scispace - formally typeset
Open AccessJournal ArticleDOI

Rapid leukocyte migration by integrin-independent flowing and squeezing

Reads0
Chats0
TLDR
It is shown here that functional integrins do not contribute to migration in three-dimensional environments, and these cells migrate by the sole force of actin-network expansion, which promotes protrusive flowing of the leading edge.
Abstract
All metazoan cells carry transmembrane receptors of the integrin family, which couple the contractile force of the actomyosin cytoskeleton to the extracellular environment In agreement with this principle, rapidly migrating leukocytes use integrin-mediated adhesion when moving over two-dimensional surfaces As migration on two-dimensional substrates naturally overemphasizes the role of adhesion, the contribution of integrins during three-dimensional movement of leukocytes within tissues has remained controversial We studied the interplay between adhesive, contractile and protrusive forces during interstitial leukocyte chemotaxis in vivo and in vitro We ablated all integrin heterodimers from murine leukocytes, and show here that functional integrins do not contribute to migration in three-dimensional environments Instead, these cells migrate by the sole force of actin-network expansion, which promotes protrusive flowing of the leading edge Myosin II-dependent contraction is only required on passage through narrow gaps, where a squeezing contraction of the trailing edge propels the rigid nucleus

read more

Content maybe subject to copyright    Report

ARTICLES
Rapid leukocyte migration by integrin-
independent flowing and squeezing
Tim La
¨
mmermann
1
, Bernhard L. Bader
3
, Susan J. Monkley
4
, Tim Worbs
5
, Roland Wedlich-So
¨
ldner
2
, Karin Hirsch
1
,
Markus Keller
3
, Reinhold Fo
¨
rster
5
, David R. Critchley
4
, Reinhard Fa
¨
ssler
1
& Michael Sixt
1
All metazoan cells carry transmembrane receptors of the integrin family, which couple the contractile force of the
actomyosin cytoskeleton to the extracellular environment. In agreement with this principle, rapidly migrating leukocytes use
integrin-mediated adhesion when moving over two-dimensional surfaces. As migration on two-dimensional substrates
naturally overemphasizes the role of adhesion, the contribution of integrins during three-dimensional movement of
leukocytes within tissues has remained controversial. We studied the interplay between adhesive, contractile and protrus ive
forces during interstitial leukocyte chemotaxis in vivo and in vitro. We ablated all integrin heterodimers from murine
leukocytes, and show here that functional integrins do not contribute to migration in three-dimensional environments.
Instead, these cells migrate by the sole force of actin-network expan sion, which promotes protrusive flowing of the leading
edge. Myosin II-dependent contraction is only required on passage through narrow gaps, where a squeezing contraction of
the trailing edge propels the rigid nucleus.
The current model of metazoan cell migration is frequently described
as a multistep cycle: F-actin polymerization at the cell front pushes
out a membrane protrusion that subsequently becomes anchored to
an extracellular substrate by transmembrane receptors of the integrin
family. Integrins are dynamically coupled to the cytoskeleton and
transduce the internal force that is generated when myosin II con-
tracts the actin network. Contraction imposes retrograde pulling
forces on the integrins, which in turn facilitates forward locomotion
of the cell body
1–3
. The mammalian integrin family consists of 24
different functional heterodimers with individual binding specifici-
ties for cellular and extracellular ligands
4
. The integrin repertoire of
each cell type defines which substrate it can use for this ‘haptokinetic’
(adhesion driven) mode of migration. Such intimate linkage between
substrate-specific adhesion and migration restricts the migrating
cells to preformed pathways, and thereby creates the determinism
that is essential for many of the precise cell trafficking and positioning
processes underlying compartmentalization and patterning during
development and regeneration. As all mammalian cells with the
exception of erythrocytes carry integrins on their surface, it is rea-
sonable to assume that haptokinesis is a universal phenomenon
4,5
.
Leukocytes are outstanding cells, as they are scattered throughout
the body and have the potential to infiltrate any type of tissue. Rather
than following exact routes and being restricted to specific compart-
ments, leukocytes frequently undergo stochastic swarming with single
cell migration velocities that are up to 100 times faster than mesench-
ymal and epithelial cell types
6,7
. What makes leukocytes so quick and
flexible? In contrast to slow cells, migrating leukocytes undergo fre-
quent shape changes and were therefore morphologically described as
‘‘amoeboid’’
8
. How these shape changes relate to actual locomotion is
poorly investigated, and it is currently unknown if amoeboid migra-
tion represents a specialized strategy or just an accelerated variant of
the above introduced migration paradigm that is only well established
for slow cells. We therefore studied the inter-dependency of adhesion,
protrusion and contraction in inflammatory cells. We demonstrate
that integrin-mediated adhesion is only necessary to overcome tissue
barriers like the endothelial layer, while interstitial migration is auto-
nomous from the molecular composition of the extracellular environ-
ment. Such adhesion-independent migration is driven by protrusive
flowing of the anterior actin network of the cell and supported by
squeezing actomyosin contractions of the trailing edge to propel the
rigid nucleus through narrow spaces.
Dendritic cells migrate without integrins
As a model system to study interstitial leukocyte migration in vivo,we
focused on dendritic cells (DCs), phagocytes that reside within peri-
pheral tissues such as skin. They become activated upon wounding or
infection, sample antigens and quickly migrate via the afferent lymph-
atic vessels into the draining lymph node where they act as antigen
presenting cells
9
. DC migration is primarily guided by the two chemo-
kines CCL19 and CCL21; these are expressed in lymphatic endothe-
lium and the lymph nodes’ T-cell area, and bind the CC-chemokine
receptor 7 (CCR7), which is upregulated on DCs upon activation
10
.
Flow cytometric analysis of DCs generated from mouse bone-
marrow-derived stem cells revealed expression of the b
1
, b
2
, b
7
and
a
v
integrin families (Fig. 1a). To investigate the role integrins play in
DC migration in vivo, we used a combinatorial mouse genetics
approach to delete all integrin heterodimers from the surface of
DCs, thereby generating phenotypically normal pan-integrin-
deficient (integrin
2/2
) DCs (Fig. 1a, see Supplementary Material,
Methods, and Supplementary Figs 1, 2, 4c). We co-injected 1:1 mix-
tures of differentially labelled integrin
2/2
and wild-type DCs into the
dermis of mouse footpads and quantified their arrival in T-cell areas
of the draining lymph nodes. Surprisingly, wild-type and integrin
2/2
DCs arrived and localized in the T-cell area in an indistinguishable
manner (Fig. 1b, c, Supplementary Fig. 3), whereas CCR7
2/2
DCs,
which cannot interpret the directional information, failed to enter
the lymph node (Fig. 1c) as previously described
10
. We next targeted
integrin functionality indirectly by deleting the talin1 gene in DCs
1
Department of Molecular Medicine,
2
Junior Research Group Cellular Dynamics and Cell Patterning, Max Planck Institute of Biochemistry, 82152 Martinsried, Germany.
3
Department
of Nutritional Medicine, Technische Universita
¨
tMu¨nchen, Munich, 85350 Freising, Germany.
4
Department of Biochemistry, University of Leicester, Leicester LE1 9HN, UK.
5
Institute
of Immunology, Hannover Medical School, 30625 Hannover, Germany.
Vol 453
|
1 May 2008
|
doi:10.1038/nature06887
51
Nature
Publishing
Group
©2008

(Supplementary Fig. 4a). The interaction of talin1 with integrin
cytoplasmic domains is required for integrin activation, ligand bind-
ing and coupling of F-actin to adhesion sites
11
. Accordingly, talin1
gene ablation in DCs did not affect the levels of integrin expression,
but completely abolished their ability to bind ligands (Supple-
mentary Fig. 4b–d). Again, arrival of talin1
2/2
DCs in lymph nodes
did not significantly differ from wild-type DCs (Fig. 1c).
To allow for a detailed comparison of the cellular dynamics of
wild-type and integrin
2/2
DCs within the different physiological
environments of their migration route, we employed two in situ
imaging approaches (Fig. 2). Representing the starting point of DC
migration, the interstitial space of the dermis is dominated by fibrillar
arrays of collagen bundles
12
. Using a newly developed set-up of ex vivo
live cell imaging within explanted ear dermis, we could directly visu-
alize DCs moving towards and entering the afferent lymphatic vessels
by wide-field microscopy (Fig. 2b, c, Supplementary Fig. 5). Single
cell tracking revealed no difference in the migratory behaviour of
wild-type, integrin
2/2
and talin1
2/2
DCs (Fig. 2c, d).
The lymph node constitutes a fundamentallydifferent environment
for cellular migration, as it contains almost no freely accessible extra-
cellular matrix molecules but is densely packed with lymphocytes and
stroma cells
13,14
. To analyse the dynamics of intranodal DC migration
from the subcapsular sinus towards the T-cell area, we employed a
set-up of intravital two-photon microscopy of the popliteal lymph
nodes
15
. Evaluating a range of motility parameters, including cell
velocity, directional persistence, mean square displacement, motility
coefficients and morphology, we found that the movement behaviour
of wild-type and integrin
2/2
DCs was indistinguishable (Fig. 2e, f,
Supplementary Videos 1–3, data not shown).
2D but not 3D migration is integrin-dependent
On their way into the lymph node, DCs do not cross significant tissue
barriers, such as continuous endothelial or epithelial linings
12
.As
these processes have been shown to be integrin-dependent
16
, we also
tested for extravasation of integrin
2/2
leukocytes from the blood
stream into the inflamed dermis, and found that it was indeed abo-
lished (Fig. 2g, Supplementary Fig. 6). This finding corroborates the
extravasation model in which integrin-mediated tight immobiliza-
tion of leukocytes to the luminal endothelial surface is necessary to
counteract the shear forces imposed by the blood flow
16
. In line with
this in vivo finding, we found that integrin
2/2
DCs were entirely
unable to adhere to and migrate on two-dimensional (2D) substrates.
Gravity was obviously not sufficient to confine the non-adherent cells
to planar surfaces and allow the transduction of traction forces
b
a
1.0
0.5
0.0
48 h
WT DC
Itg
–/–
DC
Laminin
CCR7
–/–
Itg
–/–
DC
WT
DC
β
1
β
2
β
3
β
7
α
4
α
5
α
L
α
M
α
v
α
X
Itg
–/–
Tln1
–/–
32 h 48 h 48 h32 h
c
Ratio of –/– DC in
T cortex
B
B
T
B
Figure 1
|
Migration of integrin-deficient dendritic cells into lymph nodes.
a
, Flow cytometric analysis of integrin subunits of wild-type (WT) and
integrin-deficient (Itg
2/2
) dendritic cells (DCs) (integrin, unshaded; isotype
control, grey-shaded).
b, c, In vivo migration of DCs after subcutaneous
injection into hind footpads of WT mice. At the time points indicated,
arrival of differentially labelled DCs in the lymph node was analysed.
b, Localization of DCs after 48 h. Composite of six separate images of a single
lymph node. Scale bar, 200 mm; B, B-cell follicle; T, T-cell cortex.
c, Quantitative histology. Dotted line at 0.5, 50% of DCs that migrated into
the lymph node are knockout DCs; Tln1, talin1. Red dots indicate single
experiments (1 lymph node); box shows median, 25%, 75% percentile;
whiskers show minimum, maximum.
T
B
B
E
LN
LV
BV
D
LYVE-1
DC
Speed (µm min
–1
)
F
2,303
= 2.1, P = 0.13
WT DC
Itg
–/–
DC
t
129
= 1.35,
P = 0.18
Granulocytes
per mm
2
*
*
H
2
= 13.3, P = 0.001
Interstitial migration within skin and entry into lymphatics
Migration within lymph node
Extravasation from blood vessel
a
g
c
WTItg
–/–
Tln1
–/–
b
d
2
4
6
8
Tln1
–/–
0.76
±0.13
Itg
–/–
0.76
±0.17
WT
0.79
±0.14
DP:
0
0
2,000
4,000
6,000
8,000
Tln1
–/–
Itg
–/–
WT
e
2
4
6
8
Speed
(µm min
–1
)
f
0.41
±0.15
Itg
–/–
0.41
±0.18
WT
DP:
0
DCLYVE-1 LYVE-1/DC
Figure 2
|
Integrin-independent
interstitial leukocyte migration
in vivo.a
, Scheme of the dendritic
cell (DC) migration path from the
skin via lymphatic vessels to the
lymph node; B, B-cell follicle; BV,
blood vessel; D, dermis; E, epidermis;
LN, lymph node; LV, lymphatic
vessel; T, T-cell cortex.
bd, Migration of DCs within the
dermis of ear explants.
b, Tracks of
DCs (red) entering LYVE-1
1
LVs
(green).
c,Confocalimagesofwhole
mount ear explants 2 h after addition
of DCs (red). DCs locate within the
lumen of LVs; all scale bars, 50 mm.
d, Speed and directional persistence
(DP, mean 6 s.d.) of single cells (dots)
for wild-type (WT), integrin
2/2
(Itg
2/2
)andtalin1
2/2
(Tln1
2/2
)
DCs. Red line, mean.
e, f, Intravital
two-photon microscopy.
e,3D
tracking of WT (green, blue line) and
Itg
2/2
(red, yellow line) DCs
migrating in the interfollicular areas.
f, Quantification of the average speed
and DP (mean 6 s.d.) of tracked
single cells (dots). Red line, mean.
g, Quantification of granulocytes
that extravasated from blood into ear
dermis (n 5 6 for each mouse line).
Red line, median. *P , 0.05 (post hoc).
ARTICLES NATURE
|
Vol 453
|
1 May 2008
52
Nature
Publishing
Group
©2008

(Fig. 3a, b, Supplementary Fig. 4d, Supplementary Video 4). We
conclude that extravasation requires integrin-mediated adhesion.
To better mimic the interstitial microenvironment, we established
chemotaxis assays within artificial three-dimensional (3D) matrices
of fibrin (a ligand for b
2
and b
3
integrins) and collagen I (a ligand for
several members of the b
1
integrin subfamily). In this system, DCs
persistently migrated along soluble gradients of CCL19 and showed
amoeboid morphology and velocities that were comparable to our
(and previously published
17
) in vivo observations. In agreement with
our in vivo data and in contrast to 2D migration, integrin
2/2
and
talin1
2/2
DCs migrated with speed and directional persistence that
were indistinguishable from wild-type cells (Fig. 3c, Supplementary
Fig. 7a, Supplementary Video 5). We also observed integrin-
independent migration for chemotaxing B cells and granulocytes,
suggesting a broader prevalence of this functional principle (Fig. 3d,
Supplementary Fig. 7b).
Functional dissociation of front and back
To mechanistically understand how cells can move in the absence of
integrin-mediated traction forces, we compared the integrin-
dependent movement pattern of fibroblasts with that of leukocytes
and found a striking difference. Leading edge protrusions of
fibroblasts translocated collagen fibres towards the cell body, indi-
cating the presence of rearward pulling forces and demonstrating
that protrusion, adhesion and contraction are tightly coupled in this
cell type (Supplementary Video 6)
18,19
. In contrast, chemotaxing DCs
protruded without signs of anterior pulling forces while the trailing
edge displayed an irregularly alternating contractile pattern. Con-
traction was characterized by shrinkage of the cell rear with con-
comitant forward-streaming of cytoplasmic matter (Supplementary
Video 7). When we dynamically visualized activated myosin II, the
motor system mediating contraction
1,2,20
, by time-lapse imaging of
DCs expressing a myosin light chain–GFP fusion protein, we found
accumulation at the cell rear during contractile phases (Fig. 4a,
Supplementary Fig. 8a, Supplementary Video 8). During non-
contractile phases, the trailing edge remained motionless and
appeared to be passively dragged by the protruding cell front
(Supplementary Video 7). These observations suggest that in
contrast to slow moving cells, protrusion and contraction are
spatio-temporally dissociated in leukocytes.
Receptor-mediated force transduction can only support forward
movement if retrograde (contractile) but not anterograde (protrus-
ive) forces are coupled to the environment
2,3
. To address how
actomyosin contraction functionally contributes to leukocyte loco-
motion, we pharmacologically inhibited myosin II or its upstream-
activator, Rho kinase. Irrespective of the presence of integrin
function, either treatment severely reduced migration speeds of
DCs, granulocytes and B cells (Fig. 4b, Supplementary Fig. 8e, data
not shown). Nevertheless, the chemotactic gradient was still able to
polarize the cell population. The leading edges of cells remained
dynamic and protruded with normal speed towards the chemokine
source. However, the cells were unable to move their trailing edges
ICAM-1
No. of adherent cells
Speed (µm min
–1
)
ICAM-1
H
2
=12.3
P<0.001
H
2
=16.3
P<0.001
F
2,9
=2.8, P=0.12
t
64
=0.6,
P=0.53
t
65
=1.6,
P=0.11
*
*
*
*
FN
WT
0
400
800
2D 3D
a
bc d
2
0
DP:
4
8
10
6
Speed (µm min
–1
)
4
0
12
16
8
Itg
–/–
Tln1
–/–
WT
WT
0.81
± 0.09
0.78
± 0.08
0.76
± 0.1
Itg
–/–
Tln1
–/–
WT
Itg
–/–
Tln1
–/–
Itg
–/–
WT
Granulocyte B blast
Itg
–/–
WT Itg
–/–
Tln1
–/–
Figure 3
|
Integrin-independent leukocyte migration in 3D networks
in vitro.a
, 2D adhesion versus 3D migration of DCs. b, 2D adhesion on
ICAM-1 and fibronectin (FN) 1 h after LPS stimulation. Top, quantification
of the number of adherent cells (ICAM-1, n 5 6; FN, n 5 8 for each
experiment). Bars, median values with interquartile range, *P , 0.05 (post
hoc). Bottom, morphology of wild-type (WT), integrin
2/2
(Itg
2/2
) and
talin1
2/2
(Tln1
2/2
) DCs. c, d, Velocities of chemotaxing leukocytes in 3D
collagen matrices.
c, Top, DCs (4 experiments per group); d, granulocytes, B
cells. Dots, single cells; red line, mean.
c, Bottom, single cell tracks of
chemotaxing DCs; values indicate mean DP 6 s.d.
CCL19
CCL19
CCL19 CCL19 CCL19 CCL19
Bleb
H
2
= 184.2, P < 0.001
T = 378
P < 0.001
T = 300
P < 0.001
*
*
R
e
ar
Front
R
e
ar
Front
Bleb
Speed (µm min
–1
)
Blebbistatin
WTItg
–/–
Blebbistatin
Y27632 Y27632
8
6
4
2
0
Speed (µm min
–1
)
6
4
2
0
a
b
d
c
0:30 0:35 0:45
Y27632
Bleb
0:00
0:00
0:00 1:00 3:10 5:00 6:40 8:10 10:00
0:00 1:00 3:10 5:00 6:40 8:10 10:00
1:50 3:00 5:00 6:40 8:10 10:00
3:00 5:00 6:40 8:10 10:001:50
0:55 1:05 1:10
MLC–GFP
intensity
Figure 4
|
Myosin II-dependent nuclear squeezing at the cell rear.
a
, Sequence of a myosin light chain (MLC)–GFP expressing wild-type (WT)
dendritic cell (DC) migrating towards CCL19 in a 3D collagen gel; upper
row, differential interference contrast (DIC) microscopy; lower row,
MLC–GFP intensity profiles (encoded in pseudo-colours with red
representing highest levels). Time in min:s.
b, Left, velocities of single DCs
(dots; red line, median), and right, DC morphology (coloured red, DIC
microscopy), chemotaxing towards CCL19 in 3D collagen gels upon
pharmacological inhibition.
c, Speed of cell bodies (rear) versus cell
protrusions (front) of DCs in the presence of inhibitor; red line, median.
Bleb, blebbistatin; *P , 0.05 (post hoc).
d, e, Time-lapse sequence of a WT
(
d) and integrin
2/2
(e) DC either with (lower row) or without (upper row)
blebbistatin; DIC microscopy, nuclei (green). Time in min:s; scale bars, 5 mm
(
a), 10 mm(b, d, e).
NATURE
|
Vol 453
|
1 May 2008 ARTICLES
53
Nature
Publishing
Group
©2008

(Fig. 4b–e, Supplementary Fig. 8b–d, Supplementary Video 9). This
functional dissociation between front and back caused up to 30-fold
cell elongation, and demonstrates that the leading edge migrates
autonomously and without a need for receptor-mediated coupling
of contractile forces to the extracellular matrix.
Trailing edge contraction propels the nucleus
How does trailing edge contraction contribute to cell body loco-
motion? In fibroblasts and leukocytes moving on 2D substrates,
actomyosin contraction at the back is required to disassemble recep-
tor binding-sites and subsequently retract the membrane
2,21,22
.
Hence, blocking contraction causes membrane tethers at the trailing
edge. Consistent with an adhesion-independent migration mode, we
could not observe membrane tethers in myosin II-inhibited leuko-
cytes migrating in 3D gels. We considered that the elongated pheno-
type with its rounded back was caused by the inability to move an
internal resistance through narrow gaps within the gel. As the nucleus
is the least elastic cellular compartment
23
, we visualized DNA within
chemotaxing DCs and granulocytes. In untreated cells, we observed
continuous shape changes of nuclei, indicating deformation forces,
while upon myosin II inhibition, spherical nuclei were immobilized
at the rear ends of the cells (Figs 4d, e, Supplementary Fig. 8b–d,
Supplementary Videos 10, 11).
Protrusion drives basal locomotion
Although ‘locked’ nuclei caused migration arrest in most cells, a few
still showed locomotion. This residual migration was often charac-
terized by cell body elongation and dragging of the nucleus (Fig. 5a,
Supplementary Video 12). Such non-contractile movement was in
line with our previous observations showing motile phases in the
absence of trailing edge contractions. We hypothesized that such
myosin II-independent migration might occur in areas of the col-
lagen gel where external resistance was low due to increased spacing
of the collagen fibres. To establish the relationship between internal
contractile force and external resistance, we analysed DC chemotaxis
in collagen gels of varying fibre spacing (Fig. 5b). In all collagen
densities, myosin II-inhibited DCs were slower than untreated cells
but importantly, they ‘caught up’ at lowest gel densities (Fig. 5b,
Supplementary Video 13). A comparison of instantaneous velocities
further showed that myosin II inhibition did not prevent DCs from
reaching the same peak values as untreated cells (Supplementary
Fig. 9a–c).
We conclude that, in contrast to the ‘blebbing’ model of amoeboid
cell migration
24
, cortical actin network contraction does not mediate
locomotion itself but rather facilitates a protrusive mode of migra-
tion in confined environments where protrusion alone is unable to
act against counter-forces. So in constricted areas, the cell overcomes
internal and external resistance (rigidity of the nucleus and fibre
density) by contraction (myosin II). Contractile force deforms the
resistance and facilitates a purely protrusive mode of migration.
To challenge the concept of protrusive movement, we treated che-
motaxing DCs with latrunculin B, which interferes with actin poly-
merization by depleting the pool of available functional G-actin
monomers
25
. We found that this treatment reduced cell migration
velocity in a dose-dependent manner (Fig. 5c). At intermediate con-
centrations, the trailing edge remained contractile and the nucleus
was ‘pushed’ to the cell front (Fig. 5d, Supplementary Video 14).
Consistent with a model where protrusion determines the speed of
cell migration while contraction is only activated to overcome
external resistances, the speed reduction of latrunculin B-treated cells
was independent of gel density (Fig. 5b, c).
Discussion
We show that leukocytes migrate in the absence of specific adhesive
interactions with the extracellular environment. This subversion of
the metazoan principle makes them autonomous from the tissue
context, and allows them to quickly and flexibly navigate through
any organ without adaptations to alternating extracellular ligands.
This strategy is in stark contrast to the haptokinetic migration prin-
ciple that leads along preformed pathways and therefore promotes
deterministic positioning. Adhesion independency better suits stoch-
astic movement or chemotaxis where cells randomly migrate or fol-
low soluble cues. Astonishingly, the protrusive flow of F-actin
appears sufficient to drive rapid leukocyte locomotion in environ-
ments with large pore-size. This resembles the locomotion principle
of nematode sperm cells that is entirely driven by treadmilling poly-
mers of major sperm protein
26
. Only in narrow areas do leukocytes
activate the contractile module to squeeze and propel the internal
resistance of the nucleus in a manner resembling neuronal
nucleokinesis
27
. This ‘flowing and squeezing’ migration model
fulfils a central requirement for immune cell movement: the pericel-
lular environment is transiently deformed but never digested
28
or
CCL19
F
2,147
= 18.9, P < 0.001
H
2
= 53.6, P < 0.001
H
2
= 3.1, P < 0.22
ab
c
d
Blebbistatin
Blebbistatin 100 nM Lat B 100 nM Lat B
Speed (µm min
–1
)
Speed (µm min
–1
)CCL19
00
718
14 34
25 45
33 49
38 58
46 61
51
6
4
2
0
0:00 1:50 2:40 3:50 5:40 6:20 7:40
6
4
2
0
6
4
2
0
6
4
2
0
0 100
H
2
= 120.1, P < 0.001
*
*
*
*
*
*
ns
0.75 1.5
Collagen (mg ml
–1
)
3
*
*
Latrunculin B (nM)
500
64
CCL19
Figure 5
|
Myosin II-independent protrusive migration of dendritic cells.
a
, Two types of residual migration of wild-type (WT) dendritic cells (DCs)
treated with blebbistatin in a standard 3D collagen gel. DCs either drag the
nucleus behind with elongated appearance and low speed (right column) or
appear morphologically normal with high speed (left column). Time in min.
b, Top panel, confocal reflection microscopy of 3D collagen networks of
different densities. Lower panels, velocities of single chemotaxing WT DCs
(dots) in collagen gels with varying densities. Upper graph, no inhibitor; red
line, mean. Middle graph, blebbistatin; red line, median. Bottom graph,
100 nM latrunculin B (partial actin depolymerization); red line, median.
c, Velocity of single WT DCs (dots) in the presence of differentconcentrations
of latrunculin B (1.6 mg ml
21
collagen); red line,median. In b and c, *P , 0.05
(post hoc).
d, Time-lapse sequence of a 100 nM latrunculin B-treated WT DC;
DIC microscopy, nuclei(green). Time, min:s. Blue boxes(in
b and c) highlight
experimental conditions that were used for follow-up experiments indicated
by the blue arrow. Scale bars: 10 mm(
a, d), 50 mm(b).
ARTICLES NATURE
|
Vol 453
|
1 May 2008
54
Nature
Publishing
Group
©2008

otherwise permanently remodelled which avoids ‘collateral damage’,
caused by infiltrating cells.
The fact that leukocytes are able to move autonomously generates
a new level of regulatory possibilities. Because surface bound chemo-
kines and other immobilized extracellular signals do trigger integrin
affinity
16
(unlike soluble chemokines), leukocyte integrins should no
longer be viewed as force transducers during locomotion but as
switchable immobilizing anchors that stop, slow down or confine
high intrinsic motility to specifically assigned surfaces
29,30
. The role
of integrins is therefore mostly to mediate retention, invasion, cell–
cell communication and cell–cell adhesion
31
.
METHODS SUMMARY
Generation of integrin-deficient leukocytes. Integrins and talin1 were targeted
by generating mice with the genotype a
v
flox/flox
(Supplementary Fig. 1), b
1
flox/flox
(ref. 32), b
2
2/2
(ref. 33), b
7
2/2
(ref. 34), Mx1Cre
1/2
(ref. 35) and talin1
flox/flox
(ref. 36), Mx1Cre
1/2
, respectively. Cre expression in the haematopoietic system
was induced by intraperitoneal injection of 250 mg Poly (I)?Poly (C) (Amersham
Biosciences). 10–14 d after knockout induction DCs were generated from bone
marrow suspension and matured with lipopolysaccharide (LPS). The DC culture
was depleted for granulocytes and remaining integrin-positive contaminants by
magnetic sorting (Miltenyi Biotech). DCs used for migration assays were .99%
enriched for b
1
and a
v
integrin knockout cells.
In situ live cell imaging. For dermal ex vivo microscopy, mouse ears were
mechanically separated in dorsal and ventral halves, fluorescently stained with
LYVE-1 antibody and immobilized with the epidermal side down. The dermis
was overlaid with fluorescently labelled DCs and time lapse movies were
recorded with an automated Leica MZ 16 FA stereomicroscope (Visitron
Systems).
In vitro 3D chemotaxis assays. Cells were suspended in PureCol (INAMED) and
cast in custom built migration chambers (standard collagen concentration:
1.6 mg ml
21
). After polymerization, gels were overlaid with culture medium
containing the chemoattractant (CCL19, C5a and CXCL13 for DCs, granulo-
cytes and B cells, respectively) and subsequently imaged using wide-field fluor-
escence video microscopes with differential interference contrast (Zeiss). For
nucleus visualization, SYTO-dyes (Invitrogen) were used. For visualization of
myosin light chain, DCs were nucleofected using Amaxa technology. Inhibitors
were titrated and used 50 mm blebbistatin (Sigma) and 30 mm Y27632
(Calbiochem).
Statistical analysis. t-tests and analysis of variance (ANOVA) were performed
after data were confirmed to fulfil the criteria of normal distribution and equal
variance, otherwise Kruskal–Wallis tests or Mann–Whitney U-tests were
applied. If overall ANOVA or Kruskal–Wallis tests were significant, we per-
formed a post hoc test. Analyses were performed with Sigma Stat 2.03. For
further statistical details, see Supplementary Table.
Full Methods and any associated references are available in the online version of
the paper at www.nature.com/nature.
Received 22 November 2007; accepted 6 March 2008.
1. Giannone, G. et al. Lamellipodial actin mechanically links myosin activity with
adhesion-site formation. Cell 128, 561
575 (2007).
2. Lauffenburger, D. A. & Horwitz, A. F. Cell migration: A physically integrated
molecular process. Cell 84, 359
369 (1996).
3. Mitchison, T. J. & Cramer, L. P. Actin-based cell motility and cell locomotion. Cell
84, 371
379 (1996).
4. Hynes, R. O. Integrins: Bidirectional, allosteric signaling machines. Cell 110,
673
687 (2002).
5. Hynes, R. O. & Zhao, Q. The evolution of cell adhesion. J. Cell Biol. 150, 89
96
(2000).
6. Friedl, P. Prespecification and plasticity: Shifting mechanisms of cell migration.
Curr. Opin. Cell Biol. 16, 14
23 (2004).
7. Pittet, M. J. & Mempel, T. R. Regulation of T-cell migration and effector functions:
Insights from in vivo imaging studies. Immunol. Rev. 221, 107
129 (2008).
8. de Bruyn, P. P. H. The amoeboid movement of the mammalian leukocyte in tissue
culture. Anat. Rec. 95, 117
192 (1946).
9. Reis e Sousa, C. Dendritic cells in a mature age. Nature Rev. Immunol. 6, 476
483
(2006).
10. Fo
¨
rster, R. et al. CCR7 coordinates the primary immune response by establishing
functional microenvironments in secondary lymphoid organs. Cell 99, 23
33 (1999).
11. Calderwood, D. A. & Ginsberg, M. H. Talin forges the links between integrins and
actin. Nature Cell Biol. 5, 694
697 (2003).
12. Stoitzner, P., Pfaller, K., Sto
¨
ssel, H. & Romani, N. A close-up view of migrating
Langerhans cells in the skin. J. Invest. Dermatol. 118, 117
125 ( 2002).
13. Gretz, J. E., Anderson, A. O. & Shaw, S. Cords, channels, corridors and conduits:
Critical architectural elements facilitating cell interactions in the lymph node
cortex. Immunol. Rev. 156, 11
24 (1997).
14. La
¨
mmermann, T. & Sixt, M. The microanatomy of T-cell responses. Immunol. Rev.
221, 26
43 (2008).
15. Mempel, T. R., Henrickson, S. E. & Von Andrian, U. H. T-cell priming by dendritic
cells in lymph nodes occurs in three distinct phases. Nature 427, 154
159 (2004).
16. Alon, R. & Dustin, M. L. Force as a f acilitator of integrin conformational changes
during leukocyte arrest on blood vessels and antigen-presenting cells. Immunity
26, 17
27 (2007).
17. Lindquist, R. L. et al. Visualizing dendritic cell networks in vivo. Nature Immunol. 5,
1243
1250 (2004).
18. Meshel, A. S., Wei, Q., Adelstein, R. S. & Sheetz, M. P. Basic mechanism of three-
dimensional collagen fibre transport by fibroblasts. Nature Cell Biol. 7, 157
164
(2005).
19. Vanni, S., Lagerholm, B. C., Otey, C., Taylor, D. L. & Lanni, F. Internet-based image
analysis quantifies contractile behavior of individual fibroblasts inside model
tissue. Biophys. J. 84, 2715
2727 (2003).
20. Medeiros, N. A., Burnette, D. T. & Forscher, P. Myosin II functions in actin-bundle
turnover in neuronal growth cones. Nature Cell Biol. 8, 215
226 (2006).
21. Hogg, N., Laschinger, M., Giles, K. & McDowall, A. T-cell integrins: More than just
sticking points. J. Cell Sci. 116, 4695
4705 (2003).
22. Morin, N. A. et al. Nonmuscle myosin heavy chain IIA mediates integrin LFA-1 de-
adhesion during T lymphocyte migration. J. Exp. Med. 205, 195
205 (2008).
23. Hu, S., Chen, J., Butler, J. P. & Wang, N. Prestress mediates force propagation into
the nucleus. Biochem. Biophys. Res. Commun. 329, 423
428 (2005).
24. Paluch, E., Sykes, C., Prost, J. & Bornens, M. Dynamic modes of the cortical
actomyosin gel during cell locomotion and division. Trends Cell Biol. 16, 5
10
(2006).
25. Morton, W. M., Ayscough, K. R. & McLaughlin, P. J. Latrunculin alters the actin-
monomer subunit interface to prevent polymerization. Nature Cell Biol. 2,
376
378 (2000).
26. Bottino, D., Mogilner, A., Roberts, T., Stewart, M. & Oster, G. How nematode
sperm crawl. J. Cell Sci. 115, 367
384 (2002).
27. Schaar, B. T. & McConnell, S. K. Cytoskeletal coordination during neuronal
migration. Proc. Natl Acad. Sci. USA 102, 13652
13657 (2005).
28. Wolf, K., Mu¨ller, R., Borgmann, S., Bro
¨
cker, E. B. & Friedl, P. Amoeboid shape
change and contact guidance: T-lymphocyte crawling through fibrillar collagen is
independent of matrix remodeling by MMPs and other proteases. Blood 102,
3262
3269 (2003).
29. Auffray, C. et al. Monitoring of blood vessels and tissues by a population of
monocytes with patrolling behavior. Science 317, 666
670 (2007).
30. Phillipson, M. et al. Intraluminal crawling of neutrophils to emigration sites: A
molecularly distinct process from adhesion in the recruitment cascade. J. Exp.
Med. 203, 2569
2575 (2006).
31. Sixt, M., Bauer, M., La
¨
mmermann, T. & Fa
¨
ssler, R. Beta1 integrins: Zip codes and
signaling relay for blood cells. Curr. Opin. Cell Biol. 18, 482
490 (2006).
32. Potocnik, A. J., Brakebusch, C. & Fa
¨
ssler, R. Fetal and adult hematopoietic stem
cells require beta1 integrin function for colonizing fetal liver, spleen, and bone
marrow. Immunity 12, 653
663 (2000).
33. Wilson, R. W. et al. Gene targeting yields a CD18-mutant mouse for study of
inflammation. J. Immunol. 151, 1571
1578 (1993).
34. Wagner, N. et al. Critical role for beta7 integrins in formation of the gut-associated
lymphoid tissue. Nature 382, 366
370 (1996).
35. Ku¨hn, R., Schwenk, F., Aguet, M. & Rajewsky, K. Inducible gene targeting in mice.
Science 269, 1427
1429 (1995).
36. Petrich, B. G. et al. Talin is required for integrin-mediated platelet function in
hemostasis and thrombosis. J. Exp. Med. 204, 3103
3111 (2007).
Supplementary Information is linked to the online version of the paper at
www.nature.com/nature.
Acknowledgements We thank S. Cremer for help with statistical analysis, Z. Werb
and P. Friedl for critical reading of the manuscript, and M. Bauer for technical
support. This work was financed by the German Research Foundation (DFG), the
Austrian Science Foundation (FWF) and the Max Planck Society. Work in D.R .C.’s
laboratory was supported by the Wellcome Trust.
Author Contributions T.L. and M.S. designed and performed the experiments and
analysed the data. M.S. wrote the paper. T.W. and R.Fo
¨
. performed intravital
microscopy in lymph nodes. B.L.B. and M.K. generated the integrin a
v
mouse. S.J.M.
and D.R.C. generated the talin1 mouse. R.Fa
¨.
generated the integrin b
1
and the
quadruple integrin knockout mouse and provided general support. K.H. assisted
with experiments. R.W.S. contributed to data analysis and experimental design.
Author Information Reprints and permissions information is available at
www.nature.com/reprints. Correspondence and requests for materials should be
addressed to M.S. (sixt@biochem.mpg.de).
NATURE
|
Vol 453
|
1 May 2008 ARTICLES
55
Nature
Publishing
Group
©2008

Figures
Citations
More filters
Journal ArticleDOI

The extracellular matrix: A dynamic niche in cancer progression

TL;DR: The extracellular matrix (ECM), a complex network of macromolecules with distinctive physical, biochemical, and biomechanical properties, is commonly deregulated and becomes disorganized in diseases such as cancer.
Journal ArticleDOI

Cancer Invasion and the Microenvironment: Plasticity and Reciprocity

TL;DR: The cell-matrix and cell-cell adhesion, protease, and cytokine systems that underlie tissue invasion by cancer cells are described and explained to explain how the reciprocal reprogramming of both the tumor cells and the surrounding tissue structures not only guides invasion, but also generates diverse modes of dissemination.
Journal ArticleDOI

Chemokines and Chemokine Receptors: Positioning Cells for Host Defense and Immunity

TL;DR: This review focuses on recent advances in understanding how the chemokine system orchestrates immune cell migration and positioning at the organismic level in homeostasis, in acute inflammation, and during the generation and regulation of adoptive primary and secondary immune responses in the lymphoid system and peripheral nonlymphoid tissue.
Journal ArticleDOI

Plasticity of cell migration: a multiscale tuning model

TL;DR: Using a multiparameter tuning model, this work describes how dimension, density, stiffness, and orientation of the extracellular matrix together with cell determinants—including cell–cell and cell–matrix adhesion, cytoskeletal polarity and stiffness, etc.—interdependently control migration mode and efficiency.
Journal ArticleDOI

Physical limits of cell migration: Control by ECM space and nuclear deformation and tuning by proteolysis and traction force

TL;DR: The physical limits of cell migration in dense porous environments are dependent upon the available space and the deformability of the nucleus and are modulated by matrix metalloproteinases, integrins and actomyosin function.
References
More filters
Journal ArticleDOI

Integrins: Bidirectional, Allosteric Signaling Machines

TL;DR: Current structural and cell biological data suggest models for how integrins transmit signals between their extracellular ligand binding adhesion sites and their cytoplasmic domains, which link to the cytoskeleton and to signal transduction pathways.
Journal ArticleDOI

Cell Migration: A Physically Integrated Molecular Process

TL;DR: The authors are grateful for financial support from the National Institutes of Health (grants GM23244 and GM53905), and to very helpful comments on the manuscript from Elliot Elson, Vlodya Gelfand, Paul Matsudaira, Julie Theriot, and Sally Zigmond.
Journal ArticleDOI

An advanced culture method for generating large quantities of highly pure dendritic cells from mouse bone marrow.

TL;DR: This method allows by simple means the generation of high numbers of murine DC with very low B cell or granulocyte contaminations, which will be valuable to study DC biology notably at the molecular level.
Journal ArticleDOI

Derivation of completely cell culture-derived mice from early-passage embryonic stem cells

TL;DR: Fully potent early passage R1 cells and the R1-S3 subclone should be very useful not only for ES cell-based genetic manipulations but also in defining optimal in vitro culture conditions for retaining the initial totipotency of ES cells.
Journal ArticleDOI

CCR7 coordinates the primary immune response by establishing functional microenvironments in secondary lymphoid organs.

TL;DR: In this paper, the chemokine receptor CCR7 was identified as an important organizer of the primary immune response in mice, and severely delayed kinetics regarding the antibody response and lack contact sensitivity and delayed type hypersensitivity reactions.
Related Papers (5)
Frequently Asked Questions (15)
Q1. What contributions have the authors mentioned in the paper "Rapid leukocyte migration by integrin- independent flowing and squeezing" ?

The authors studied the interplay between adhesive, contractile and protrusive forces during interstitial leukocyte chemotaxis in vivo and in vitro. The authors ablated all integrin heterodimers from murine leukocytes, and show here that functional integrins do not contribute to migration in three-dimensional environments. 

Because surface bound chemokines and other immobilized extracellular signals do trigger integrin affinity16 (unlike soluble chemokines), leukocyte integrins should no longer be viewed as force transducers during locomotion but as switchable immobilizing anchors that stop, slow down or confine high intrinsic motility to specifically assigned surfaces29,30. 

In fibroblasts and leukocytes moving on 2D substrates, actomyosin contraction at the back is required to disassemble receptor binding-sites and subsequently retract the membrane2,21,22. 

Statistical analysis. t-tests and analysis of variance (ANOVA) were performed after data were confirmed to fulfil the criteria of normal distribution and equal variance, otherwise Kruskal–Wallis tests or Mann–Whitney U-tests were applied. 

The authors considered that the elongated phenotype with its rounded back was caused by the inability to move aninternal resistance through narrow gaps within the gel. 

To address how actomyosin contraction functionally contributes to leukocyte locomotion, the authors pharmacologically inhibited myosin II or its upstreamactivator, Rho kinase. 

Only in narrow areas do leukocytes activate the contractile module to squeeze and propel the internal resistance of the nucleus in a manner resembling neuronal nucleokinesis27. 

This functional dissociation between front and back caused up to 30-fold cell elongation, and demonstrates that the leading edge migrates autonomously and without a need for receptor-mediated coupling of contractile forces to the extracellular matrix. 

To better mimic the interstitial microenvironment, the authors established chemotaxis assays within artificial three-dimensional (3D) matrices of fibrin (a ligand for b2 and b3 integrins) and collagen The author(a ligand for several members of the b1 integrin subfamily). 

W. M., Ayscough, K. R. & McLaughlin, P. J. Latrunculin alters the actinmonomer subunit interface to prevent polymerization. 

This finding corroborates the extravasation model in which integrin-mediated tight immobilization of leukocytes to the luminal endothelial surface is necessary to counteract the shear forces imposed by the blood flow16. 

In all collagen densities, myosin II-inhibited DCs were slower than untreated cells but importantly, they ‘caught up’ at lowest gel densities (Fig. 5b, Supplementary Video 13). 

16. Alon, R. & Dustin, M. L. Force as a facilitator of integrin conformational changes during leukocyte arrest on blood vessels and antigen-presenting cells. 

This subversion of the metazoan principle makes them autonomous from the tissue context, and allows them to quickly and flexibly navigate through any organ without adaptations to alternating extracellular ligands. 

R. et al. CCR7 coordinates the primary immune response by establishing functionalmicroenvironments insecondary lymphoidorgans.