that links the β-I-domain to
the
PSI
(plexin/semaphorin/integrin)
domain, four EGF (epidermal growth factor)
repeats, and an intracellular tail. Ligands
bind either to a metal ion-dependent
adhesion site (MIDAS) in both I-domains, or
Figure 3. Structural organization of integrins to the MIDAS in the β-I-domain and the
(from Moser et al., 2009).
propeller domain of the α-subunit, if no α-Idomain is present (Fig 3).
Integrin -tails harbor a relatively conserved membrane-proximal (MP) NPxY motif and
a membrane-distal (MD) NPxY (in β1) or NxxY motif (in β3, β5, and β6) (Table II). NPxY
motifs form β-turns, which are recognized by proteins containing a phosphotyrosinebinding domain (PTB). The NPxY or NxxY motifs serve as docking sites for a multitude
of proteins including talin, kindlins, Src, Dok1/2, tensin, filamin, Numb, Disabled (Dab)1, Dab-2, Eps8, and the Ras-related proteins in brain (Rabs). Whereas Dab1/2, 3endonexin, Src, and kindlins bind the MD-NxxY motif, talin, Numb, Dok1/2, and tensin
bind the MP-NPxY motif. In this way, these motifs are involved in integrin activation,
internalization, and recruitment of integrins to FAs. In addition, the β-tail of most
integrins contains a serine/threonine-rich sequence, which has also been implicated in
integrin activation. Here, we review how the -tails control integrin functions, with
particular emphasis on the regulation of integrin activation and trafficking by the NPxY
and NxxY motifs in β1 and β3.
β1A: KLLMII HDRREFAKFEKEKMNAKWDTGE NPIYKSAVTTVV
β1D: KLLMII HDRREFAKFEKEKMNAKWDTGE NPIYKSPINNFK
NPKYEGK
NPNYGRKAGL
β2:
KALIHL SDLREYRRFEKEKLKSQWNND. NPLFKSATTTVM
NPKFAES
β3:
KLLITI HDRKEFAKFEEERARAKWDTAN NPLYKEATSTFT
NITYRGT
β5:
KLLVTI HDRREFAKFQSERSRARYEMAS NPLYRKPISTHTVDFTFNKF NKSYNGTVD
β6:
KLLVSF HDRKEVAKFEAERSKAKWQTGT NPLYRGSTSTFK
NVTYKHREK QKVDLSTDC
β7:
RLSVEI YDRREYSRFEKEQQQLNWKQDS NPLYKSAITTTI
NPRFQEADSPTL
β8:
RQVILQ WNSNKIKSSSDYRVSASKKDKLILQSVCTRAVTYRREKPEEIKMDISKLNAHETFRCNF
Table II. Amino acid sequences of the cytoplasmic tails of all integrin β-subunits apart from β4. The
MP-NPxY and MD-NxxY motifs are highlighted in bold, and the serine/threonine-rich region is highlighted
in red and bold.
6
2 Integrin activation
Integrins on non-adherent cells exist on the plasma membrane in a low-affinity state, an
intermediate-affinity state, or a high-affinity state for ligand. The different states are in a
dynamic equilibrium, and the conformational change that mediates the transition from
the low-affinity state to the intermediate-/high-affinity state is called ‘integrin
activation’. Integrin activation can be induced by inside-out signals. After ligand binding,
outside-in signals are generated, which regulate a variety of cellular processes.
2.1 Inside-out activation
Integrin activation by inside-out mechanisms has been studied mostly in circulating
leukocytes and platelets, which should not adhere under normal physiological
circumstances, and therefore have inactive integrins by default. The main integrin on
platelets is αIIbβ3, and a wealth of evidence indicates that αIIbβ3 is regulated tightly by
conformational changes induced by ‘inside-out’ signals. αIIbβ3 is considered an
archetype for all integrins, and it is often assumed that the mechanisms of activation of
other integrins are similar to those of αIIbβ3. Whereas a bent form with a closed head is
thought to represent the low-affinity state, an extended form with a closed head
represents the intermediate state, and the extended form with an open head represents
activated- and ligand-bound integrins (Fig. 4; Xiong et al., 2001; Nishida et al., 2006;
Takagi et al., 2002). Stabilization of the low-affinity conformation is thought to occur by
interactions between the α- and β-subunit (sometimes referred to as a ‘clasp’),
particularly a salt bridge between R995 in αIIb and D723 in β3 (Vinogradova). The
disruption of this salt bridge leads to separation of the cytoplasmic tails or the
‘unclasping’ of the αIIb and the β3 subunits, and is thus important in the transition of
the low-/intermediate- to the high-affinity state (Nishida et al., 2006; Takagi et al.,
2002).
Figure 4. Structural rearrangements occurring in integrin activation. The α-headpiece domains are
displayed in red, the α-tailpiece in pink, the β-headpiece domains are displayed in blue, and the βtailpiece in cyan. The transmembrane and cytoplasmic domains are represented in black (adapted from
Takagi et al., 2002).
Inside-out activation of αIIbβ3 is initiated by ligands for G-protein coupled receptors
(GPCRs) including thrombin, sphingosine-1-phosphate, lysophosphatidic acid (LPA),
7
and phorbol myristate acetate. Ligand binding to a GPCR leads to an increase of
intracellular Ca2+ and formation of diacylglycerol (DAG). This results in activation of
protein kinase C (PKC) and a Rap guanine nucleotide exchanger (Rap-GEF), which then
cooperate to activate Rap1. Activated Rap1 interacts with RIAM, after which talin
associates with RIAM to form a complex inducing integrin activation (Fig 5; Han et al.,
2006).
Figure 5. Inside-out integrin
activation. Ligand binding to a GPCR
results in increased intracellular Ca2+
and DAG. PKC and a Rap-GEF are
then activated, which induce Rap1
activation. RIAM associates with
talin upon interaction with Rap1,
resulting in integrin activation due to
the unmasking of an integrin-binding
site in talin (from Han et al, 2006).
Inside-out activation of IIb3 and other integrins on non-adherent cells is of vital
importance, as illustrated by the human disorders Glanzmann thrombasthenia (GT) and
leukocyte adhesion deficiency syndrome-I and -III (LAD-I,-III). GT is a rare bleeding
disorder caused by reduced binding of αIIbβ3 to fibrinogen or vWF, which leads to
defective platelet adhesion and aggregation during a bleeding. The patients suffer from
mild bruising to severe hemorrhages. GT is caused by mutations in either the α- or βsubunit of αIIbβ3. Many different mutations have been found across the entire sequence
of the α- or the β-subunit, leading to quantitative or qualitative defects (Nurden et al.,
2006). Whereas many mutations lead to reduced surface expression of αIIbβ3, probably
by compromising the transport of integrins to the platelet surface, other mutations
affect the ability to bind to ligands. Some mutations affect surface expression, as well as
activation and subsequent outside-in signaling (e.g. the S752P mutation in the β3subunit; Chen et al., 1994). LAD-I results from defects in functions of β2 integrins on
leukocytes due to mutations in the gene encoding the β2-subunit. As a consequence,
LAD-I patients suffer from recurrent bacterial infections and impaired wound healing.
LAD-III is caused by defects in the functions of β1, β2, and β3 integrins on leukocytes
and platelets, leading to LAD-I-like infections and additional GT-like bleeding
abnormalities.
Although agonist-induced activation of αIIbβ3 has been studied mostly in adherent
Chinese Hamster ovary (CHO) cells, which express talin and PKC at levels comparable
to those in platelets, it is unclear whether the same inside-out mechanisms exist to
activate endogenous integrins in adherent cells, and even whether the same
conformational states exist for integrins in adherent cells. A recent study suggests that
α5β1 in adherent cells does adopt either the bent, inactive state or the extended state
with separated legs, and that the conversion from the former to the latter occurs in FAs
(Askari et al., 2010). Integrin activation in adherent cells is probably induced by
cytoskeletal tension generated by myosin-II, in response to ECM stiffness. Depending on
the degree of tension, binding of α5β1 to Fn can occur in two ways. Under low tension,
51 forms ‘relaxed’ bonds solely to the RGD site of Fn, whereas under high tension,
51 binds both the RGD and additional sites on Fn, leading to stronger, ‘tensioned’
bonds (Friedland et al., 2009).
8
2.1.1 Regulation of inside-out activation by talin and kindlins
Inside-out activation of integrins occurs by talin binding to the MP-NPxY motif. Talin
consists of an N-terminal head domain and a C-terminal rod or tail. The head is
constituted by a FERM (4.1, ezrin, radixin, moesin) domain, of which the F3 domain
harbors the PTB-site (Fig 6; Moser et al., 2009). The PTB-site recognizes the MP-NPxY
motif in β1, β2, β3, β5, and β7 (Pfaff et al., 1998, Calderwood et al., 2003). Subsequently,
the talin-head binds additional membrane-proximal residues. Talin exists in the
cytoplasm in an auto-inhibitory conformation, in which the C-terminal domain interacts
with the PTB-domain. Talin recruitment to the membrane and disruption of the autoinhibitory conformation are mediated by phosphatidylinositol-4,5-biphosphate
(PI4,5P2), generated by phosphatidylinositol phosphate kinase (PIPK1; Martel et al.,
2001, Goksoy et al., 2008). Interestingly, talin can in turn bind to and activate PIPK,
generating a positive feedback loop. Moreover, PI4,5P2 is important for the recruitment
of additional proteins important for FA assembly (Ling et al., 2003).
Figure 6. Structural organization
of talin and kindlin (from Moser et
al., 2009).
Talin binding disrupts the ‘clasp’ between the α- and the β-tail, and requires the D723
residue in β3 (Vinogradova et al., 2002; Anthis et al., 2009). Both the binding of talin
and the disruption of the salt bridge are required for activation of αIIbβ3 (Anthis et al.,
2009; Tadokoro et al., 2003; Wegener et al., 2007; Lu et al., 2001; Takagi et al., 2001).
The salt bridge appears to have no apparent function in β1-integrins under normal
physiological conditions in vivo, as mice carrying a D759A substitution in β1, which
impairs formation of the salt bridge, had no obvious defects (Czuchra et al., 2006).
Although binding of the talin-head alone is sufficient for integrin activation, the talinrod is required to connect integrins to the actin cytoskeleton, and thus to target
integrins to FAs (Calderwood et al., 2003; Tanentzapf and Brown, 2006; Bouaouina et
al., 2008; Zhang et al., 2008). The importance of talin for integrin function is underlined
by in vivo studies; disruption of the talin gene in mice leads to death during gastrulation,
due to a failure of integrin-mediated cell migration (Monkley et al., 2000). In addition,
platelet-restricted deletion of talin in mice impairs integrin α2β1-mediated adhesion to
collagen and αIIbβ3-mediated adhesion to Fn, and causes severe hemostatic defects
(Nieswandt et al., 2007; Petrich et al., 2007). It should be noted that in Drosophila
melanogaster, loss of talin expression results in wing blistering, much like the integrinknockout phenotype, but integrin binding to ECM is not impaired (Brown et al., 2002). A
similar phenotype is observed in Drosophila mutants expressing talin-head alone
(Tanentzapf and Brown, 2006). These observations suggest that talin-requirement for
inside-out activation in adherent cells can be bypassed. Indeed, integrin activation can
also be induced by ligand from the outside (Du et al., 1991). However, integrin-ligand
binding crucially depends on cytoskeletal association, as in the absence of the talin-rod,
loss of integrin function is observed.
9
Although one study suggests that the MP-NPxY in β3 is not essential for talin binding
and that the actual binding site for talin is located more proximal to the membrane,
most studies find that the NPxY motif is necessary for talin binding, and that talin
binding is abolished by Y-A substitution in the MP-NPxY motif (Patil et al., 1999;
Calderwood et al., 2002; Calderwood et al., 2003; García-Alvarez et al., 2003; Bouaouina
et al., 2008). Downstream of the MP-NPxY, a W-A substitution also impairs talin binding
(Bouaouina et al., 2008). Interestingly, talin binding was not prevented upon a Y-S
substitution in the MP-NPxY of β1, which is probably because this mutation still allows
the formation of a β-turn (Vignoud et al., 1997).
Other proteins that have been implicated in integrin activation are kindlin-1, -2 and -3.
Kindlins contain a C-terminal FERM domain, which is highly homologous to the FERM of
talin, but binds the MD-NxxY motif in β-tails (Meves et al., 2009). Kindlin-1 is expressed
in epithelial cells, kindlin-2 is ubiquitously expressed, and kindlin-3 is primarily present
in hematopoietic cells (Ussar et al., 2006). Mutations in kindlin-1 lead to Kindler
Syndrome (KS), which is a rare autosomal, recessive genodermatosis. Hallmarks of KS
are skin blistering, photosensitivity, poikiloderma, abnormal pigmentation and thinning
of the skin (Siegel et al., 2003). Histologically, KS is characterized by duplication of the
basement membrane and microblistering at the dermal-epidermal junction. The clinical
and histological features of KS resemble the defects caused by loss of β1 integrins from
the epidermis, in particular of α3β1 (Margadant et al., 2009; Raghavan et al., 2000;
Brakebusch et al., 2000). Moreover, mutations in kindlin-3 have been disclosed in LADIII patients, illustrating the importance of kindlin-3 in regulating integrin functions in
leukocytes and platelets. Genetic disruption of kindlin-3 recapitulates the LAD- and GTlike symptoms in mice, whereas the disruption of kindlin-1 partially resembles KS
(Moser et al., 2008; Ussar et al., 2008). Kindlin-2 knockout mice die during development
due to defective adhesion of the endoderm and epiblast to the basement membrane
(Montanez et al., 2008). Thus, it is clear that the kindlins are crucial for integrin
functions, however, their precise role is unclear and controversial. A number of studies
suggest that talin and kindlin cooperate to regulate inside-out integrin activation
(Moser et al., 2009). In many cell lines, expression of talin alone is sufficient to induce
integrin activation, but activation is enhanced when kindlin is co-expressed, whereas
kindlin alone only very weakly activates integrins (Ma et al., 2008; Bledzka et al., 2010;
Ye et al., 2010; Montanez et al., 2008). An intriguing possibility is that the F3 domain of
kindlin binds directly to the talin rod, thereby releasing the auto-inhibitory
conformation and thus activating talin. In contrast, other data suggest that expression of
kindlin-1 or -2 alone inhibits the activation of both αIIbβ3 and α5β1, whereas coexpression with talin results in activation of αIIbβ3 but inhibition of α5β1, suggesting
that kindlins may exert integrin-specific effects (Harburger et al., 2009). In summary,
whereas talin seems absolutely required for integrin activation, the effects of kindlins
on integrin functions are less clear. It is suggested that kindlins can indirectly favor talin
binding over the inhibitory binding of filamin, that kindlins stabilize integrins in FAs, or
that kindlins function as adaptors for other proteins (Ye et al., 2010; Harburger et al.,
2009).
2.1.2 Regulation of inside-out activation by the NPxY motifs and the
serine/threonine-rich region
It is controversial whether tyrosine–phoshorylation of the NPxY motifs is required for
inside-out activation. The role of the NPxY motifs in integrin activation has been
10
WT:
KLLMII HDRREFAKFEKEKMNAKWDTGE
D759A:
HARREFAKFEKEKMNAKWDTGE
Y783F:
Y783A:
Y795F:
Y795A:
YY783,795FF:
YY783,795AA:
β1A
β1A
β1A
β1A
β1A
β1A
β1A
β1A
β3
β3
β3
β3
WT:
KLLITI HDRKEFAKFEEERARAKWDTAN
Y747A:
Y759A:
YY747,759FF:
MP-NPxY
NPIYKSAVTTVV
MD-NxxY
NPKYEGK
NPIF
NPIA
NPIY
NPIY
NPIF
NPIA
NPKY
NPKY
NPKF
NPKA
NPKF
NPKA
MP-NPxY
NPLYKEATSTFT
NPLA
NPLY
NPLF
MD-NxxY
NITYRGT
NITY
NITA
NITF
Table III. Amino acid sequences of β1A and β3 cytoplasmic tails. The positions of Y-F and Y-A
mutations are indicated. The NPxY and NxxY motifs are highlighted in bold, and the serine/threonine-rich
region is highlighted in red and bold.
addressed using either chimeric integrins overexpressed in CHO cells, or mutated β1tails expressed in β1-deficient cells (Table III).
Results obtained in CHO cells suggest that structural integrity, rather than the
phosphorylation status of the MP-NPxY motif is required for integrin activation, because
conservative Y-F substitutions in the MP-NPxY of β1 and β3 do not affect integrin
activation or cell adhesion, whereas more disruptive Y-A substitutions do (O’Toole et al.,
1995; Romzek et al., 1998). Y-A substitutions in the MD-NPxY motifs also slightly
reduced integrin activation (O’Toole et al., 1995). Also in fibroblasts, a Y-F substitution
in the MD-NPxY motif did not affect adhesion, although there was a loss of directed
migration (Sakai et al., 1998). In contrast, Y-F substitutions in both NPxY motifs
impaired adhesion of T-lymphoma cells (Stroeken et al., 2000). This probably reflects
differences in the regulation of integrin activation between adherent and non-adherent
cells. However, mice carrying Y-F substitutions in both NPxY motifs (YY/FF) do not
display any defects (Czuchra et al., 2006). In contrast, YY/AA mice are not viable, and
keratinocyte-restricted YY/AA substitutions cause epidermal defects similar to the loss
of β1 expression in the epidermis (Czuchra et al., 2006). The complete loss of β1
function upon YY/AA mutations in mice was confirmed in another study, while YY/FF
substitutions had no effects (Chen et al., 2006). Thus, most in vitro and in vivo evidence
supports the concept that structural integrity of the NPxY motifs is required for integrin
activation, but tyrosine-phosphorylation is not. Instead, it has even been suggested that
phosphorylation may act as a negative regulator of integrin activation, either by
reducing the affinity of talin, or by increasing the affinity of other PTB-domaincontaining proteins such as Dok1, which then compete with talin for the binding site
(Czuchra et al., 2006; Oxley et al., 2007; McCleverty et al., 2007; Wegener et al., 2007). In
this way, phosphorylation may serve as a molecular switch. Dok1 binds β3 but not to
β1–tails (Calderwood et al., 2003). The differential effects of Dok1 and talin in integrin
activation are probably explained by differences in the structural properties of their
PTB-domains (Wegener et al., 2007). Tyrosine-phosphorylation of the NPxY motifs in
1 can occur by Src, and reduces adhesion to Fn and laminin (Sakai et al, 2001).
Y-A substitutions in the NPxY motifs impair the binding of a number of proteins
including Numb, Dab1, Dab2, EPS8, tensin, Dok1 and talin (Calderwood et al., 2003).
Although the NPxY motifs are highly conserved among β1A, β2, β3, β5 and β7, there is a
11
difference in which interactions occur, since not all proteins were able to bind to all βtails. Additional residues (-5 relative to the N, and +2 relative to the Y in the MP-NPxY
motif) influence the specificity of the interactions, for example both tensin and talin
require W776 downstream of the MP-NPxY motif (McCleverty et al., 2007; GarcíaAlvarez et al., 2003). Whereas talin initially binds the MP-NPxY motif, additional binding
to W776 is thought to result in the actual separation of the α- and the β-tail, following a
two-step activation pathway (García-Alvarez et al., 2003).
The serine/threonine region in between the two NPxY motifs is also considered
important for the binding of regulatory proteins and integrin activation. The region is
phosphorylated in β1, β2, β3 and β7 after agonist stimulation (Hilden et al., 2003;
Valmu and Gahmberg, 1995; Willigen et al., 1996; Nilsson et al., 2006). Phosphorylation
can occur by PKC and seems to regulate the exposure of the ligand-binding sites in
αIIbβ3 (Willigen et al., 1996). Conservative and non-conservative mutations of the
serines or threonines in β3 or β1 decreased integrin activation (O’Toole et al., 1995).
Consequently, an S752P substitution in αIIbβ3 resulted in impaired adhesion to Fn, and
T788A or T789A mutations in β1 decreased cell adhesion and invasion in T-lymphoma
cells (Ylänne et al., 1995; Stroeken et al., 2000). In spite of these observations however,
the β1 splice variant β1D lacks the serine/threonine stretch, and binds much stronger
to talin than β1A, suggesting that the serine/threonine-region negatively impacts talinintegrin binding (Belkin et al., 1997).
In summary, most studies indicate that the MP-NPxY motif is crucial, whereas the MDNPxY motif only subtly affects integrin activation. Morover, it seems that
phosphorylation of the NPxY motifs is not essential for activation, but instead may
regulate inactivation.
2.2 Outside-in signaling
Upon ligand binding, integrins cluster together and actin filaments are recruited, which
mediates cell spreading over the substrate, assembly of cell-matrix adhesions, and cell
migration. Collectively, these events are referred to as outside-in signaling. Three
different types of cell-matrix adhesions are distinguished, being focal complexes (FCs),
FAs, and fibrillar adhesions (Table IV). These adhesions differ in appearance and
molecular composition, and the differences are best-exemplified in the context of the
Fn-binding integrins αvβ3 and α5β1. FCs are small adhesions found at the cell
periphery, generally containing αvβ3, talin, vinculin and paxillin. FAs are also located at
the cell periphery and also contain αvβ3, vinculin, paxillin and talin, as well as many
proteins that are highly phosphorylated on tyrosines. By contrast, fibrillar adhesions
are found at more central positions and hardly contain tyrosine-phosphorylated
proteins. Fibrillar adhesions are associated with Fn fibrils, and contain α5β1 and tensin.
The three adhesion complexes probably represent different stages of development; FCs
being early adhesions, which over time develop into FAs, whereas fibrillar adhesions
develop as a result of Fn-associated α5β1 displacement from FAs (reviewed by Geiger et
al; 2001).
12
Table IV. Characteristic features of different types of cell-matrix adhesions (from Geiger et al., 2001).
2.2.1 Focal adhesions
Integrin connection to the cytoskeleton can occur in multiple ways and involves a
number of proteins, including talin, α-actinin, vinculin, FAK, paxillin, the IPP complex,
kindlin-2, and probably also kindlin-1 (Fig 7). Talin links integrins to actin filaments
either directly or via additional proteins like vinculin and α-actinin. The latter crosslinks
actin filaments to provide stiffness to actin filament networks. Vinculin interacts with
talin, α-actinin and the actin cytoskeleton, and recruits additional proteins like paxillin
and vinexin (Ziegler et al., 2006, 2008). Vinculin is believed to regulate FA dynamics, as
increased levels of phosphoinositides inhibit the interaction between vinculin and actin,
which promotes FA turnover and cell motility (Ziegler et al., 2006). In addition, vinculin
controls the interaction between paxillin and FAK, and FAK is also believed to regulate
FA turnover (Subauste et al., 2004). The IPP complex is formed in the cytosol by binding
of integrin-linked kinase (ILK) to the adaptor proteins PINCH and parvin. Recruitment
of the IPP complex to FAs requires all three components (Legate et al., 2006; Zhang et
al., 2002; Sepulveda et al., 2005). Paxillin can bind both ILK and parvin and kindlin-2
can bind ILK (Nikolopoulos and Turner, 2000, 2001; Montanez et al., 2008). Both
paxillin and kindlin-2 are required for localization of the IPP complex to FAs, since an
ILK mutant that does not interact with paxillin does not localize in FAs, and FA-like
structures in the absence of kindlin-2 do not contain ILK (Nikolopoulos and Turner,
2001; Montanez et al., 2008). The IPP complex binds to β1 and β3 tails, and the actin
cytoskeleton and regulates several signaling pathways (Legate et al., 2006). Kindlin-2
links integrins to the actin cytoskeleton via migfilin and filamin (Tu et al., 2003). It is
unclear whether this interaction is important, as the targeted deletion of migfilin in
mice does not cause any abnormalities (Moik et al., 2011). Whether kindlin-1 also
connects integrins to the cytoskeleton remains to be determined.
13
Figure 7. Integrin activation and ligand binding leads to assembly of FAs. An inactive integrin (A) is
activated by talin (B) and then binds to ECM (C) Subsequently, integrin clustering into FAs and
connection to the actin cytoskeleton occurs via talin, paxillin, vinculin, FAK, and the IPP complex (from
Legate et al., 2006).
2.2.2 Fibrillar adhesions
The high phosphorylation degree of FAs -in contrast to fibrillar adhesions- suggests that
FAs are the major sites of signaling, whereas fibrillar adhesions probably provide the
cells with a firm anchorage (Zamir, et al., 2000). αvβ3 forms a tight bond with
vitronectin, whereas α5β1 forms a tight bond with Fn. It is thought that fibrillar
adhesions arise from the specific translocation of α5β1 and tensin from FAs, which
starts at the margins of the FA towards the cell centre. Of all integrins, α5β1 is most
efficient in binding to soluble Fn dimers and supporting a fibroblast-like, contractile cell
shape by driving cytoskeletal organization (Danen et al., 2002). The binding of α5β1 to
soluble Fn dimers results in a reorientation of FAs, increased Rho activity, and
subsequent assembly of Fn dimers into fibrils, a process designated Fn fibrillogenesis
(Huveneers et al., 2008; Pankov et al., 2000). It is suggested that actomyosin-driven
contraction drives the formation of fibrillar adhesions. αvβ3 integrins bind to
immobilized vitronectin or Fn and are not much affected by contraction forces from the
cytoskeleton. By contrast, α5β1 binds soluble Fn, and is translocated out of FAs because
of low tensional forces in a tensin-dependent manner (Fig 8; Zamir et al., 2000; Pankov
et al., 2000). Fn fibrillogenesis is important for structural support from the ECM, and
probably for ‘sensing’ the environment, or (Clark et al., 2005).
14
Figure 8. Differential distribution
of FN-binding integrins αvβ3 and
α5β1. (A) FAs contain mainly αvβ3
and additional proteins as vinculin
and paxillin. Only small amounts of
α5β1 and tensin are present. B)
Actomyosin-driven
contraction
forces drive the formation of
fibrillar adhesions. Whereas αvβ3
stays in place due to its association
with stretched vitronectin or Fn
(providing high tension), α5β1
binding to soluble Fn (low tension)
allows its translocation. While
tensin co-translocates with α5β1,
vinculin and paxillin stay in FAs
(from Zamir et al., 2000)
2.2.3 Regulation of outside-in signaling by the NPxY motifs
As discussed in the previous chapter, inside-out activation of integrins is regulated by
the cytoplasmic sequences of the β-subunit. Downstream of integrin activation and
ligand binding, the same sequences are likely to also mediate outside-in signaling
events. Talin-binding to the MP-NpxY motif promotes cell spreading either directly,
because of its association with actin filaments, or indirectly by preventing the binding of
filamin. Filamin, in addition to inhibiting integrin activation, also inhibits cell spreading
by recruiting FilGAP, which prevents the activation of Rac required for cell spreading
(Kiema et al., 2006; Nieves et al., 2010). Because talin-binding is essential for both
integrin activation and cell spreading, a recent study made use of constitutively active
chimeric integrins (containing a deletion of the GFFKR sequence in the α-subunit) to
address the relative contribution of talin to outside-in signaling, independent of its
effects on integrin activation (Nieves et al., 2010). Interestingly, prevention of talinbinding by a Y-A substitution in the MP-NPxY motif did not impair cell spreading, FA
assembly or stress fiber formation in constitutively active integrins, implying that talinbinding is not absolutely required for outside-in signaling and that other proteins that
do not bind to the NPxY motif can also regulate these processes. However, cell
spreading was impaired when talin was depleted in cells expressing constitutively
active integrins without Y-A mutations in the β-tail, but not in the presence of these
mutations (Nieves et al., 2010). This inhibition could be relieved by depletion of filamin,
suggesting that talin binding to the NPxY motif is important downstream of integrin
activation to prevent the binding of filamin (Nieves et al., 2010). Some difference is
observed between different integrins in the ability to bind talin or filamin, providing
differential regulation of integrin activation, cell spreading and FA formation by talin
15
and filamin. How the differential binding of talin and filamin to integrins is regulated is
not known yet (Nieves et al., 2010). The kindlins are also thought to mediate outside-in
signaling events, as kindlin-3-deficient platelets are unable to spread even in the
presence of Mn2+ to induce integrin activation (Moser et al., 2008). Similarly, Mn2+
stimulation of kindlin-2-deficient cells does not completely rescue cell spreading or FA
assembly, whereas the few FAs that are formed contain paxillin but not ILK, suggesting
that kindlin-1 is required for recruitment of the IPP complex (Montanez et al., 2008).
Thus, it seems that in addition to integrin activation, the NPxY motifs are also important
for outside-in signaling. The function of tyrosine-phosphorylation is again uncertain. In
platelets, tyrosine-phosphorylation of both NPxY motifs is observed during platelet
aggregation, secondary to agonist-induced αIIbβ3 activation and fibrinogen binding,
and is thus an outside-in effect (Law et al., 1999). Platelets with Y-F substitutions in
both NxxY motifs of β3 were defective in aggregation and clot-retraction in vitro, and in
mice, the same mutations cause a mild bleeding defect (Law et al., 1999). These
observations suggest that tyrosine-phosphorylation of both NPxY motifs is required for
αIIbβ3 functions. In contrast, phosphorylation of the MD-tyrosine disrupts kindlin-2
binding to the β3-tail, suggesting that phosphorylation rather negatively impacts β3
functions (Bledzka et al., 2010). Furthermore, phosphorylation of the MP-NPxY in β1
impaired adhesion, assembly of FCs, and migration, whereas phosphorylation of the
MD-NxxY motif was important for optimal migration (Sakai et al., 1998; Sakai et al.,
2001). To reconcile conflicting results, it has been suggested that the dynamic
regulation of phosphorylation and dephosphorylation is important, for example for
integrin cycling in and out FAs (Sakai et al., 1998, 2001). In support of this is the
observation that a Y-S mutation in either NPxY motif impairs integrin recruitment to
FAs (Vignoud et al., 1997). Phosphorylation of the MP-NPxY motif occurs during the
maturation of FAs, and promotes the binding of tensin (McCleverty et al., 2007).
Phosphorylation does not increase the affinity of tensin, but promotes its binding
probably indirectly, due to a decrease of talin affinity. Tensin recruits various kinases
and RhoGAPs to FAs and thereby contributes to integrin signaling (Legate and Fässler,
2009). Because tensin is important for the translocation of α5β1 into fibrillar adhesions
and for Fn fibrillogenesis, it has been suggested that phosphorylation of the MP-tyrosine
in β1 is involved in Fn fibrillogenesis. However, a Y-F substitution in the MP-NxxY motif
of β1 caused increased Fn matrix assembly, and it therefore seems that the
phosphorylation of this motif negatively regulates Fn fibrillogenesis (Sakai et al., 2001).
Thus, in vitro reports on the importance of tyrosine-phosphorylations of the NPxY
motifs of β1 are controversial. Perhaps the best evidence that tyrosinephosphorylations are not important for β1 functions stems from mice carrying YY/FF
mutations in both NPxY motifs of β1. As discussed in the previous chapter, these mice
are completely normal, which does not support a role for tyrosine-phosphorylation
either in integrin activation, or in outside-in signaling (Czuchra et al., 2006; Chen et al.,
2006).
In summary, both NPxY motifs are required for integrin-mediated events downstream
of activation, and the MP-NPxY motif seems most important. The importance of
tyrosine-phosphorylations of the NPxY motifs remains controversial, and probably
differs between integrins; whereas for αIIbβ3 tyrosine-phosphorylation seems
important, this is not the case for β1-integrins.
16
3 Integrin trafficking
Internalization of integrins occurs by endocytosis (Fig 9). Endocytosed integrins travel
from early endosomes (EEs) via late endosomes to multivesicular endosomes (MVEs),
and ultimately end up in lysosomes for degradation. Alternatively, they are recycled to
the plasma membrane. Newly synthesized integrins are delivered to the plasma
membrane via the biosynthetic-secretory pathway. This pathway involves trafficking
from the endoplasmic reticulum (ER) via the Golgi network to the plasma membrane. In
the Golgi, integrins can be redirected to the ER, or targeted for lysosomal degradation,
for example when they are not properly folded.
Figure 9. A model for integrin traffic. Integrin-ECM interaction leads to clustering of integrins and
formation of FCs. β1 integrins and ECM proteins are internalized with the help of dynamin and activated
PKCα. Internalized β1 integrins travel to the early endosomes, from where they can be either degraded
in lysosomes recycled to the PM via a Rab4-dependent mechanism (short loop) or continued to the
perinuclear recycling compartment (PNRC; long loop). (from Pellinen and Ivaska, 2006)
3.1 Integrin endocytosis
Endocytosis of integrins occurs mainly in a clathrin-dependent manner, although
clathrin-independent mechanisms such as caveolin have been implicated as well (Table
V; Shi & Sottile, 2008; Caswell et al., 2009). During clathrin-mediated endocytosis, three
clathrin subunits form a clathrin triskelion. The triskelions assemble into pentagons or
hexagons, thereby forming a coat around the vesicles (clathrin-coated vesicles; CCVs).
Recruitment of integrins into CCVs requires the action of bridging proteins or adaptors
17
that bind both clathrin and the integrin (reviewed by Caswell et al., 2009). In addition,
clathrin-mediated endocytosis requires proteins that mediate vesicle formation and
fusion with target membranes. Regulators of clathrin-mediated endocytosis include
dynamin, the AP1-AP4 adaptor complexes, Numb, Dab2, SNAREs (soluble NSF (Nethylmaleimide-sensitive fusion protein) attachment protein receptor), and Rabs. The
GTPase dynamin assembles around the neck of the bud and influences the rate by which
vesicles pinch off. AP2 is a heterotetrameric complex that is crucial for linking proteins
to clathrin and for recruiting additional proteins required during endocytosis (Pearse et
al., 2000). AP2 and other components of the endocytic machinery like dynamin are
recruited by PIPK (Chao et al., 2009). AP2 binds directly to PIPK, and disruption of this
interaction causes AP2 to recognize endocytosis signals on receptors and subsequent
assembly of a clathrin coat (Bairstow et al., 2006). SNAREs regulate the docking and
fusing of vesicles with target membranes and provide specificity by selecting for the
appropriate target membrane. There are two types of SNAREs; vesicle-SNAREs (vSNAREs) and target-SNAREs (t-SNAREs). Together they bring the vesicle and the target
membrane in close proximity to facilitate membrane fusion. Disrupting SNARE function
impairs trafficking of α5β1, which interferes indirectly with integrin signaling,
lamellipodium extension, cell migration, and FA turnover. Integrin signaling is probably
impaired due to a lack of activation of a FAK/Src/PI3kinase-dependent pathway, which
requires SNARE-dependent trafficking of Src to the membrane and subsequent FAK-Src
interaction. Overexpression of membrane-targeted Src or phosphatidylinositol 3-kinase
(PI3K) rescued cell spreading and FA turnover in the absence of SNARE function
(Skalski et al., 2011). Different SNAREs act in different parts of the endocytic or the
biosynthetic-secretory pathway (Skalski et al., 2010; Xu et al., 2002; Mallard et al.,
2002).
The Rabs are proteins from the Ras superfamily of GTPases, which are essential for both
the delivery of newly synthesized proteins to the plasma membrane, as well as for the
endocytosis and recycling of plasma membrane proteins. Rab-dependent trafficking
occurs along microtubules (Pellinen et al., 2006). The two main Rabs involved in
Table V. Pathways for the internalization of integrins. (from
Caswell et al., 2009) ‡ ‡A direct physical association with integrins
has been shown, §Active conformation. AP2, adaptor protein 2;
DAB2, disabled 2; HAX1, HCLS1-associated protein X1; L1CAM, L1
cell adhesion molecule; NRP1, neuropilin 1; GIPC, GAIP carboxy
terminus-interacting protein; PKC, protein kinase C
18
integrin recycling are Rab4 and Rab11, which regulate a short-loop or a long-loop
recycling route, respectively. The growth factor PDGF caused Rab4-dependent rapid
recycling of αvβ3 but not α5β1, while in serum-starved cells internalized αvβ3 is
trafficked from the Rab4-positive EEs to the Rab11-positive perinuclear recycling
compartment (PNRC) and then recycled back to the plasma membrane, thus following
the long-loop (Roberts et al., 2001). Protein kinase B (PKB)/Akt and PKC- are involved
in the Rab11-dependent delivery of integrins to the plasma membrane, by promoting
the release of integrins from the PNRC (Caswell & Norman, 2006). Only a small fraction
of all integrins is present in CCVs, since around 0.1 to 1.3% (depending on the integrin)
of the plasma membrane pool is internalized every minute, and CCVs have a half-life of
1-3 minutes (Bretscher 1992; Teckchandani et al., 2009; Puthenveedu & von Zastrow,
2006). A particularly actively cycling integrin is α5β1 (Bretscher, 1992). A fraction of
the endocytosed α5β1 is trafficked from EEs via MVEs to lysosomes for degradation (Ng
et al., 1999). This fraction is probably small and consists of ligand-bound integrins, since
α5β1 and Fn colocalize in MVEs, whereas integrins that did not colocalize with Fn were
mostly detected in EEs, suggesting that these are recycled to the plasma membrane. The
α5-tail of α5β1 is ubiquitinated upon binding of Fn, which is required for the
degradation, probably by constituting a sorting-signal in MVEs (Lobert and Stenmark,
2010). Probably, not all ligand-bound integrins are ubiquitinated for degradation, but a
part may recycle via the long-loop, after being released from their ligand. Further
experiments are needed to determine whether all ubiquitinated integrins are degraded
or not, whether this is general for all integrins or cell types, and if and how integrin
signaling proceeds after internalization. Ubiquitination-induced degradation ensures
that occupied integrins are freed from their ligand, and is required for proper cell
migration FA disassembly, whereas integrin recycling induces the formation of new
adhesions to the ECM. Integrin activation is necessary for degradation but not for
endocytosis, suggesting that both active and inactive integrins are internalized (Lobert
and Stenmark, 2010; Teckchandani et al., 2009). Indeed, internalization of unligated
integrins may be particularly important in migrating cells, to transfer them to the
leading edge in order to form new adhesion sites (Jones et al., 2006; Caswell et al., 2009;
Nishimura et al., 2007). This process does not seem to involve trafficking from the
lagging to the leading edge, but rather the localized transport of integrins over short
distances towards the leading edge (Rappoport and Simon, 2003; Nishimura et al.,
2007). Dab2 may be required for bulk endocytosis of inactive integrins, since it
colocalizes with β1 integrins spread over the entire cell surface, instead of only with
active integrins residing in FAs (Teckchandani, et al. 2009). Endocytosis of unligated
integrins is significantly delayed compared to the rapid endocytosis of active integrins
after microtubule-induced FA disassembly (Chao et al., 2009). Thus, integrin trafficking
is emerging as an important regulator of adhesion assembly and disassembly, and in
this way it regulates cell adhesion and migration. Furthermore, remodeling of the
matrix also requires endocytosis, and it seems that for the turnover of matrix-Fn,
endocytosis of Fn-binding integrins via caveolae is important (Shi and Sottile, 2008).
Caveolae are plasma membrane invaginations containing lipid rafts, the structural
protein caveolin, and caveolin-associated cavins (Hansen and Nichols, 2010). At
present, caveolin-mediated integrin endocytosis is poorly understood, but it seems to
involve a direct association of PKCα with the β1-tail (Ng et al., 1999; Upla et al., 2004).
19
3.2 Anterograde integrin transport
Both the biosynthetic-secretory pathway and the recycling of internalised integrins
deliver integrins to the plasma membrane, and the combination of the two is referred to
as anterograde transport. Protein kinase D1 (PKD1) binds specifically to the
cytoplasmic tail of β3 and thus regulates anterograde transport of β3 integrins.
Interestingly, depletion of PKD1 impairs all anterograde αvβ3 transport, whereas
prevention of phosphorylation on residue 916 specifically prevents integrin trafficking
via the Golgi network, but not integrin recycling (Liljedahl et al., 2001; Woods et al.,
2004; White et al., 2007). Interestingly, talin is also involved in anterograde transport of
newly synthesized integrins (Albiges-Rizo et al., 1995; Martel et al., 2000). Talin
controls the exit of integrins from the ER to the Golgi, for which different explanations
are given. One explanation is that talin might link integrin-containing vesicles to myosin
VII or myosin X and thereby regulates the transport along the actin cytoskeleton.
Alternatively, talin binding exposes a GFFKR sequence in the α-subunit that may act as
an integrin export signal (Martel et al., 2000; O’Toole et al., 1991, 1994). It remains
unclear how talin contributes to transport of newly synthesized integrins, and also how
talin-integrin interaction at the ER are regulated.
3.3 Regulation of integrin trafficking by the β-subunit cytoplasmic tails
Internalization of plasma membrane receptors is regulated by internalization signals in
the cytoplasmic tails. Three types of internalization signals are distinguished; tyrosinebased sorting signals like NPxY or YxxØ motifs (Ø is a residue with a bulky hydrophobic
side chain), dileucine-based sorting signals like [DE]xxxL[LI] or DxxL motifs, and
ubiquitination of lysine residues (Bonifacino and Traub, 2003). Whereas NPxY motifs
are required for the endocytosis of several receptors, it is controversial how integrin
internalization is regulated by these motifs. Initial studies using Y-S substitutions in
either one of the NPxY motifs did not impair endocytosis of α5β1, which may be
explained because these mutations still allow formation of β-turns and thus not prevent
binding of PTB-containing proteins, as discussed above (Vignoud et al., 1994). In β3, the
MD-NITY motif seems to stimulate internalization. Of the three β3 integrin variants
(β3A-C), only β3A contains NxxY motifs, and β3A is expressed at lower levels on the cell
surface than the other isoforms (Gawaz et al., 2001). Internalization of β3A integrins is
mediated by binding of endonexin to the MD-NITY motif, and a Y-A substitution in this
motif or expression of dominant-negative endonexin increase the surface levels of β3A
integrins. In another study it was shown that Y-A substitutions in the NPxY motifs of β3
did not diminish constitutive endocytosis whereas the internalization of ligand-coated
particles was impaired suggesting that these motifs mainly regulate endocytosis of
ligand-bound integrins (Ylänne et al., 1995). In addition, substituting either one of the
NPxF motifs in β2 integrins for NPxA results in a loss of internalization, and Y-F
mutations in both NpxY motifs of β1 also reduces endocytosis, probably as a result of a
defect in clathrin binding (Rabb et al., 1993; Pellinen et al., 2008). The latter defect
could be rescued by overexpression of Rab21, suggesting a clathrin-independent
pathway of endocytosis (Pellinen et al., 2008).
20
3.4 Functional consequences of integrin trafficking
It is becoming increasingly clear that regulation of integrin trafficking is important for
optimal integrin functions during cell adhesion, cell spreading, and cell migration
(Caswell et al., 2009). For example, inhibition of integrin recycling impairs cell adhesion
and migration (Proux-Gillardeaux et al., 2005). The importance of integrin transport for
cell migration has been studied in most detail for the Fn-binding integrins αvβ3 and
α5β1. When αvβ3 is actively recycled, recycling of α5β1 is slow and vice versa, and it is
not until the recycling of αvβ3 is diminished that α5β1 recycling increases (Caswell and
Norman, 2008; White et al., 2007). As described above, PKD1 regulates both the shortloop recycling of αvβ3 as well as the transport of new αvβ3 to the membrane (Woods et
al., 2004). PKD1 contributes to directionally-persistent migration by controlling
recycling of αvβ3 integrins, whereas the PKD1-mediated regulation of integrin
transport to the membrane probably enhances the speed of migration (White et al.,
2007). Random cell migration occurs when α5β1 is actively recycled. If both αvβ3
integrins and α5β1 integrins are blocked, cells migrate in a directionally-persistent
manner (White et al., 2007).
Figure 10. Effect of integrin endocytosis on cell migration. A) Phosphorylation of Numb results in loss
of Numb binding to integrin and subsequent loss of endocytosis, leading to directional persistent
migration. B/C) Active recycling of α5β1 integrins results in random cell migration, while active recycling
of αvβ3 integrins leads to directional persistent migration. (from Petrie et al., 2009)
Numb localizes at FAs near CCVs and associates with several intracellular proteins
involved in integrin endocytosis (Nishimura et al., 2007). Numb binds to the MP-NPxY
motif, although one study showed binding to β1 and β3 integrins, while another only
finds Numb associated with motif of β3 and β5 but not with β1A (Nishimura et al., 2007;
Calderwood et al., 2003). Numb probably forms a complex with the polarity protein
Par3 and atypical-PKC (aPKC), and phosphorylation of Numb by aPKC prevents
association of Numb with integrins. Endocytosis is impaired when Numb is not
associated with the integrin, leading to directionally persistent cell migration (Petrie et
al., 2009; Nishimura et al., 2007). It is unclear whether this is caused only by impaired
β1 endocytosis, as displayed in figure 10, or whether it is due to impaired endocytosis of
21
both β1 and β3 integrins (Nishimura et al., 2007; White et al., 2007). Suppression of
Numb impairs clathrin-dependent integrin endocytosis, but to a lesser extent than
suppression of AP2 and clathrin together, demonstrating that in addition to Numb,
other clathrin-associated proteins such as Dab2 regulate integrin internalization
(Nishimura et al., 2007). Like Numb, Dab2 also binds directly to the MD-NPxY motif of
β-tails; according to one study binding does occur to β3 and β5 integrins, but not β1A
integrins, while binding to β1 integrins was observed in two other studies (Calderwood
et al., 2003; Teckchandani et al., 2009; Prunier & Howe, 2005). Dab-2 depletion leads to
increased integrin surface expression, but reduction of the integrin intracellular pool.
Reduced migration but not excess adhesion is observed in Dab2-deficient cells, and it is
thought that maintenance of the intracellular integrin pool by Dab2 normally enables
trafficking of integrins to the leading edge of the cell surface during cell migration to
form new adhesion sites (Teckchandani et al., 2009). Numb and Dab2 differ in
localization; whereas Numb localizes in FAs near CCSs at the leading edge, Dab2 is
diffusely distributed over the cell surface (Nishimura et al., 2007; Teckchandani et al.,
2009).
In conclusion, integrins can follow different routes inside the cell, resulting in either
recycling or degradation. It is likely that both active and inactive integrins can be
internalized, although there may be some difference in the proteins involved in
endocytosis. Since several studies give contradicting results, it remains unclear whether
the NxxY motifs are important as internalization signals.
22
4 Discussion
Integrins exist on the membrane in a conformation of low, intermediate, or high affinity
for ligand, and the shift from low to high affinity is called integrin activation. This
conformational change can occur from the inside, which is referred to as ‘inside-out’
signaling or activation. After ligand binding, integrins connect to the actin cytoskeleton,
which is accompanied by the formation of adhesion plaques, cell spreading over the
substrate, and the initiation of several signaling cascades (‘outside-in’ signaling). The
paradigm of inside-out integrin activation has been developed mainly from studies on
the platelet integrin αIIbβ3. Integrin αIIbβ3 and other integrins on non-adherent cells
are ‘locked’ in the low-affinity conformation by a salt bridge between the cytoplasmic
sequences of the - and the -subunits. Activation of αIIbβ3 occurs by GPCR agonists
that trigger an intracellular pathway, of which the final step is talin binding to the -tail,
which then induces a conformational change across the membrane, exposing the ligandbinding site. Although the mechanisms of αIIbβ3 activation are often assumed to be
generic to all integrins, several findings especially from in vivo studies indicate
otherwise. For example, YY/FF substitutions in the NPxY motifs of β3 in mice impair in
vivo integrin function, while these substitutions do not affect the functions of β1
integrins in mice (Law et al., 1999; Czuchra et al., 2006; Chen et al., 2006). In addition,
the salt bridge between the α- and β-subunit that stabilizes the inactive conformation is
crucial for αIIbβ3, while mice carrying a D759A substitution in β1, which impairs the
formation of a salt bridge, do not have any obvious defects (Anthis et al., 2009; Czuchra
et al., 2006). Differences in the regulation of activation may reflect differences in specific
integrin heterodimers, or differential modes of regulation existing between integrins on
non-adherent versus integrins on adherent cells. The MP-NPxY and the MD-NxxY motifs
in the cytoplasmic tail sequence of integrins are implicated both in integrin activation
and outside-in signaling (table VI). Many studies have demonstrated that the MP-NPxY
is important for talin binding, and that loss of talin binding impairs integrin activation
(Calderwood et al., 2002; García-Alvarez et al., 2003; Bouaouina et al., 2008; Anthis et
al., 2009; Tadokoro et al., 2003; Wegener et al., 2007). It seems that the talin-head is
both necessary and sufficient for activation, while full-length talin (including the rod) is
required for downstream events including FA assembly and cell adhesion and migration
(Calderwood et al., 2003; Bouaouina et al., 2008; Zhang et al., 2008; Monkley et al.,
2000; Nieswandt et al., 2007; Petrich et al., 2007). Interestingly, talin disruption or
expression of only the talin-head in Drosophila does not impair integrin-ligand
interaction but results nevertheless in an integrin-knockout phenotype, suggesting that
inside-out events are less important than outside-in events in adherent cells in tissue.
This is further underlined by studies that show that the conformational change can also
be triggered by ligand, irrespective of cytoplasmic events (Du et al., 1991). It is thus
unclear to what extent inside-out mechanisms contribute to integrin activation in
adherent cells.
The MD-NxxY motif is bound by kindlins, which are essential modulators of integrin
function, as is clearly illustrated by the human disorders KS and LAD-III, as well as the
phenotypes of kindlin-knockout mice (Moser et al., 2008; Ussar et al., 2008; Montanez et
al., 2008). Whereas kindlin-3 is clearly essential for talin-induced activation of integrins
on non-adherent cells, it is controversial whether kindlins are essential for inside-out
activation in adherent cells. For instance, kindlin-3 does not enhance β1 activation in
CHO cells, and kindlin-2 in αIIbβ3-expressing CHO cells has been reported to either
23
weakly activate or inhibit αIIbβ3 activation (Shi et al., 2007; Ma et al., 2008; Montanez
et al., 2008; Harburger et al., 2009). In addition, talin binding alone can activate
integrins without additional binding of kindlin, whereas kindlin binding alone cannot
activate integrins (Ma et al., 2008). Similarly, a Y-A substitution in the MD-NxxY motif,
which inhibits kindlin binding, does not prevent activation of αIIbβ3, but a Y-A mutation
in the M-P-NpxY completely abolishes activation (Ye et al., 2010). An interesting
possibility is that kindlins have integrin-specific effects, as suggested by a number of
studies. Both kindlin-1 and kindlin-2 can activate αIIbβ3 with talin, while they inhibit
talin-mediated α5β1 activation, and knockdown of kindlin-2 but not kindlin-3 decreases
adhesion to vitronectin, while adhesion to fibronectin is affected by knockdown of
kindlin-3 but not kindlin-2 (Harburger et al., 2009; Bialkowska et al., 2010).
Accumulating evidence suggests that the kindlins regulate cell surface expression of
integrins. Kindlin-3-deficient platelets express low surface levels of αIIbβ3 and α5β1
compared to wild-type platelets, and kindlin-3-deficient macrophages also express less
integrins than wild-type macrophages (Moser et al., 2008; Schmidt et al., 2011).
Similarly, expression of αIIbβ3 and β1 decreases upon knockdown of kindlin-2, and
expression of β1 integrins in the epidermis of KS patients is low with respect to normal
individuals (Lai-Cheong et al., 2009; Qu et al., 2011). Conversely, overexpression of
kindlin-1 and kindlin-2 boosts the surface expression of both α5β1 and αIIbβ3
(Harburger et al., 2009). Interestingly, surface expression of β1 integrins in mouse
keratinocytes isolated from mice carrying a YY/AA substitution in the epidermis is also
significantly reduced, which is most likely due to a loss of kindlin binding, in light of the
observations described above (Czuchra et al., 2006). Reduction of total integrin levels,
in the absence of events that disrupt activation, was studied in hypomorphic mice
expressing reduced β1 integrin levels in the epidermis. These mice develop a skin
phenotype similar to mice that completely lack β1 in the epidermis, albeit delayed and
less severe (Piwko-Czuchra et al., 2009). In vitro, reduced β1 integrin levels were
sufficient to support keratinocyte adhesion, whereas outside-in signaling events as
cytoskeletal reorganization, cell spreading and proliferation were impaired. This
indicates that downstream of ligand binding and initial cell adhesion, maintenance of
optimal integrin levels is required for a number of cell functions. Cell surface levels are
maintained by trafficking mechanisms, and it is indeed increasingly recognized that
regulation of integrin trafficking is important for integrin-mediated processes including
cell adhesion, cell spreading, and cell migration (Caswell et al., 2009). A role for kindlins
in the control of integrin trafficking may also explain the integrin-specific effects of
kindlins, as the α-subunits are very important in determining the trafficking properties
of the heterodimer. For example, the α-subunit GFFKR sequence, present in all αsubunits, is suggested to serve as an export signal, and mutations in this sequence cause
reduced surface expression, as observed in some GT patients (Peyruchaud et al., 1998;
Martel et al., 2000). Interestingly, exposure of the GFFKR upon talin-binding has been
postulated to regulate export of newly synthesized integrins from the ER to the plasma
membrane (Martel et al., 2000). In addition to membrane delivery, endocytosis and
recycling likely depend on the α-subunits. For instance, whereas rapid and constitutive
endocytosis occurs for α5β1 and αMβ2, no or low endocytosis is observed for α3β1,
α4β1 and αLβ2 (Bretscher et al., 1992). In addition, the stabilizing effect of Dab2 on
surface expression is restricted to α1β1, α2β1, and α3β1, but does not apply to α5β1 or
αv-containing integrins (Teckchandani et al., 2009). The α-subunit may also stimulate
endocytosis and lysosomal degradation by ubiquitination, as observed for α5 (Lobert et
al., 2010).
24
In conclusion, it seems that talin-binding to the MP-NPxY motif is crucial for both insideout integrin activation and connection to the cytoskeleton. Whereas the importance of
inside-out activation in non-adherent cells is clear, its relevance in adherent cells
remains questionable. Kindlin-3 binding to the MD-NxxY motif is also essential for
inside-out activation of integrins in non-adherent cells, whereas the effects of kindlins
on talin-dependent activation in adherent cells are controversial, and can be
stimulatory as well as inhibitory. Alternatively, the kindlins may control integrin
functions by stabilizing their cell surface expression, probably by regulating integrin
trafficking. Ultimately, integrin function will depend on the sum of integrin activation
mechanisms -both from the inside and the outside- as well as efficient integrin turnover.
The specific requirements for optimal integrin function are likely different for each
specific heterodimer.
25
β1 Mutation
/β3
In vivo
β1 D759A
β1 Y783F
Phenotype
Mice/cell lines
Author
viable, normal β1 integrin function
viable, normal β1 integrin function
mice
mice
β1 Y795F
viable, normal β1 integrin function
mice
β1 YY783,795FF
viable, normal β1 integrin function
mice
Czuchra et al., 2006
Czuchra et al., 2006;
Chen et al., 2006
Czuchra et al., 2006;
Chen et al., 2006
Czuchra et al., 2006;
Chen et al., 2006
Czuchra et al., 2006;
Chen et al., 2006
Czuchra et al., 2006
β1 YY783,795AA not viable
β1 YY783,795AA inactive β1 integrins, reduced expression
mice
mice; basal
keratinocytes
β1 YY783,795FF impaired function, active β1 integrins, normal expression mice; in platelets only
β1 YY783,795AA inactive β1 integrins, reduced expression
mice; in platelets only
β1 YY783,795FF reduced invasion and adhesion, normal metastatic capacity ESb lymphoma cells
β1 T788,789A
impaired invasion, adhesion, metastatis to liver, normal
ESb lymphoma cells
metastasis to spleen
β3 YY747,759FF tendency to rebleed, normal β3 expression, normal
mice
platelet count
In vitro
β1 D759A
normal β1 integrin function
mouse keratinocytes
β1 YY783,795FF normal β1 integrin function
mouse keratinocytes
β1 YY783,795FF normal-improved adhesion to Fn, laminin and collagen
mouse ES cells
β1 YY783,795AA inactive β1 integrins, impaired adhesion to Fn, laminin and mouse ES cells
collagen
β1 Y783F
normal β1 integrin function
ESb lymphoma cells
β1 Y795F
failed to express Y795F in Esb lymphoma cells
ESb lymphoma cells
β1 YY783,795FF reduced invasion, poor adhesion
ESb lymphoma cells
β1 TT788,789AA impaired invasion, reduced adhesion
ESb lymphoma cells
β1 Y795F
normal β1 integrin function
Jurkat T cell A1
β1 Y795A
adhesion to Fn reduced, inactive β1 integrins, normal
Jurkat T cell A1
expression
β1 del MP-NPxY adhesion to Fn reduced, normal β1 integrin expression
Jurkat T cell A1
β1 del MD-NPxY adhesion to Fn reduced, normal β1 integrin expression
Jurkat T cell A1
β1 D759A
active β1 integrin, normal migration
GD25 cells
β1 Y783F
active β1 integrin, normal adhesion, Fn matrix formation, GD25 cells
reduced migration
β1 Y795F
active β1 integrin, normal adhesion, Fn matrix formation, GD25 cells
reduced migration
β1 YY783,795FF active β1 integrin, normal adhesion, Fn matrix formation, GD25 cells
impaired migration
β1 Y783S
lack of recruitment to FAs
CHO cells
β1 Y795S
lack of recruitment to Fas
CHO cells
β1 YY783,795SS lack of recruitment to FAs
CHO cells
β1 Y783F
active β1 integrin
CHO cells
β1 Y783A
inactive β1 integrin
CHO cells
β1 Y795F
active β1 integrin
CHO cells
β1 Y795A
active β1 integrin
CHO cells
β3 Y747A
inactive β3 integrin, impaired recruitment to FAs, impaired CHO cells
cell spreading on Fn
β3 Y759A
β3 integrin activation reduced, normal recruitment to FAs, CHO cells
reduced cell spreading on Fn
β3 YY747,759FF no aggregation upon thrombin-stimulation
mouse platelets
Chen et al., 2006
Chen et al., 2006
Stroeken et al., 2000
Stroeken et al., 2000
Law et al., 1999
Czuchra et al., 2006
Czuchra et al., 2006
Chen et al., 2006
Chen et al., 2006
Stroeken et al., 2000
Stroeken et al., 2000
Stroeken et al., 2000
Stroeken et al., 2000
Romzek et al., 1998
Romzek et al., 1998
Romzek et al., 1998
Romzek et al., 1998
Sakai et al., 1998
Sakai et al., 1998, 2001
Sakai et al., 1998, 2001
Sakai et al., 1998, 2001
Vignoud et al., 1997
Vignoud et al., 1997
Vignoud et al., 1997
O'Toole et al., 1995
O'Toole et al., 1995
O'Toole et al., 1995
O'Toole et al., 1995
O'Toole et al., 1995;
Ylänne et al., 1995
O'Toole et al., 1995;
Ylänne et al., 1995
Law et al., 1999
Table VI. Implications of mutations in the NPxY motifs on integrin functions in vivo and in vitro.
26
5 Abbreviations
βTD
CHO
Col
CCV
Dab
DAG
E-Cadherin
ECM
EE
EGF
ER
FA
FAK
Fb
FC
FERM
Fg
Fn
FX
GT
GPCR
HD
ICAM
iC3b
ILK
KS
LAD
LAP-TGF-β
Ln
MAdCAM-1
MD
MIDAS
MP
MVE
Nn
Op
PECAM-1
PIPK
PI4,5P2
PKC
PKD1
PNRC
PSI
PTB
Rap-GEF
RGD
TGN
Tn
TSP
VCAM-1
Vn
vWF
= β-tail domain
= Chinese Hamster ovary
= collagen
= clathrin-coated vesicles
= disabled
= diacylglycerol
= epithelial cadherin
= extracellular matrix
= early endosome
= epidermal growth factor
= endoplasmic reticulum
= focal adhesion
= focal adhesion kinase
= fibrillin
= focal complex
= four-point-one, ezrin, radixin, moesin
= fibrinogen
= fibronectin
= factor X
= Glanzmann thrombasthenia
= G-protein coupled receptor
= hemidesmosome
= intercellular adhesion molecule
= inactive complement factor 3b
= integrin-linked kinase
= Kindler syndrome
= leukocyte adhesion deficiency
= latency-associated peptide of transforming growth factor-β
= laminin
= mucosal addressin cell adhesion molecule-1
= membrane-distal
= metal ion-dependent adhesion site
= membrane-proximal
= multivesicular endosome
= nephronectin
= osteopontin
= platelet endothelial cell adhesion molecule-1
= phosphatidylinositol phosphate kinase
= phosphatidylinositol-4,5-biphosphate
= protein kinase C
= protein kinase D1
= perinuclear recycling compartment
= plexin/semaphorin/integrin
= phosphotyrosine-binding domain
= Rap-guanine exchange factor
= Arg-Gly-Asp
= trans Golgi network
= tenascin
= thrombospondin
= vascular cell adhesion molecule-1
= vitronectin
= von Willebrand Factor
27
6 Reference list
1.
Albigès-Rizo, C., Frachet, P., & Block, M. R. (1995) Down regulation of talin alters cell
adhesion and the processing of the alpha 5 beta 1 integrin. J Cell Sci 108 ( Pt 10), 3317-29.
2.
Anthis, N. J., Wegener, K. L., Ye, F., Kim, C., Goult, B. T., Lowe, E. D., Vakonakis, I., Bate, N.,
Critchley, D. R., Ginsberg, M. H., & Campbell, I. D. (2009) The structure of an integrin/talin
complex reveals the basis of inside-out signal transduction. EMBO J 28, 3623-32.
3.
Arias-Salgado, E. G., Lizano, S., Sarkar, S., Brugge, J. S., Ginsberg, M. H., & Shattil, S. J.
(2003) Src kinase activation by direct interaction with the integrin beta cytoplasmic domain.
Proc Natl Acad Sci U S A 100, 13298-302.
4.
Askari, J. A., Tynan, C. J., Webb, S. E. D., Martin-Fernandez, M. L., Ballestrem, C., &
Humphries, M. J. (2010) Focal adhesions are sites of integrin extension. J Cell Biol 188, 891903.
5.
Bairstow, S. F., Ling, K., Su, X., Firestone, A. J., Carbonara, C., & Anderson, R. A. (2006) Type
Igamma661 phosphatidylinositol phosphate kinase directly interacts with AP2 and regulates
endocytosis. J Biol Chem 281, 20632-42.
6.
Belkin, A. M., Retta, S. F., Pletjushkina, O. Y., Balzac, F., Silengo, L., Fassler, R., Koteliansky,
V. E., Burridge, K., & Tarone, G. (1997) Muscle beta1D integrin reinforces the cytoskeletonmatrix link: modulation of integrin adhesive function by alternative splicing. J Cell Biol 139,
1583-95.
7.
Bialkowska, K., Ma, Y.-Q., Bledzka, K., Sossey-Alaoui, K., Izem, L., Zhang, X., Malinin, N.,
Qin, J., Byzova, T., & Plow, E. F. (2010) The integrin co-activator Kindlin-3 is expressed and
functional in a non-hematopoietic cell, the endothelial cell. J Biol Chem 285, 18640-9.
8.
Bledzka, K., Bialkowska, K., Nie, H., Qin, J., Byzova, T., Wu, C., Plow, E. F., & Ma, Y.-Q.
(2010) Tyrosine phosphorylation of integrin beta3 regulates kindlin-2 binding and integrin
activation. J Biol Chem 285, 30370-4.
9.
Bonifacino, J. S. & Traub, L. M. (2003) Signals for sorting of transmembrane proteins to
endosomes and lysosomes. Annu Rev Biochem 72, 395-447.
10. Bouaouina, M., Lad, Y., & Calderwood, D. A. (2008) The N-terminal domains of talin cooperate
with the phosphotyrosine binding-like domain to activate beta1 and beta3 integrins. J Biol Chem
283, 6118-25.
11. Brakebusch, C., Grose, R., Quondamatteo, F., Ramirez, A., Jorcano, J. L., Pirro, A., Svensson,
M., Herken, R., Sasaki, T., Timpl, R., Werner, S., & Fässler, R. (2000) Skin and hair follicle
integrity is crucially dependent on beta 1 integrin expression on keratinocytes. EMBO J 19,
3990-4003.
12. Brandenberger, R., Schmidt, A., Linton, J., Wang, D., Backus, C., Denda, S., Müller, U., &
Reichardt, L. F. (2001) Identification and characterization of a novel extracellular matrix protein
nephronectin that is associated with integrin alpha8beta1 in the embryonic kidney. J Cell Biol
154, 447-58.
13. Bretscher, M. S. (1992) Circulating integrins: alpha 5 beta 1, alpha 6 beta 4 and Mac-1, but not
alpha 3 beta 1, alpha 4 beta 1 or LFA-1. EMBO J 11, 405-10.
14. Brown, N. H., Gregory, S. L., Rickoll, W. L., Fessler, L. I., Prout, M., White, R. A. H., &
Fristrom, J. W. (2002) Talin is essential for integrin function in Drosophila. Dev Cell 3, 569-79.
15. Burridge, K. & Chrzanowska-Wodnicka, M. (1996) Focal adhesions, contractility, and signaling.
Annu Rev Cell Dev Biol 12, 463-518.
28
16. Calderwood, D. A., Fujioka, Y., de Pereda, J. M., García-Alvarez, B., Nakamoto, T., Margolis,
B., McGlade, C. J., Liddington, R. C., & Ginsberg, M. H. (2003) Integrin beta cytoplasmic
domain interactions with phosphotyrosine-binding domains: a structural prototype for diversity
in integrin signaling. Proc Natl Acad Sci U S A 100, 2272-7.
17. Calderwood, D. A., Yan, B., de Pereda, J. M., Alvarez, B. G., Fujioka, Y., Liddington, R. C., &
Ginsberg, M. H. (2002) The phosphotyrosine binding-like domain of talin activates integrins. J
Biol Chem 277, 21749-58.
18. Caswell, P. & Norman, J. (2008) Endocytic transport of integrins during cell migration and
invasion. Trends Cell Biol 18, 257-63.
19. Caswell, P. T. & Norman, J. C. (2006) Integrin trafficking and the control of cell migration.
Traffic 7, 14-21.
20. Caswell, P. T., Vadrevu, S., & Norman, J. C. (2009) Integrins: masters and slaves of endocytic
transport. Nat Rev Mol Cell Biol 10, 843-53.
21. Chao, W.-T. & Kunz, J. (2009) Focal adhesion disassembly requires clathrin-dependent
endocytosis of integrins. FEBS Lett 583, 1337-43.
22. Chen, H., Zou, Z., Sarratt, K. L., Zhou, D., Zhang, M., Sebzda, E., Hammer, D. A., & Kahn, M.
L. (2006) In vivo beta1 integrin function requires phosphorylation-independent regulation by
cytoplasmic tyrosines. Genes Dev 20, 927-32.
23. Chen, Y. P., O'Toole, T. E., Ylänne, J., Rosa, J. P., & Ginsberg, M. H. (1994) A point mutation
in the integrin beta 3 cytoplasmic domain (S752-->P) impairs bidirectional signaling through
alpha IIb beta 3 (platelet glycoprotein IIb-IIIa). Blood 84, 1857-65.
24. Clark, K., Pankov, R., Travis, M. A., Askari, J. A., Mould, A. P., Craig, S. E., Newham, P.,
Yamada, K. M., & Humphries, M. J. (2005) A specific alpha5beta1-integrin conformation
promotes directional integrin translocation and fibronectin matrix formation. J Cell Sci 118, 291300.
25. Cluzel, C., Saltel, F., Lussi, J., Paulhe, F., Imhof, B. A., & Wehrle-Haller, B. (2005) The
mechanisms and dynamics of (alpha)v(beta)3 integrin clustering in living cells. J Cell Biol 171,
383-92.
26. Czuchra, A., Meyer, H., Legate, K. R., Brakebusch, C., & Fässler, R. (2006) Genetic analysis of
beta1 integrin "activation motifs" in mice. J Cell Biol 174, 889-99.
27. Danen, E. H. J., Sonneveld, P., Brakebusch, C., Fassler, R., & Sonnenberg, A. (2002) The
fibronectin-binding integrins alpha5beta1 and alphavbeta3 differentially modulate RhoA-GTP
loading, organization of cell matrix adhesions, and fibronectin fibrillogenesis. J Cell Biol 159,
1071-86.
28. Du, X. P., Plow, E. F., Frelinger, 3rd, A. L., O'Toole, T. E., Loftus, J. C., & Ginsberg, M. H.
(1991) Ligands "activate" integrin alpha IIb beta 3 (platelet GPIIb-IIIa). Cell 65, 409-16.
29. Franchini, M., Favaloro, E. J., & Lippi, G. (2010) Glanzmann thrombasthenia: an update. Clin
Chim Acta 411, 1-6.
30. Friedland, J. C., Lee, M. H., & Boettiger, D. (2009) Mechanically activated integrin switch
controls alpha5beta1 function. Science 323, 642-4.
31. García-Alvarez, B., de Pereda, J. M., Calderwood, D. A., Ulmer, T. S., Critchley, D., Campbell,
I. D., Ginsberg, M. H., & Liddington, R. C. (2003) Structural determinants of integrin
recognition by talin. Mol Cell 11, 49-58.
32. Gawaz, M., Besta, F., Ylänne, J., Knorr, T., Dierks, H., Böhm, T., & Kolanus, W. (2001) The
NITY motif of the beta-chain cytoplasmic domain is involved in stimulated internalization of the
beta3 integrin A isoform. J Cell Sci 114, 1101-13.
29
33. Geiger, B., Bershadsky, A., Pankov, R., & Yamada, K. M. (2001) Transmembrane crosstalk
between the extracellular matrix--cytoskeleton crosstalk. Nat Rev Mol Cell Biol 2, 793-805.
34. Ginsberg, M. H., Du, X., & Plow, E. F. (1992) Inside-out integrin signalling. Curr Opin Cell Biol
4, 766-71.
35. Goksoy, E., Ma, Y.-Q., Wang, X., Kong, X., Perera, D., Plow, E. F., & Qin, J. (2008) Structural
basis for the autoinhibition of talin in regulating integrin activation. Mol Cell 31, 124-33.
36. Han, J., Lim, C. J., Watanabe, N., Soriani, A., Ratnikov, B., Calderwood, D. A., PuzonMcLaughlin, W., Lafuente, E. M., Boussiotis, V. A., Shattil, S. J., & Ginsberg, M. H. (2006)
Reconstructing and deconstructing agonist-induced activation of integrin alphaIIbbeta3. Curr
Biol 16, 1796-806.
37. Hansen, C. G. & Nichols, B. J. (2010) Exploring the caves: cavins, caveolins and caveolae.
Trends Cell Biol 20, 177-86.
38. Harburger, D. S., Bouaouina, M., & Calderwood, D. A. (2009) Kindlin-1 and -2 directly bind the
C-terminal region of beta integrin cytoplasmic tails and exert integrin-specific activation effects.
J Biol Chem 284, 11485-97.
39. Hilden, T. J., Valmu, L., Kärkkäinen, S., & Gahmberg, C. G. (2003) Threonine phosphorylation
sites in the beta 2 and beta 7 leukocyte integrin polypeptides. J Immunol 170, 4170-7.
40. Humphries, J. D., Byron, A., & Humphries, M. J. (2006) Integrin ligands at a glance. J Cell Sci
119, 3901-3.
41. Huveneers, S., Truong, H., Fässler, R., Sonnenberg, A., & Danen, E. H. J. (2008) Binding of
soluble fibronectin to integrin alpha5 beta1 - link to focal adhesion redistribution and contractile
shape. J Cell Sci 121, 2452-62.
42. Hynes, R. O. (1992) Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69,
11-25.
43. Hynes, R. O. (2002) Integrins: bidirectional, allosteric signaling machines. Cell 110, 673-87.
44. Ivaska, J., Pallari, H.-M., Nevo, J., & Eriksson, J. E. (2007) Novel functions of vimentin in cell
adhesion, migration, and signaling. Exp Cell Res 313, 2050-62.
45. Jones, M. C., Caswell, P. T., & Norman, J. C. (2006) Endocytic recycling pathways: emerging
regulators of cell migration. Curr Opin Cell Biol 18, 549-57.
46. Kiema, T., Lad, Y., Jiang, P., Oxley, C. L., Baldassarre, M., Wegener, K. L., Campbell, I. D.,
Ylänne, J., & Calderwood, D. A. (2006) The molecular basis of filamin binding to integrins and
competition with talin. Mol Cell 21, 337-47.
47. Lai-Cheong, J. E., Parsons, M., Tanaka, A., Ussar, S., South, A. P., Gomathy, S., Mee, J. B.,
Barbaroux, J.-B., Techanukul, T., Almaani, N., Clements, S. E., Hart, I. R., & McGrath, J. A.
(2009) Loss-of-function FERMT1 mutations in kindler syndrome implicate a role for fermitin
family homolog-1 in integrin activation. Am J Pathol 175, 1431-41.
48. Law, D. A., DeGuzman, F. R., Heiser, P., Ministri-Madrid, K., Killeen, N., & Phillips, D. R.
(1999) Integrin cytoplasmic tyrosine motif is required for outside-in alphaIIbbeta3 signalling and
platelet function. Nature 401, 808-11.
49. Legate, K. R. & Fässler, R. (2009) Mechanisms that regulate adaptor binding to beta-integrin
cytoplasmic tails. J Cell Sci 122, 187-98.
50. Legate, K. R., Montañez, E., Kudlacek, O., & Fässler, R. (2006) ILK, PINCH and parvin: the
tIPP of integrin signalling. Nat Rev Mol Cell Biol 7, 20-31.
51. Liljedahl, M., Maeda, Y., Colanzi, A., Ayala, I., Van Lint, J., & Malhotra, V. (2001) Protein
kinase D regulates the fission of cell surface destined transport carriers from the trans-Golgi
network. Cell 104, 409-20.
30
52. Ling, K., Doughman, R. L., Iyer, V. V., Firestone, A. J., Bairstow, S. F., Mosher, D. F., Schaller,
M. D., & Anderson, R. A. (2003) Tyrosine phosphorylation of type Igamma phosphatidylinositol
phosphate kinase by Src regulates an integrin-talin switch. J Cell Biol 163, 1339-49.
53. Lobert, V. H., Brech, A., Pedersen, N. M., Wesche, J., Oppelt, A., Malerød, L., & Stenmark, H.
(2010) Ubiquitination of alpha 5 beta 1 integrin controls fibroblast migration through lysosomal
degradation of fibronectin-integrin complexes. Dev Cell 19, 148-59.
54. Lu, C., Takagi, J., & Springer, T. A. (2001) Association of the membrane proximal regions of the
alpha and beta subunit cytoplasmic domains constrains an integrin in the inactive state. J Biol
Chem 276, 14642-8.
55. Ma, Y.-Q., Qin, J., Wu, C., & Plow, E. F. (2008) Kindlin-2 (Mig-2): a co-activator of beta3
integrins. J Cell Biol 181, 439-46.
56. Mallard, F., Tang, B. L., Galli, T., Tenza, D., Saint-Pol, A., Yue, X., Antony, C., Hong, W.,
Goud, B., & Johannes, L. (2002) Early/recycling endosomes-to-TGN transport involves two
SNARE complexes and a Rab6 isoform. J Cell Biol 156, 653-64.
57. Margadant, C., Frijns, E., Wilhelmsen, K., & Sonnenberg, A. (2008) Regulation of
hemidesmosome disassembly by growth factor receptors. Curr Opin Cell Biol 20, 589-96.
58. Margadant, C., Raymond, K., Kreft, M., Sachs, N., Janssen, H., & Sonnenberg, A. (2009)
Integrin alpha3beta1 inhibits directional migration and wound re-epithelialization in the skin. J
Cell Sci 122, 278-88.
59. Martel, V., Racaud-Sultan, C., Dupe, S., Marie, C., Paulhe, F., Galmiche, A., Block, M. R., &
Albiges-Rizo, C. (2001) Conformation, localization, and integrin binding of talin depend on its
interaction with phosphoinositides. J Biol Chem 276, 21217-27.
60. Martel, V., Vignoud, L., Dupé, S., Frachet, P., Block, M. R., & Albigès-Rizo, C. (2000) Talin
controls the exit of the integrin alpha 5 beta 1 from an early compartment of the secretory
pathway. J Cell Sci 113 ( Pt 11), 1951-61.
61. McCleverty, C. J., Lin, D. C., & Liddington, R. C. (2007) Structure of the PTB domain of
tensin1 and a model for its recruitment to fibrillar adhesions. Protein Sci 16, 1223-9.
62. Meves, A., Stremmel, C., Gottschalk, K., & Fässler, R. (2009) The Kindlin protein family: new
members to the club of focal adhesion proteins. Trends Cell Biol 19, 504-13.
63. Moik, D. V., Janbandhu, V. C., & Fässler, R. (2011) Loss of migfilin expression has no overt
consequences on murine development and homeostasis. J Cell Sci 124, 414-21.
64. Monkley, S. J., Zhou, X. H., Kinston, S. J., Giblett, S. M., Hemmings, L., Priddle, H., Brown, J.
E., Pritchard, C. A., Critchley, D. R., & Fässler, R. (2000) Disruption of the talin gene arrests
mouse development at the gastrulation stage. Dev Dyn 219, 560-74.
65. Montanez, E., Ussar, S., Schifferer, M., Bösl, M., Zent, R., Moser, M., & Fässler, R. (2008)
Kindlin-2 controls bidirectional signaling of integrins. Genes Dev 22, 1325-30.
66. Moser, M., Legate, K. R., Zent, R., & Fässler, R. (2009) The tail of integrins, talin, and kindlins.
Science 324, 895-9.
67. Moser, M., Nieswandt, B., Ussar, S., Pozgajova, M., & Fässler, R. (2008) Kindlin-3 is essential
for integrin activation and platelet aggregation. Nat Med 14, 325-30.
68. Ng, T., Shima, D., Squire, A., Bastiaens, P. I., Gschmeissner, S., Humphries, M. J., & Parker, P.
J. (1999) PKCalpha regulates beta1 integrin-dependent cell motility through association and
control of integrin traffic. EMBO J 18, 3909-23.
69. Nieswandt, B., Moser, M., Pleines, I., Varga-Szabo, D., Monkley, S., Critchley, D., & Fässler, R.
(2007) Loss of talin1 in platelets abrogates integrin activation, platelet aggregation, and thrombus
formation in vitro and in vivo. J Exp Med 204, 3113-8.
31
70. Nieves, B., Jones, C. W., Ward, R., Ohta, Y., Reverte, C. G., & LaFlamme, S. E. (2010) The
NPIY motif in the integrin beta1 tail dictates the requirement for talin-1 in outside-in signaling. J
Cell Sci 123, 1216-26.
71. Nikolopoulos, S. N. & Turner, C. E. (2001) Integrin-linked kinase (ILK) binding to paxillin LD1
motif regulates ILK localization to focal adhesions. J Biol Chem 276, 23499-505.
72. Nilsson, S., Kaniowska, D., Brakebusch, C., Fässler, R., & Johansson, S. (2006) Threonine 788
in integrin subunit beta1 regulates integrin activation. Exp Cell Res 312, 844-53.
73. Nishida, N., Xie, C., Shimaoka, M., Cheng, Y., Walz, T., & Springer, T. A. (2006) Activation of
leukocyte beta2 integrins by conversion from bent to extended conformations. Immunity 25, 58394.
74. Nishimura, T. & Kaibuchi, K. (2007) Numb controls integrin endocytosis for directional cell
migration with aPKC and PAR-3. Dev Cell 13, 15-28.
75. Nurden, A. T. (2006) Glanzmann thrombasthenia. Orphanet J Rare Dis 1, 10.
76. O'Toole, T. E., Katagiri, Y., Faull, R. J., Peter, K., Tamura, R., Quaranta, V., Loftus, J. C.,
Shattil, S. J., & Ginsberg, M. H. (1994) Integrin cytoplasmic domains mediate inside-out signal
transduction. J Cell Biol 124, 1047-59.
77. O'Toole, T. E., Mandelman, D., Forsyth, J., Shattil, S. J., Plow, E. F., & Ginsberg, M. H. (1991)
Modulation of the affinity of integrin alpha IIb beta 3 (GPIIb-IIIa) by the cytoplasmic domain of
alpha IIb. Science 254, 845-7.
78. O'Toole, T. E., Ylanne, J., & Culley, B. M. (1995) Regulation of integrin affinity states through
an NPXY motif in the beta subunit cytoplasmic domain. J Biol Chem 270, 8553-8.
79. Pankov, R., Cukierman, E., Katz, B. Z., Matsumoto, K., Lin, D. C., Lin, S., Hahn, C., &
Yamada, K. M. (2000) Integrin dynamics and matrix assembly: tensin-dependent translocation of
alpha(5)beta(1) integrins promotes early fibronectin fibrillogenesis. J Cell Biol 148, 1075-90.
80. Patil, S., Jedsadayanmata, A., Wencel-Drake, J. D., Wang, W., Knezevic, I., & Lam, S. C. (1999)
Identification of a talin-binding site in the integrin beta(3) subunit distinct from the NPLY
regulatory motif of post-ligand binding functions. The talin n-terminal head domain interacts
with the membrane-proximal region of the beta(3) cytoplasmic tail. J Biol Chem 274, 28575-83.
81. Pearse, B. M., Smith, C. J., & Owen, D. J. (2000) Clathrin coat construction in endocytosis. Curr
Opin Struct Biol 10, 220-8.
82. Pellinen, T. & Ivaska, J. (2006) Integrin traffic. J Cell Sci 119, 3723-31.
83. Pellinen, T., Tuomi, S., Arjonen, A., Wolf, M., Edgren, H., Meyer, H., Grosse, R., Kitzing, T.,
Rantala, J. K., Kallioniemi, O., Fässler, R., Kallio, M., & Ivaska, J. (2008) Integrin trafficking
regulated by Rab21 is necessary for cytokinesis. Dev Cell 15, 371-85.
84. Petrich, B. G., Marchese, P., Ruggeri, Z. M., Spiess, S., Weichert, R. A. M., Ye, F., Tiedt, R.,
Skoda, R. C., Monkley, S. J., Critchley, D. R., & Ginsberg, M. H. (2007) Talin is required for
integrin-mediated platelet function in hemostasis and thrombosis. J Exp Med 204, 3103-11.
85. Petrie, R. J., Doyle, A. D., & Yamada, K. M. (2009) Random versus directionally persistent cell
migration. Nat Rev Mol Cell Biol 10, 538-49.
86. Peyruchaud, O., Nurden, A. T., Milet, S., Macchi, L., Pannochia, A., Bray, P. F., Kieffer, N., &
Bourre, F. (1998) R to Q amino acid substitution in the GFFKR sequence of the cytoplasmic
domain of the integrin IIb subunit in a patient with a Glanzmann's thrombasthenia-like syndrome.
Blood 92, 4178-87.
87. Pfaff, M., Liu, S., Erle, D. J., & Ginsberg, M. H. (1998) Integrin beta cytoplasmic domains
differentially bind to cytoskeletal proteins. J Biol Chem 273, 6104-9.
32
88. Piwko-Czuchra, A., Koegel, H., Meyer, H., Bauer, M., Werner, S., Brakebusch, C., & Fässler, R.
(2009) Beta1 integrin-mediated adhesion signalling is essential for epidermal progenitor cell
expansion. PLoS One 4, e5488.
89. Proux-Gillardeaux, V., Gavard, J., Irinopoulou, T., Mège, R.-M., & Galli, T. (2005) Tetanus
neurotoxin-mediated cleavage of cellubrevin impairs epithelial cell migration and integrindependent cell adhesion. Proc Natl Acad Sci U S A 102, 6362-7.
90. Prunier, C. & Howe, P. H. (2005) Disabled-2 (Dab2) is required for transforming growth factor
beta-induced epithelial to mesenchymal transition (EMT). J Biol Chem 280, 17540-8.
91. Puthenveedu, M. A. & von Zastrow, M. (2006) Cargo regulates clathrin-coated pit dynamics.
Cell 127, 113-24.
92. Qu, H., Tu, Y., Shi, X., Larjava, H., Saleem, M. A., Shattil, S. J., Fukuda, K., Qin, J., Kretzler,
M., & Wu, C. (2011) Kindlin-2 regulates podocyte adhesion and fibronectin matrix deposition
through interactions with phosphoinositides and integrins. J Cell Sci 124, 879-91.
93. Rabb, H., Michishita, M., Sharma, C. P., Brown, D., & Arnaout, M. A. (1993) Cytoplasmic tails
of human complement receptor type 3 (CR3, CD11b/CD18) regulate ligand avidity and the
internalization of occupied receptors. J Immunol 151, 990-1002.
94. Raghavan, S., Bauer, C., Mundschau, G., Li, Q., & Fuchs, E. (2000) Conditional ablation of
beta1 integrin in skin. Severe defects in epidermal proliferation, basement membrane formation,
and hair follicle invagination. J Cell Biol 150, 1149-60.
95. Rappoport, J. Z. & Simon, S. M. (2003) Real-time analysis of clathrin-mediated endocytosis
during cell migration. J Cell Sci 116, 847-55.
96. Roberts, M., Barry, S., Woods, A., van der Sluijs, P., & Norman, J. (2001) PDGF-regulated rab4dependent recycling of alphavbeta3 integrin from early endosomes is necessary for cell adhesion
and spreading. Curr Biol 11, 1392-402.
97. Romzek, N. C., Harris, E. S., Dell, C. L., Skronek, J., Hasse, E., Reynolds, P. J., Hunt, 3rd, S.
W., & Shimizu, Y. (1998) Use of a beta1 integrin-deficient human T cell to identify beta1
integrin cytoplasmic domain sequences critical for integrin function. Mol Biol Cell 9, 2715-27.
98. Sakai, T., Jove, R., Fässler, R., & Mosher, D. F. (2001) Role of the cytoplasmic tyrosines of beta
1A integrins in transformation by v-src. Proc Natl Acad Sci U S A 98, 3808-13.
99. Sakai, T., Zhang, Q., Fässler, R., & Mosher, D. F. (1998) Modulation of beta1A integrin
functions by tyrosine residues in the beta1 cytoplasmic domain. J Cell Biol 141, 527-38.
100. Schmidt, S., Nakchbandi, I., Ruppert, R., Kawelke, N., Hess, M. W., Pfaller, K., Jurdic, P.,
Fässler, R., & Moser, M. (2011) Kindlin-3-mediated signaling from multiple integrin classes is
required for osteoclast-mediated bone resorption. J Cell Biol 192, 883-97.
101. Sepulveda, J. L., Gkretsi, V., & Wu, C. (2005) Assembly and signaling of adhesion complexes.
Curr Top Dev Biol 68, 183-225.
102. Shattil, S. J. (2009) The beta3 integrin cytoplasmic tail: protein scaffold and control freak. J
Thromb Haemost 7 Suppl 1, 210-3.
103. Shi, F. & Sottile, J. (2008) Caveolin-1-dependent beta1 integrin endocytosis is a critical regulator
of fibronectin turnover. J Cell Sci 121, 2360-71.
104. Shi, X., Ma, Y.-Q., Tu, Y., Chen, K., Wu, S., Fukuda, K., Qin, J., Plow, E. F., & Wu, C. (2007)
The MIG-2/integrin interaction strengthens cell-matrix adhesion and modulates cell motility. J
Biol Chem 282, 20455-66.
105. Siegel, D. H., Ashton, G. H. S., Penagos, H. G., Lee, J. V., Feiler, H. S., Wilhelmsen, K. C.,
South, A. P., Smith, F. J. D., Prescott, A. R., Wessagowit, …, McLean, W. H. I., McGrath, J. A.,
& Epstein, E. H. (2003) Loss of kindlin-1, a human homolog of the Caenorhabditis elegans actin-
33
extracellular-matrix linker protein UNC-112, causes Kindler syndrome. Am J Hum Genet 73,
174-87.
106. Skalski, M., Sharma, N., Williams, K., Kruspe, A., & Coppolino, M. G. (2011) SNAREmediated membrane traffic is required for focal adhesion kinase signaling and Src-regulated
focal adhesion turnover. Biochim Biophys Acta 1813, 148-58.
107. Skalski, M., Yi, Q., Kean, M. J., Myers, D. W., Williams, K. C., Burtnik, A., & Coppolino, M.
G. (2010) Lamellipodium extension and membrane ruffling require different SNARE-mediated
trafficking pathways. BMC Cell Biol 11, 62.
108. Stroeken, P. J., van Rijthoven, E. A., Boer, E., Geerts, D., & Roos, E. (2000) Cytoplasmic
domain mutants of beta1 integrin, expressed in beta 1-knockout lymphoma cells, have distinct
effects on adhesion, invasion and metastasis. Oncogene 19, 1232-8.
109. Subauste, M. C., Nalbant, P., Adamson, E. D., & Hahn, K. M. (2005) Vinculin controls PTEN
protein level by maintaining the interaction of the adherens junction protein beta-catenin with the
scaffolding protein MAGI-2. J Biol Chem 280, 5676-81.
110. Tadokoro, S., Shattil, S. J., Eto, K., Tai, V., Liddington, R. C., de Pereda, J. M., Ginsberg, M. H.,
& Calderwood, D. A. (2003) Talin binding to integrin beta tails: a final common step in integrin
activation. Science 302, 103-6.
111. Takagi, J., Erickson, H. P., & Springer, T. A. (2001) C-terminal opening mimics 'inside-out'
activation of integrin alpha5beta1. Nat Struct Biol 8, 412-6.
112. Takagi, J., Petre, B. M., Walz, T., & Springer, T. A. (2002) Global conformational
rearrangements in integrin extracellular domains in outside-in and inside-out signaling. Cell 110,
599-11.
113. Tanentzapf, G. & Brown, N. H. (2006) An interaction between integrin and the talin FERM
domain mediates integrin activation but not linkage to the cytoskeleton. Nat Cell Biol 8, 601-6.
114. Teckchandani, A., Toida, N., Goodchild, J., Henderson, C., Watts, J., Wollscheid, B., & Cooper,
J. A. (2009) Quantitative proteomics identifies a Dab2/integrin module regulating cell migration.
J Cell Biol 186, 99-111.
115. Tu, Y., Wu, S., Shi, X., Chen, K., & Wu, C. (2003) Migfilin and Mig-2 link focal adhesions to
filamin and the actin cytoskeleton and function in cell shape modulation. Cell 113, 37-47.
116. Upla, P., Marjomäki, V., Kankaanpää, P., Ivaska, J., Hyypiä, T., Van Der Goot, F. G., & Heino,
J. (2004) Clustering induces a lateral redistribution of alpha 2 beta 1 integrin from membrane
rafts to caveolae and subsequent protein kinase C-dependent internalization. Mol Biol Cell 15,
625-36.
117. Ussar, S., Moser, M., Widmaier, M., Rognoni, E., Harrer, C., Genzel-Boroviczeny, O., &
Fässler, R. (2008) Loss of Kindlin-1 causes skin atrophy and lethal neonatal intestinal epithelial
dysfunction. PLoS Genet 4, e1000289.
118. Ussar, S., Wang, H.-V., Linder, S., Fässler, R., & Moser, M. (2006) The Kindlins: subcellular
localization and expression during murine development. Exp Cell Res 312, 3142-51.
119. Valmu, L. & Gahmberg, C. G. (1995) Treatment with okadaic acid reveals strong threonine
phosphorylation of CD18 after activation of CD11/CD18 leukocyte integrins with phorbol esters
or CD3 antibodies. J Immunol 155, 1175-83.
120. Van Willigen, G., Hers, I., Gorter, G., & Akkerman, J. W. (1996) Exposure of ligand-binding
sites on platelet integrin alpha IIB/beta 3 by phosphorylation of the beta 3 subunit. Biochem J
314, 769-79.
121. Van der Flier, A. & Sonnenberg, A. (2001) Function and interactions of integrins. Cell Tissue
Res 305, 285-98.
34
122. Vignoud, L., Albigès-Rizo, C., Frachet, P., & Block, M. R. (1997) NPXY motifs control the
recruitment of the alpha5beta1 integrin in focal adhesions independently of the association of
talin with the beta1 chain. J Cell Sci 110 ( Pt 12), 1421-30.
123. Vignoud, L., Usson, Y., Balzac, F., Tarone, G., & Block, M. R. (1994) Internalization of the
alpha 5 beta 1 integrin does not depend on "NPXY" signals. Biochem Biophys Res Commun 199,
603-11.
124. Vinogradova, O., Velyvis, A., Velyviene, A., Hu, B., Haas, T., Plow, E., & Qin, J. (2002) A
structural mechanism of integrin alpha(IIb)beta(3) "inside-out" activation as regulated by its
cytoplasmic face. Cell 110, 587-97.
125. Wegener, K. L., Partridge, A. W., Han, J., Pickford, A. R., Liddington, R. C., Ginsberg, M. H., &
Campbell, I. D. (2007) Structural basis of integrin activation by talin. Cell 128, 171-82.
126. White, D. P., Caswell, P. T., & Norman, J. C. (2007) alpha v beta3 and alpha5beta1 integrin
recycling pathways dictate downstream Rho kinase signaling to regulate persistent cell
migration. J Cell Biol 177, 515-25.
127. Woods, A. J., White, D. P., Caswell, P. T., & Norman, J. C. (2004) PKD1/PKCmu promotes
alphavbeta3 integrin recycling and delivery to nascent focal adhesions. EMBO J 23, 2531-43.
128. Xiong, J. P., Stehle, T., Diefenbach, B., Zhang, R., Dunker, R., Scott, D. L., Joachimiak, A.,
Goodman, S. L., & Arnaout, M. A. (2001) Crystal structure of the extracellular segment of
integrin alpha Vbeta3. Science 294, 339-45.
129. Xu, H., Boulianne, G. L., & Trimble, W. S. (2002) Membrane trafficking in cytokinesis. Semin
Cell Dev Biol 13, 77-82.
130. Ye, F., Hu, G., Taylor, D., Ratnikov, B., Bobkov, A. A., McLean, M. A., Sligar, S. G., Taylor, K.
A., & Ginsberg, M. H. (2010) Recreation of the terminal events in physiological integrin
activation. J Cell Biol 188, 157-73.
131. Ylänne, J., Huuskonen, J., O'Toole, T. E., Ginsberg, M. H., Virtanen, I., & Gahmberg, C. G.
(1995) Mutation of the cytoplasmic domain of the integrin beta 3 subunit. Differential effects on
cell spreading, recruitment to adhesion plaques, endocytosis, and phagocytosis. J Biol Chem 270,
9550-7.
132. Zamir, E. & Geiger, B. (2001) Molecular complexity and dynamics of cell-matrix adhesions. J
Cell Sci 114, 3583-90.
133. Zamir, E., Katz, M., Posen, Y., Erez, N., Yamada, K. M., Katz, B. Z., Lin, S., Lin, D. C.,
Bershadsky, A., Kam, Z., & Geiger, B. (2000) Dynamics and segregation of cell-matrix
adhesions in cultured fibroblasts. Nat Cell Biol 2, 191-6.
134. Zhang, X., Jiang, G., Cai, Y., Monkley, S. J., Critchley, D. R., & Sheetz, M. P. (2008) Talin
depletion reveals independence of initial cell spreading from integrin activation and traction. Nat
Cell Biol 10, 1062-8.
135. Zhang, Y., Chen, K., Tu, Y., Velyvis, A., Yang, Y., Qin, J., & Wu, C. (2002) Assembly of the
PINCH-ILK-CH-ILKBP complex precedes and is essential for localization of each component to
cell-matrix adhesion sites. J Cell Sci 115, 4777-86.
136. Ziegler, W. H., Gingras, A. R., Critchley, D. R., & Emsley, J. (2008) Integrin connections to the
cytoskeleton through talin and vinculin. Biochem Soc Trans 36, 235-9.
137. Ziegler, W. H., Liddington, R. C., & Critchley, D. R. (2006) The structure and regulation of
vinculin. Trends Cell Biol 16, 453-60.
35