ArticlePDF Available

mDia2 Induces the Actin Scaffold for the Contractile Ring and Stabilizes Its Position during Cytokinesis in NIH 3T3 Cells

Authors:

Abstract and Figures

mDia proteins are mammalian homologues of Drosophila diaphanous and belong to the formin family proteins that catalyze actin nucleation and polymerization. Although formin family proteins of nonmammalian species such as Drosophila diaphanous are essential in cytokinesis, whether and how mDia proteins function in cytokinesis remain unknown. Here we depleted each of the three mDia isoforms in NIH 3T3 cells by RNA interference and examined this issue. Depletion of mDia2 selectively increased the number of binucleate cells, which was corrected by coexpression of RNAi-resistant full-length mDia2. mDia2 accumulates in the cleavage furrow during anaphase to telophase, and concentrates in the midbody at the end of cytokinesis. Depletion of mDia2 induced contraction at aberrant sites of dividing cells, where contractile ring components such as RhoA, myosin, anillin, and phosphorylated ERM accumulated. Treatment with blebbistatin suppressed abnormal contraction, corrected localization of the above components, and revealed that the amount of F-actin at the equatorial region during anaphase/telophase was significantly decreased with mDia2 RNAi. These results demonstrate that mDia2 is essential in mammalian cell cytokinesis and that mDia2-induced F-actin forms a scaffold for the contractile ring and maintains its position in the middle of a dividing cell.
Effects of depletion of mDia isoforms in NIH 3T3 cells. (A) Depletion of mDia proteins (mDia1, mDia2 and mDia3) by siRNA treatment. NIH 3T3 cells were treated with siRNA specific for mDia1, 2, and 3 (A6, siRNAmDia2#1 and siRNAmDia3#2, respectively) or scrambled control siRNA. After 24 and 48 h, the cells were harvested and subjected to immunoblot for each protein. (B) Fluorescence staining for DNA (blue), F-actin (red), and tubulin (green) in NIH 3T3 cells treated with siRNAmDia2#1 for 48 h. All images show stacks of 10 different focal planes. Bar, 50 ␮ m. Arrows represent examples of the binucleate cells. (C) Frequency of binucleation in NIH 3T3 cells subjected to RNAi for mDia isoforms. NIH 3T3 cells were subjected to RNAi as indicated and the number of binucleate cells in each population was determined. The results are from three independent experiments, in each of which N Ͼ 200 cells were examined. Values are shown as percentage of binucleate cells. Error bars, SD. *p Ͻ 0.01 versus control RNAi, mDia1 RNAi, and mDia3 RNAi cells. (D) Rescue of the binucleate phenotype by expression of RNAi-resistant full-length mDia2 construct. NIH 3T3 cells were transfected with pEGFP or pEGFP-mDia2 r#1, and after 24 h, the cells were subjected to control or mDia2 RNAi (siRNAmDia2#1) for 48 h. The cells were either subjected to Western blotting with anti-mDia2 (top), GFP (middle), or tubulin (bottom, left) or subjected to fluorescence staining as described above for determining the number of binucleate cells in each population. Values are shown as percentage of binucleate cells expressing GFP per total cells expressing GFP (right). Error bars, SD. *p Ͻ 0.01. (E) Effects of microinjection of N1 antibody to mDia2. Control IgG or N1 antibody to mDia2 were microinjected into randomly growing NIH 3T3 cells along with Alexa Fluor 594 Dextran as a marker, and after 10 h the cells were fixed and stained for tubulin (green) and DNA (blue). Left, immunofluorescence images of control IgG- (top) or anti-mDia2 antibody- (bottom) injected NIH 3T3 cells. Bar, 50 ␮ m. Right, values are shown as percentage of binucleate cells per injected cells. Error bars, SD. *p Ͻ 0.01. The results are from three independent experiments, in each of which N Ͼ 100 cells were examined.
… 
Content may be subject to copyright.
Molecular Biology of the Cell
Vol. 19, 2328–2338, May 2008
mDia2 Induces the Actin Scaffold for the Contractile Ring
and Stabilizes Its Position during Cytokinesis in NIH 3T3
Cells
Sadanori Watanabe,* Yoshikazu Ando,* Shingo Yasuda,* Hiroshi Hosoya,
Naoki Watanabe,* Toshimasa Ishizaki,* and Shuh Narumiya*
*Department of Pharmacology, Kyoto University Faculty of Medicine, Kyoto 606-8501, Japan; and
Department of Biological Science, Graduate School of Science, Hiroshima University, Higashi-Hiroshima
739-8526, Japan
Submitted October 29, 2007; Revised January 25, 2008; Accepted February 7, 2008
Monitoring Editor: Fred Chang
mDia proteins are mammalian homologues of Drosophila diaphanous and belong to the formin family proteins that
catalyze actin nucleation and polymerization. Although formin family proteins of nonmammalian species such as
Drosophila diaphanous are essential in cytokinesis, whether and how mDia proteins function in cytokinesis remain
unknown. Here we depleted each of the three mDia isoforms in NIH 3T3 cells by RNA interference and examined this
issue. Depletion of mDia2 selectively increased the number of binucleate cells, which was corrected by coexpression of
RNAi-resistant full-length mDia2. mDia2 accumulates in the cleavage furrow during anaphase to telophase, and concen-
trates in the midbody at the end of cytokinesis. Depletion of mDia2 induced contraction at aberrant sites of dividing cells,
where contractile ring components such as RhoA, myosin, anillin, and phosphorylated ERM accumulated. Treatment with
blebbistatin suppressed abnormal contraction, corrected localization of the above components, and revealed that the
amount of F-actin at the equatorial region during anaphase/telophase was significantly decreased with mDia2 RNAi.
These results demonstrate that mDia2 is essential in mammalian cell cytokinesis and that mDia2-induced F-actin forms
a scaffold for the contractile ring and maintains its position in the middle of a dividing cell.
INTRODUCTION
Cytokinesis is the final step in cell division that physically
separates a dividing cell into two. In a somatic cell, separa-
tion, i.e., cleavage occurs in the middle of a dividing cell
between the two spindle poles to ensure each set of segre-
gated chromosomes inherited to each daughter cell. Al-
though cytokinesis is a multistep process under coordinated
control of cell cycle progression, cytoskeletal dynamics, and
vesicle transport, the actomyosin-based constriction by the
contractile ring that is constructed in the equatorial region of
a dividing cell is recognized as a major driving force for
physical separation till abscission (Balasubramanian et al.,
2004). However, how the position of the contractile ring, i.e.,
cleavage plane, is determined and maintained through cy-
tokinesis and how and where actin filaments are produced
and assembled with myosin and other molecules into the
contractile ring in mammalian cells remain largely un-
known.
The small GTPase Rho functions in several organisms and
several lines of cultured mammalian cells as a molecular
switch linking nuclear division and cytokinesis; Rho is acti-
vated in anaphase to telophase and induces the contractile
ring in dividing cells (Mabuchi et al., 1993; Piekny et al., 2005;
Narumiya and Yasuda, 2006). In mammalian cells, the GTP-
bound, activated form of Rho acts on two downstream ef-
fectors to induce actomyosin bundles; one is ROCK/Rho-
kinase that activates myosin for cross-linking of anti-parallel
actin filaments, and the other is mammalian homolog of
Drosophila diaphanous (mDia) protein that induces actin
filaments by catalyzing actin nucleation and polymerization
(Watanabe et al., 1997; Sagot et al., 2002). mDia belongs to the
formin family of proteins and there are three isoforms,
mDia1-3 (Higgs, 2005). mDia has multiple domains, GBD
(GTPase-binding domain) in the N-terminus, FH (formin
homology) domains, FH1 and FH2, in the middle, and DAD
(diaphanous auto-regulatory domain) in the C-terminus
(Higgs, 2005; Rose et al., 2005). The FH2 domain binds to the
barbed end of an actin filament and catalyzes actin nucle-
ation and polymerization. The FH1 domain accelerates actin
elongation by the FH2 domain through binding to the actin
monomer-binding protein, profilin (Watanabe et al., 1997;
Romero et al., 2004; Kovar et al., 2006). By this action, mDia
induces long unbranched actin filaments in contrast to
Arp2/3 complex that induces actin meshwork (Goode and
Eck, 2007). In addition to the action on actin, mDia has been
reported to stabilize and orient microtubules in interphase
and mitotic cells (Ishizaki et al., 2001; Palazzo et al., 2001;
Yasuda et al., 2004). Intriguingly, although involvement of
ROCK in cytokinesis has been examined previously (Kosako
et al., 2000), whether mDia protein is involved in cytokinesis
of mammalian cells and if so, which mDia isoform functions
in this process have not yet been examined thoroughly,
though its nonmammalian orthologues such as Cdc12p in
Schizosaccharomyces pombe and Diaphanous in Drosophila
This article was published online ahead of print in MBC in Press
(http://www.molbiolcell.org/cgi/doi/10.1091/mbc.E07–10–1086)
on February 20, 2008.
Address correspondence to: Shuh Narumiya (snaru@mfour.med.kyoto-u.
ac.jp).
2328 © 2008 by The American Society for Cell Biology
melanogaster have been shown essential for cytokinesis in
each species (Castrillon and Wasserman, 1994; Tominaga et
al., 2000; Pelham and Chang, 2002; Dean et al., 2005). To
examine this issue, we have used RNA interference (RNAi)
to deplete each mDia isoform in NIH 3T3 cells and identified
that one of the mDia isoforms, mDia2, is essential for cyto-
kinesis of this cell line. We have also performed fluorescence
microscopy for contractile ring components such as RhoA,
F-actin, myosin, anillin, and phosphorylated ERM (pERM),
as well as live cell-imaging for myosin and mDia2, and
examined localization and actions of mDia2 in cytokinesis.
We now show that mDia2 localizes in the equatorial region
of a dividing cell in anaphase and induces F-actin there to
provide an actin scaffold for assembly of the contractile ring
and stabilize its position during cytokinesis.
MATERIALS AND METHODS
Materials
Short interfering double-stranded RNA oligomers (siRNAs) K2 and A6 (Ara-
kawa et al., 2003; Yamana et al., 2006) were used for RNAi for mDia1. Three
different siRNAs, siRNAmDia2#1, #2 and #3, corresponding to nucleotide
sequences of 289-313, 1889-1907, and 2150-2168, respectively, were used for
RNAi for mDia2 (NM_019670). Two different siRNAs, siRNAmDia3#1 and #2,
corresponding to nucleotide sequences of 967-991 and 1295-1319, respectively,
were used for RNAi for mDia3 (AY312280). Stealth RNAi negative control
duplexes (Invitrogen, Carlsbad, CA) were used for control RNAi. Block-iT
Alexa Fluor Red fluorescent Oligo (Invitrogen) was used for determination of
efficiency of siRNA transfection. pEGFP-MRLC was described previously
(Miyauchi et al., 2006). pDsRed2-Histone H2BK has been described previously
(Yasuda et al., 2004). pEGFP-mDia2 was prepared as follows. cDNA for
mDia2 (Yasuda et al., 2004) was subcloned into pCR-Blunt vector (Invitrogen;
pCR-Blunt-mDia2). The plasmid was digested with XhoI and KpnI and then
subcloned into pEGFP-C1 (Clontech, San Jose, CA) to generate pEGFP-C1-
mDia2 (XhoI-3), lacking a fragment 5of the XhoI site of the full-length
mDia2. The 5-XhoI fragment of mDia2 was obtained by PCR amplification
using the pCR-Blunt-mDia2 as a template, with 5-GAACTCGAGCTATG-
GAGAGGCACCGGGC-3as the forward primer and 5-GTCCCTCTGCTC-
GAGTTTCC-3as the reverse primer. The resultant PCR fragment was di-
gested with XhoI and then subcloned into pEGFP-C1-mDia2 (XhoI-3)to
generate pEGFP-C1-mDia2. To prepare mDia2-RNAi–resistant pEGFP-
mDia2, pEGFP-mDia2 r#1 and r#2, we introduced silent mutations into the
region of pEGFP-C1-mDia2 targeted by mDia2siRNA #1 using the
QuikChange Site-Directed Mutagenesis Kit II (Stratagene, La Jolla, CA) with
5-GAAAACAACCCAAAGGCGCTGCCCGAAAGCGAGGTGTTGAAGCTTTT-
TGAGAAGATG-3as a template for pEGFP-mDia2 r#1 and 5-CCCAAAG-
GCGCTGCCAGAGTCCGAAGTCTTGAAGCTTTTTGAG-3for pEGFP-mDia2
r#2. To prepare pEGFP-mDia2 I704A, we introduced mutations into the codon
of Ile 704 with 5-GCTCAGAACCTTTCAGCCTTCCTGAGCTCCTTCCG-3
as a template as described above.
Primary antibodies (Abs) used were mouse DM1A monoclonal Ab (mAb)
to
-tubulin, fluorescein isothiocyanate (FITC)-conjugated DM1A, FITC-con-
jugated Ab to
-actin, and rabbit polyclonal Ab to myosin IIA (Sigma-Aldrich,
St. Louis, MO); rat mAb to
-tubulin (Chemicon, Temecula, CA); goat Ab to
mDia3 (N15), rabbit polyclonal Ab to green fluorescent protein (GFP; FL), and
mouse 26C4 mAb to RhoA (Santa Cruz Biotechnology, Santa Cruz, CA); and
rabbit polyclonal Ab to GFP from MBL (Nagoya, Japan). Rabbit polyclonal
antibody to mDia1 was described previously (Watanabe et al., 1997). Poly-
clonal C1 Ab to mDia2 was generated in rabbits against a glutathione S-
transferase (GST) fusion protein of the C-terminal fragment of mDia2 (amino
acid residues, 1056-1171). The fragment was subcloned into pGEX-6P-1 (GE
Healthcare Life Science, Piscataway, NJ) to generate pGEX-mDia2-C, which
was then used for transformation of Escherichia coli BL21 (Novagen, Madison,
WI). After induction with 1 mM IPTG, the bacteria were lysed and the fusion
protein was purified with a GSH-Sepharose column (GE Healthcare Life
Science). The purified protein was injected into rabbits as antigen. The anti-
body to mDia2 was purified from antiserum by affinity chromatography
using the antigen coupled with NHS-activated Sepharose (GE Healthcare Life
Science). Polyclonal N1 Ab to mDia2 was also generated against a GST fusion
protein of the N-terminal fragment of mDia2 (amino acid residues, 33-411) as
described above and purified using the antigen coupled with CNBr-activated
Sepharose (GE Healthcare Life Science). Rabbit anti-anillin and rat anti-
phospho-ERM (pERM) Abs were kind gifts from Dr. Makoto Kinoshita
(Kyoto University) and Professor Sachiko Tsukita (Osaka University, Japan).
Cell Culture and Transfection
NIH 3T3 cells and C2C12 cells were maintained in DMEM (GIBCO, Rockville,
MD) supplemented with 10% fetal calf serum (FCS) at 37°C with an atmo-
sphere containing 10% CO
2
. Transfection of plasmids was performed using
Lipofectamine LTX Reagent (Invitrogen) according to the manufacturer’s
protocol. We diluted 1
g of each plasmid DNA and 2
l of PLUS reagent
(Invitrogen) in 400
l of Opti-MEM, subsequently mixed with 5
l of Lipo-
fectamine LTX. The lipofectamine solution was mixed with 2 ml of fresh
medium and added to cells of 50–60% confluency in one well of a six-well
plate. RNAi was performed using Lipofectamine RNAiMAX Reagent (In-
vitrogen) according to the manufacturer’s reverse transfection protocol. We
mixed 1.2
lof20
M siRNA duplex and 4
l of Lipofectamine RNAiMAX in
400
l of Opti-MEM. NIH 3T3 cells or C2C12 cells of semiconfluency were
washed and suspended with trypsin-EDTA. The siRNA mixture was added
to 1.0 10
5
cells in 2 ml of the culture medium, and the cell suspension was
then seeded in a well of a six-well plate. siRNA experiments in synchronized
cells were performed as follows. NIH 3T3 cells were seeded and cultured for
16 h in a 100-mm dish with the culture medium containing 2 mM thymidine.
The cells were then washed twice with phosphate buffered saline (PBS) and
subjected to RNAi transfection as described above. The cells were then seeded
and further cultured for 8 h. The medium was then replaced again with the
culture medium containing 2 mM thymidine and the cells were cultured for
another 16 h. After washing three times with PBS and once with fresh
medium, the cells were incubated in the culture medium alone or in that
containing 40 ng/ml nocodazole for 8 h. The cells in the latter procedure were
then washed free of nocodazole as described above for thymidine removal.
The cells were cultured in fresh medium for 2 h for time-lapse imaging for the
statistical analysis of the phenotype induced by mDia2 RNAi or for 20 min
either with or without 80
M blebbistatin (Tocris, Ballwin, MO) before being
fixed for immunofluorescence.
Microinjection
NIH 3T3 cells were seeded and cultured for 6 h. The cells were microinjected
with normal rabbit IgG (Santa Cruz) or affinity-purified N1 antibody to
mDia2 (0.01 mg/ml) in PBS with 0.5
g/ml dextran-Alexa-fluor 594 (Molec-
ular Probes, Eugene, OR) using a microinjection system (Eppendorf, Fremont,
CA) with maintenance pressure of 400 hPa and injection pressure of 20 hPa
for 0.1 s. The cells were then incubated for 10 h before fixation.
Fluorescence Microscopy
NIH 3T3 cells or C2C12 cells were plated onto a coverslip in a 35-mm culture
dish for fluorescence microscopy. We used three different fixation protocols.
For phalloidin staining for F-actin, cells were fixed with 4% paraformalde-
hyde in PBS at 37°C for 15 min. The cells were washed three times with PBS
and permeabilized with 0.1% Triton X-100 in PBS for 5 min on ice, followed
by three washes with PBS. For immunofluorescence for GFP, RhoA, anillin,
myosin IIA, pERM, and mDia2, cells were fixed with 10% TCA on ice for 15
min (Yonemura et al., 2004). The fixed cells were washed three times with PBS
containing 30 mM glycine (G-PBS) and permeabilized with 0.2% Triton X-100
in G-PBS for 5 min on ice, followed by three washes with G-PBS. For staining
for
-tubulin and for comparison of mDia1, mDia2, or mDia3 staining, the
cells were fixed with methanol at 20°C for 5 min. After fixation and perme-
abilization as described above, the cells were incubated with 3% BSA in PBS
for 1 h and were incubated at room temperature for2horat4°Covernight
with following primary antibodies: rat anti-tubulin (1:1000 dilution), anti-
pERM (1:1), anti-myosin (1:50), anti-anillin (1:200), anti-mDia1 (1:200), anti-
mDia2 (1:200), and anti-mDia3 (1:50) Abs. After three washes with Tris-
buffered saline (TBS) containing 0.1% Tween-20 (TBST), the cells were
incubated with appropriate secondary antibodies coupled to Alexa Fluor 594
(Molecular Probes), and/or either Texas-Red phalloidin (1:200) or FITC-
conjugated anti-
-tubulin (1:200). The samples were washed three times with
TBST before mounting in Prolong Antifade DAPI-Gold (Molecular Probes) on
glass slides. Staining was examined with a Leica SP5 confocal imaging system
(Plan-Apo 63/1.40 NA; Deerfield, IL). Binucleate or multinucleate cells were
identified on samples stained for tubulin and DNA. The percentage of binu-
cleate or multinucleate cells was determined in a blinded manner by an
observer without information of the identity of the samples. For the images of
interphase cells in Figure 1 and Figure S1, build-up images were obtained
from a collection of 10 Z sections of 0.5-
m step intervals from the bottom to
the top of the cells. For the images of dividing cells in Figure S3D and S4C and
Figures 4–6, build-up images were obtained from a collection of 15–25 Z
sections of 0.5-
m step intervals around the middle section of the cells. The
images were analyzed by the built-in software. Quantification of F-actin
intensity in Figure 5A was performed using MetaMorph software (Universal
Imaging, West Chester, PA) as follows. Images of 20 mitotic cells each in
control and mDia2 RNAi groups were obtained from two different experi-
ments, and the F-actin intensity in each cell was calculated by dividing the
sum of Texas-Red phalloidin fluorescence intensity in each pixel by the total
pixel counts in the cell area.
Immunoblotting
At indicated times of siRNA treatment, NIH 3T3 cells or C2C12 cells were
washed twice with PBS and lysed in 100
l of Laemmli sample buffer.
Immunoblotting was performed as described previously (Ando et al., 2007).
Primary antibodies were diluted in the blocking buffer and added: 1:1000
mDia2 in Cytokinesis
Vol. 19, May 2008 2329
Figure 1. Effects of depletion of mDia isoforms in NIH 3T3 cells. (A) Depletion of mDia proteins (mDia1, mDia2 and mDia3) by siRNA
treatment. NIH 3T3 cells were treated with siRNA specific for mDia1, 2, and 3 (A6, siRNAmDia2#1 and siRNAmDia3#2, respectively) or
scrambled control siRNA. After 24 and 48 h, the cells were harvested and subjected to immunoblot for each protein. (B) Fluorescence staining
for DNA (blue), F-actin (red), and tubulin (green) in NIH 3T3 cells treated with siRNAmDia2#1 for 48 h. All images show stacks of 10 different
focal planes. Bar, 50
m. Arrows represent examples of the binucleate cells. (C) Frequency of binucleation in NIH 3T3 cells subjected to RNAi
for mDia isoforms. NIH 3T3 cells were subjected to RNAi as indicated and the number of binucleate cells in each population was determined.
The results are from three independent experiments, in each of which N 200 cells were examined. Values are shown as percentage of
binucleate cells. Error bars, SD. *p 0.01 versus control RNAi, mDia1 RNAi, and mDia3 RNAi cells. (D) Rescue of the binucleate phenotype
by expression of RNAi-resistant full-length mDia2 construct. NIH 3T3 cells were transfected with pEGFP or pEGFP-mDia2 r#1, and after 24 h,
the cells were subjected to control or mDia2 RNAi (siRNAmDia2#1) for 48 h. The cells were either subjected to Western blotting with
anti-mDia2 (top), GFP (middle), or tubulin (bottom, left) or subjected to fluorescence staining as described above for determining the number
of binucleate cells in each population. Values are shown as percentage of binucleate cells expressing GFP per total cells expressing GFP (right).
Error bars, SD. *p 0.01. (E) Effects of microinjection of N1 antibody to mDia2. Control IgG or N1 antibody to mDia2 were microinjected
into randomly growing NIH 3T3 cells along with Alexa Fluor 594 Dextran as a marker, and after 10 h the cells were fixed and stained for
tubulin (green) and DNA (blue). Left, immunofluorescence images of control IgG- (top) or anti-mDia2 antibody- (bottom) injected NIH 3T3
cells. Bar, 50
m. Right, values are shown as percentage of binucleate cells per injected cells. Error bars, SD. *p 0.01. The results are from
three independent experiments, in each of which N 100 cells were examined.
S. Watanabe et al.
Molecular Biology of the Cell2330
dilution for the Abs to
-tubulin, mDia1, mDia2, or GFP and 1:200 dilution for
the Ab to mDia3. After overnight incubation at 4°C with these Abs except for
incubation at room temperature for 2 h with the Abs to tubulin and GFP, the
membranes were washed three times with TBS containing 0.05% Tween-20.
The bound primary Abs were detected with corresponding horseradish per-
oxidase–conjugated secondary Abs (GE Healthcare Bio-Science, Piscataway,
NJ; 1:3000 dilution in the blocking buffer) and ECL Western Blotting Detection
System (GE Healthcare Bio-Sciences).
Time-Lapse Live Cell Imaging
NIH 3T3 cells were seeded on 35-mm glass-bottom dishes (MatTek, Ashland,
MA) with or without siRNA transfection. The medium was replaced 18–24 h
after siRNA treatment with DMEM containing 10% FCS and 300 nM Syto11
(Invitrogen), and the cells were incubated for 30 min at 37°C. The dish was
then placed on a temperature-controlled stage maintained at 37°C with 5%
CO
2
. Live cell imaging was performed on an inverted microscope (model
DMIRE2; Leica) as described previously (Oceguera-Yanez et al., 2005). Se-
quential time-lapse images were acquired every 3 min for 120–225 min. For
imaging of cells expressing either EGFP-mDia2 or MRLC-EGFP together with
DsRed-histone-H2B, the cells were transfected at 8 h after seeding with
indicated plasmid DNAs. Live cell imaging was performed on the confocal
microscope (TCS-SP5; Leica) with 63/1.40 NA lenses. Sequential time-lapse
images were acquired every 30 s or 1 min for 30 min.
Statistical Analysis
Data are presented as mean SD and were analyzed by Student’s ttest. p
0.01 was considered statistically significant.
RESULTS
Effects of Depletion of Each mDia Isoform on Cytokinesis
To examine whether mDia isoforms are required for cytoki-
nesis, and, if so, to identify which mDia isoform functions in
this process, we transfected NIH 3T3 cells with siRNA spe-
cific to each mDia isoform, A6 for mDia1, siRNAmDia2#1
for mDia2, and siRNAmDia3#2 for mDia3. In these experi-
ments, transfection efficiency of siRNA determined by
Block-iT Red fluorescent Oligo was almost 100% (Figure
S1A). The cells were collected 24 and 48 h after transfection
and subjected to Western blotting. The treatment with each
siRNA specifically and time-dependently suppressed the
expression of corresponding mDia isoforms. After 48 h, only
a negligible amount of each protein was detected by West-
ern blotting; the percentage of depletion of endogenous
proteins at 48 h were 75, 95, and 70% for mDia1, mDia2, and
mDia3, respectively (Figure 1A). The siRNA treatment for
each mDia isoform did not apparently change the cell
shapes. Although RNAi for mDia1 and mDia3 did not affect
cytokinesis (Figure S1, C and D), mDia2-RNAi increased the
proportion of the cells with two nuclei (binucleate cells) as
evidenced by immunofluorescence as well as flow cytom-
etry (Figure 1B and Figure S1D). Failure of cytokinesis,
consequently, increased the size of these cells (Figure S1B).
Quantitative analysis revealed that the proportion of the
binucleate cells significantly increased in the mDia2-de-
pleted cell population compared with that in the control,
mDia1-depleted, or mDia3-depleted population; the per-
centage of binucleate cells increased to 25.8 4.9 and 39.8
1.2% 24 and 48 h after transfection with mDia2 siRNA,
respectively (Figure 1C). We confirmed these results by us-
ing additional nonoverlapping siRNAs for mDia1 and
mDia3, K2 and siRNAmDia3#1, respectively, and two addi-
tional nonoverlapping siRNAs for mDia2, siRNAmDia2#2
and #3. Although all siRNAs sufficiently depleted respective
mDia isoforms, only treatment with siRNAs specific for
mDia2 significantly increased the rates of binucleate cells
(data not shown). To further validate that binucleation is
caused by mDia2 depletion, we attempted to rescue the
phenotype by cotransfecting a RNAi-resistant full-length
construct of mDia2 fused to enhanced GFP (EGFP) (EGFP-
mDia2 r#1) with siRNAmDia2#1. Expression of EGFP was
used as a control in these experiments. Depletion of endog-
enous mDia2 and expression of EGFP-mDia2 r#1 were ver-
ified by immunoblotting (Figure 1D, left). Expression of the
EGFP-mDia2 r#1 fully rescued the effects of mDia2 RNAi on
cytokinesis; the rate of binucleate cells returned to the level
in the control cells after the expression of this construct
(Figure 1D, right). These results suggest that the generation
of binucleate cells was specifically induced by reduction of
mDia2 protein. We further microinjected the antibody N1 to
mDia2 and evaluated its effect on cytokinesis. Ten hours
after we microinjected the mDia2 antibody into randomly
growing NIH 3T3 cells, we observed significant increase in
the percentage of binucleate cells compared with control IgG
injected cells (Figure 1E). We obtained similar results when
we injected the mDia2 antibody into NIH 3T3 cells synchro-
nized in G2 by double thymidine block and observed 5 h
later (data not shown). These results that depletion of mDia2
induces cytokinesis failure demonstrate that an mDia iso-
form is required for cytokinesis of NIH 3T3 cells and that it
is mDia2 that functions as an mDia isoform essential in this
process. In contrast to these findings, Tominaga et al. (2000)
previously microinjected antibodies to mDia1 or mDia2 into
dividing NIH 3T3 cells and found that the injection of anti-
mDia1 antibody and not the antibody to mDia2 interfered
with cytokinesis. Curiously, they did not find significant
expression of mDia2 in these cells, and yet they rescued
cytokinesis failure induced by anti-mDia1 antibody by over-
expression of full-length mDia2. Although the reason for
such apparent discrepancy is currently unknown, we exam-
ined a possibility that mDia1 functions redundantly with
mDia2 in cytokinesis. We therefore depleted mDia isoforms
in every possible combination and examined cytokinesis
failure. We found significant increases in the number of
multi(bi)-nucleate cells upon depletion of mDia1 or mDia3
only combined with mDia2 depletion, i.e., combined deple-
tion of mDia1 and mDia2, that of mDia2 and mDia3 or that
of mDia1, mDia2 and mDia3, and not that of mDia1 and mDia3
(Figure S1E). Importantly, combination of mDia1 and mDia2
depletion did not enhance the rate of binucleate cells compared
with that found on mDia2 depletion alone. These results sug-
gest that at least in the NIH 3T3 cells we used, mDia2 is
primarily important in cytokinesis among mDia isoforms.
Localization of mDia2 during Cell Division
To obtain insights into action mechanism of mDia2 in cyto-
kinesis, we next investigated the localization of endogenous
mDia2 in different phases of cell division by staining for
mDia2, tubulin, and DNA in dividing NIH 3T3 cells using
the C1 antibody to mDia2. This analysis revealed that
signals of mDia2 localized around the basal part of cell
cortex of rounding cells in prometaphase to metaphase,
appeared at equatorial cell cortex in late anaphase, accu-
mulated in the cleavage furrow in telophase, and finally
concentrated in the intercellular bridge at the end of
cytokinesis (Figure 2). The specificity of these signals was
verified by their loss with mDia2 RNAi both on Western blot
analysis and immunofluorescence (Figure 1A and Figure
S2A). We observed similar localization of mDia2 by using a
different mDia2 antibody N1 (data not shown). The localiza-
tion of mDia2 described above was also confirmed by ex-
pressing pEGFP-mDia2 and monitoring the GFP signal. The
GFP signals began to localize at the equatorial surface dur-
ing late anaphase and concentrated in the cleavage furrow
during cytokinesis (Movie S1). For comparison, we exam-
ined the localization of mDia1 and mDia3 during cell divi-
sion. Weak diffuse signals for mDia1 were observed over the
cell bodies and intercellular bridge at the end of telophase
mDia2 in Cytokinesis
Vol. 19, May 2008 2331
(Figure S2B); signals for mDia3 were found in association
with the central spindle in anaphase, and the associated
mDia3 signals were concentrated in the midbody as central
spindle microtubules were bundled in cytokinesis (Figure
S2C). Staining for neither mDia1 nor mDia3 yielded signals
in the cleavage furrow as that for mDia2. These localization
patterns of mDia1 and mDia3 are consistent with our pre-
vious results (Kato et al., 2001; Yasuda et al., 2004). These
results suggest that mDia2 accumulates specifically in the
cleavage furrow during cytokinesis.
We next wondered whether this function of mDia2 in
cytokinesis is conserved in other cell lines. We observed that
mDia2 RNAi also induced cytokinesis failure in C2C12
mouse myoblast cells and mDia2 localized to the cleavage
furrow of these cells (Figure S3), supporting the conserved
localization and function of mDia2 during cytokinesis.
Abnormal Contraction in mDia2-depleted Cells
To investigate how mDia2 depletion induces cytokinesis
failure in NIH 3T3 cells, we monitored the progression of
cell division of mDia2-RNAi cells by videomicroscopy. We
first used randomly growing cells subjected to control and
mDia2 RNAi. In control cells, the cleavage furrow appeared
6 min after anaphase onset and ingressed progressively
thereafter to separate two daughter cells within several min-
utes, and the separated daughter cells linked by the inter-
cellular bridge began to spread at 18 min (Figure 3A and
Movie S2). In contrast, mDia2 RNAi cells exhibited abnor-
mal cytokinesis behavior. In one group of the cells, the
cleavage furrow appeared between two daughter cells in
anaphase and began to ingress. However, the ingression was
not properly maintained but followed by robust contraction
at aberrant sites of daughter cells that prevented the ingres-
sion at the cleavage furrow and pushed separating chromo-
somes one end to the other, resulting in fusion of the daugh-
ter cells and production of a binucleate cell (Figure 3B and
Figure 2. Localization of mDia2 during cell division. NIH 3T3 cells
in different phases of cell division were stained for mDia2 (red),
tubulin (green), and DNA (blue). Each image shows a single focal
plane; the image of a prometaphase cell represents that at the basal
plane, and those of cells in anaphase, telophase and cytokinesis
represent those at the middle plane. Bar, 5
m.
Figure 3. Cytokinesis failure in mDia2-RNAi cells.
(A–C) Selected frames from time-lapse movies of
NIH 3T3 cells subjected to control RNAi (A and
Movie S2) or RNAi for mDia2 (B and C and Movies
S3 and S4). Phase contrast and green fluorescence
images were monitored by videomicroscopy after
DNA (green) was stained by Syto11. 0 min, anaphase
onset. Insets show magnified views of the DNA in
each merged image. Arrows in B and C indicate
abnormal contraction in the mDia2-depleted cells.
Bar, 10
m.
S. Watanabe et al.
Molecular Biology of the Cell2332
Movie S3). In another group of mDia2 RNAi cells, contrac-
tion was frequently observed already at metaphase, and
such contraction at abnormal sites continued in anaphase to
telophase and apparently inhibited appearance and func-
tioning of the cleavage furrow at the prospective site, lead-
ing to formation of a binucleate cell upon spreading (Figure
3C and Movie S4).
To statistically analyze the cytokinesis failure in mDia2-
RNAi cells, we next enriched mitotic cells by using thymi-
dine and nocodazole after RNAi treatment and tracked the
cell division by videomicroscopy. We confirmed a similar
phenotype of the cytokinesis failure in mDia2-RNAi cells
(data not shown). Percentages of cells showed cytokinesis
failure were 6.1 and 49.0% for control (n 33) and mDia2-
RNAi cells (n 51), respectively. Percentages of cells
showed the abnormal contraction were 3.0 and 76.5% for
control (n 33) and mDia2-RNAi cells (n 51), respec-
tively, suggesting that there is a link between the cytokinesis
failure and the abnormal contraction induced by mDia2
RNAi. These results indicate that mDia2 is required for
maintenance and functioning of the contractile ring between
two daughter cells.
Effects of mDia2 Depletion on Distribution of the
Contractile Ring Components
The contractile ring is composed of several components
including F-actin, myosin, anillin, and pERM (Straight et al.,
2003; Yokoyama et al., 2005). To examine whether these
components accumulate and are maintained at the prospec-
tive cleavage site during cytokinesis of mDia2 RNAi cells,
we stained for these contractile ring components by using
Texas-Red phalloidin and antibodies to each molecule in
cells depleted of mDia2. We stained for myosin by using
antibody to myosin heavy chain IIA. mDia2-RNAi cells
showed aberrant shapes (Figure 4), which probably reflected
abnormal contraction observed in time-lapse imaging anal-
ysis (Movies S3 and S4). In these mDia2-RNAi cells, signals
for each of F-actin, myosin, anillin, and pERM were not
observed at the cleavage furrow as in control cells, but
aberrantly localized around the cell cortex where abnormal
cell shape change was observed (Figure 4A). To examine
dynamics of such localization of the contractile ring compo-
nents, we expressed EGFP-fusion of myosin regulatory light
chain (MRLC-EGFP; Miyauchi et al., 2006) and followed its
movement during cell division. Although myosin accumu-
lated normally at the cleavage furrow from anaphase to
telophase and concentrated there as the furrow ingressed in
cytokinesis in control-RNAi cells (Figure 4B and Movies S5
and S6), significant accumulation of MRLC-EGFP was found
at sites of abnormal contraction in the cell cortex of mDia2-
RNAi cells. Intriguingly, such abnormal contraction was
observed already in prometaphase/metaphase before chro-
mosomes began to segregate and became more robust in
anaphase to telophase, and MRLC-EGFP was found to ac-
cumulate at sites of each contraction (Figure 4C and Movies
S7 and S8). This contraction-associated localization pattern
of EGFP-MRLC appears similar to localization of other com-
ponents of the contractile ring in fixed preparation of mDia2
RNAi cells, indicating that the components of the contractile
ring in mDia2-depleted cells were not maintained at the
cleavage furrow but moved together to the site of abnormal
contraction. We next examined effects of combined deple-
tion of mDia2 with other mDia isoforms in order to examine
whether this abnormal contraction is a direct consequence of
the loss of mDia2 or due to a shift in the balance of Rho
effectors caused by the mDia2 depletion. Combined deple-
tion of either mDia1 or mDia3 or both with mDia2 did not
abolish abnormal contraction caused by the loss of mDia2
(Figure 4D). Immunofluorescence study also showed that
there was no clear accumulation of mDia1 and mDia3 at the
site of abnormal contraction (data not shown). These results
suggest that mDia2 itself is important for proper positioning
and maintenance of the contractile ring components during
cell division.
Effects of Blebbistatin on Accumulation of the Contractile
Ring Components in mDia2-depleted Cells
The above results demonstrate that the contractile ring com-
ponents accumulate at various sites of mDia2-RNAi cells
where aberrant contraction occurs. On the other hand, it has
Figure 4. Abnormal localization of contrac-
tile ring components during anaphase/telo-
phase in mDia2-depleted cells. (A) Immuno-
fluorescence. NIH 3T3 cells were transfected
with siRNA for mDia2 (siRNAmDia2#1) and
synchronized with thymidine and nocoda-
zole, and at 32 h after transfection the cells
were released into DMEM containing 10%
FCS for 20 min. The cells were fixed and
stained for indicated molecules (red), tubu-
lin (green), and DNA (blue). All images
show stacks of focal planes as described in
Materials and Methods. Bar, 5
m. Arrows
indicate abnormal accumulation of indicated
molecules. (B and C) Selected frames from
time-lapse movies of NIH 3T3 cells subjected
to control (B and Movies S5 and S6) or
mDia2 RNAi (C and Movies S7 and S8), and
labeled with DsRed-histone H2Bk (red) and
MRLC-EGFP (green). Bar, 5
m. (D) Effects
of combined depletion of mDia2 and other
mDia isoforms. NIH 3T3 cells were trans-
fected with siRNA for mDia isoforms in in-
dicated combination for 48 h, and the cells
were stained for F-actin (red), tubulin
(green), and DNA (blue). A typical example is shown out of 20 cells examined. Arrows indicate abnormal accumulation of F-actin. Bar,
5
m.
mDia2 in Cytokinesis
Vol. 19, May 2008 2333
been shown that various contractile ring components utilize
apparently different mechanisms and accumulate at the
equatorial region of the dividing cell cortex in anaphase and
are assembled there to the contractile ring (Straight et al.,
2003; Yokoyama et al., 2005). We wondered whether mDia2
depletion interfered with such initial accumulation process of
the components. We could not fully examine this issue in intact
mDia2-depleted cells because of strong contraction in these
cells. We therefore used blebbistatin, a myosin II ATPase
inhibitor, to suppress contraction and examined localization
of each component (Straight et al., 2003). Immunofluores-
cence analysis showed that the accumulation of signals for
F-actin at the equatorial region appeared weaker in mDia2-
RNAi cells compared with control-RNAi cells (Figure 5A,
Figure 4. (cont).
S. Watanabe et al.
Molecular Biology of the Cell2334
left). Decrease of F-actin in mDia2-RNAi cells was verified
by quantitative fluorescence intensity measurements of
whole cells subjected to control and mDia2 RNAi (Figure
5A, right). Given that mDia proteins have actin-nucleating
and -polymerizing activity, these results indicate that mDia2
is indeed responsible for formation of F-actin at this site.
Therefore, we next examined whether this actin-nucleating/
polymerizing activity of mDia2 is indeed required for the
cytokinesis by attempting to rescue the mDia2 RNAi-in-
duced cytokinesis failure by expressing an actin-polymer-
ization–defective GFP-mDia2 mutant. Mutation into Ile
704
of
mDia2 has been shown to render mDia2 actin polymeriza-
tion-defective (Xu et al., 2004; Harris et al., 2006). We found
that GFP full-length mDia2 I704A could not rescue the
mDia2 RNAi-induced cytokinesis failure, whereas this GFP
full-length mDia2 I704A localized to the cleavage furrow
(Figure S4, B and C). These data further support that the
actin-nucleating/polymerizing activity of mDia2 is required
for cytokinesis.
In contrast to the above findings on F-actin, we found
signals for myosin, anillin, or pERM at equatorial cortex of
both control and mDia2-RNAi cells in the presence of bleb-
bistatin, but these components showed broader localization
in mDia2-RNAi cells compared with control cells (Figure
5B). The percentages of cells showing the localization of each
components wider than the 120° sector from the center of the
cell are 20 and 62% for myosin for control and mDia2-RNAi
cells (n 50 for each), 16 and 58% for anillin for control and
mDia2-RNAi cells (n 50 for each), 29 and 40% for pERM
for control and mDia2-RNAi cells, respectively (n 45 for
each). We also treated control-RNAi cells with latrunculin B
in the presence of blebbistatin during anaphase/telophase to
test the contribution of F-actin in this effect, and found that
latrunculin-treated cells showed dispersion of signals for
anillin and myosin around the cell cortex (data not shown).
Given that latrunculin-treated cells showed dispersed local-
ization of the contractile ring components including anillin,
the mDia2-driven F-actin production in the equatorial cortex
appeared important for concentration and maintenance of
the contractile ring components, myosin II, anillin, and
pERM in the cleavage furrow during anaphase/telophase.
Effects of mDia2 Depletion on Rho Localization during
Cytokinesis
The small GTPase Rho is activated in anaphase to telophase,
localizes in the cleavage furrow, and induces assembly of the
contractile ring there (Piekny et al., 2005). It is also believed
that Rho acts on downstream effectors such as ROCK and
citron kinase and produces contraction of the contractile ring
for cleavage (Madaule et al., 1998; Kosako et al., 2000; Ueda
et al., 2002; Yamashiro et al., 2003; Piekny et al., 2005). It is
therefore interesting to know whether aberrant contraction
seen in mDia2-depleted cells is associated with or dissoci-
ated from Rho activation. To examine this issue, we stained
for RhoA during anaphase/telophase in mDia2-depleted
cells in the presence or absence of blebbistatin (Figure 6, A
and B). Although signals for RhoA accumulated in the cleav-
age furrow during anaphase/telophase in control cells,
RhoA localized aberrantly at sites of the cell cortex of ap-
parently abnormal contraction in mDia2-RNAi cells without
blebbistatin (Figure 6A). On the other hand, in the presence
of blebbistatin, almost all signals for RhoA were restricted to
the equatorial region of mDia2-RNAi cells. The equatorial
localization of RhoA was found in 18 of 20 cells in mDia2-
Figure 5. Localization of contractile ring
components in blebbistatin-treated cells. NIH
3T3 cells were subjected to control or mDia2
RNAi and synchronized with thymidine and
nocodazole, and at 32 h after transfection the
cells were released into the medium contain-
ing 80
M blebbistatin for 20 min. The cells
were fixed and stained for indicated molecules
(red), tubulin (green), and DNA (blue). All
images show stacks of focal planes as de-
scribed in Materials and Methods. Bar, 5
m. (A)
Phalloidin staining for F-actin in blebbistatin-
treated cells. Left, phalloidin staining of con-
trol or mDia2-RNAi cells are shown. Right,
quantitative analysis of fluorescence intensity
for F-actin in control or mDia2-RNAi cells.
Fluorescence intensity of phalloidin staining
for total area of each cell was quantified using
MetaMorph in 20 cells of each group from two
different experiments as described in Materials
and Methods. *p 0.01. (B) Immunofluores-
cence for MyosinIIA, anillin, and pERM in
blebbistatin-treated cells.
mDia2 in Cytokinesis
Vol. 19, May 2008 2335
RNAi cells with blebbistatin compared with 8 of 20 in
mDia2-RNAi cells without blebbistatin (Figure 6B). Given
that blebbistatin was added upon mitosis in this experi-
ment, these results taken together suggest that RhoA lo-
calizes in the prospective cleavage furrow and induces the
contractile ring complex there in a manner independent of
mDia2, but that mDia2-driven F-actin is important for
maintenance of the RhoA-containing contractile ring com-
plex at the site of the cleavage furrow.
DISCUSSION
In this work, using RNAi, we have examined the involve-
ment of three mDia isoforms in cytokinesis and identified
mDia2 as an essential mDia isoform required for cytokinesis
in NIH 3T3 cells and C2C12 cells (Figure 1 and Figure S3).
We have also analyzed localization of the mDia isoforms
and found that mDia2 localizes specifically in the cleavage
furrow during cytokinesis. We have found that depletion of
mDia2 substantially reduced the amount of F-actin accumu-
lating in the cleavage furrow but did not affect accumulation
of other contractile ring components such as RhoA, myosin,
anillin, and pERM in the equatorial region. However, with-
out mDia2, these components could not be properly inte-
grated into the contractile ring, but apparently formed in-
complete complexes, which were shifted away from the
equatorial region and induced contraction at aberrant site of
a dividing cell.
Our findings summarized above can thus not only suggest
functions of mDia2 in cytokinesis but also address to some
of the major questions regarding cytokinesis. First, given
that mDia molecules can stabilize and align microtubules
(Ishizaki et al., 2001; Palazzo et al., 2001; Yasuda et al., 2004)
and that the cleavage plane is suggested to be specified by
spindle microtubules that are stabilized in the equatorial
region (Canman et al., 2003), it is possible that mDia isoforms
are involved in determination of the cleavage plane. How-
ever, the above observation that depletion of mDia2 does not
interfere with RhoA accumulation in the equator argues
against this idea and rather suggests that RhoA accumulates
first in the cleavage plane and recruit mDia2 there. This is
consistent with the property of mDia2 to bind to members of
the Rho GTPases such as RhoA, Rac1, and Cdc42 (Alberts et
al., 1998; Yasuda et al., 2004). However, binding to RhoA
cannot explain selective localization of mDia2 to the cleav-
age furrow, because other mDia isoforms can bind to RhoA
as well. Further analysis is therefore required to elucidate a
mechanism for selective localization of mDia2 to the cleav-
age furrow.
Second, it is intriguing that mDia2 localizes in the cleav-
age furrow and that its depletion reduced the F-actin
amount there. Previously, it was argued whether F-actin is
formed in the cleavage furrow or formed elsewhere and
transported to the furrow (Wang, 2005; Eggert et al., 2006).
Given that mDia molecules are capable of catalyzing actin
nucleation and polymerization, our results strongly suggest
that the majority of F-actin is produced in situ in the cleav-
age furrow by the action of mDia2 and accumulate there.
Then, what functions do actin filaments induced by mDia2
exert in cytokinesis? Abnormal contraction at aberrant sites
apparently by the contractile ring components including
RhoA, myosin, anillin, and pERM in mDia2-depleted cells
suggests that with mDia2 depletion and/or with depletion
of mDia2-induced F-actin, the contractile ring is not prop-
erly organized and is not maintained at the prospective site
of the cleavage furrow, which indicates that mDia2 and
mDia2-induced F-actin link these components together to
form the contractile ring and stabilize its position in the
equatorial region. Formins such as mDia2 can produce long,
straight actin filaments. The structure of the contractile ring
was studied by electron microscopy in fission yeast and
newt eggs, and these studies revealed that it consists of
anti-parallel bundles of straight actin filaments that are
bound to the plasma membrane through barbed ends (Ma-
buchi et al., 1988; Kamasaki et al., 2007). Although it is
argued whether such structure is applied also to the con-
tractile ring in mammalian cells (Eggert et al., 2006), it is
tempting to speculate that mDia2 induces straight actin
filaments of opposite directionality in the prospective site of
the cleavage furrow, which provide an actin-based scaffold
encircling the equatorial region of dividing cells and facili-
tate formation of the contractile ring complex by incorporat-
ing other components of the ring such as myosin, anillin,
and pERM to this actin scaffold. By such actions, mDia2 may
restrict the movement of the contractile ring and stabilize its
position. mDia2 may also function in anchoring the actin
filaments of the contractile ring to the plasma membrane,
because it accumulates in the equatorial cell cortex and its
binds to the barbed end of actin filaments. Our results are
thus consistent with and have substantially extended the
findings by Dean et al. (2005), who examined localization of
myosin in dividing Drosophila S2 cells subjected to RNAi for
diaphanous and showed that diaphanous is required for
Figure 6. Effects of mDia2 depletion on RhoA localization during
cytokinesis. (A) Aberrant localization of RhoA in mDia2-RNAi cells.
NIH 3T3 cells were transfected and synchronized as described in the
legend for Figure 4. The cells were released into the medium with-
out blebbistatin. After 20 min, the cells were fixed and stained for
RhoA (red), tubulin (green), and DNA (blue). All images show
stacks of 15–25 focal planes. Bar, 5
m. (B) Effects of blebbistatin on
RhoA localization in mDia2-RNAi cells. NIH 3T3 cells were sub-
jected to mDia2 RNAi and synchronized with double thymidine
block. The cells were treated with DMSO or 80
M blebbistatin for
30 min 6.5 h after thymidine removal and stained for RhoA (red),
tubulin (green), and DNA (blue). All images show stacks of 15–25
focal planes. Bar, 5
m.
S. Watanabe et al.
Molecular Biology of the Cell2336
maintenance of myosin II to the cleavage furrow. It has to be
mentioned, however, that construction of the contractile ring
may not be governed solely by mDia2 but by interdependent
actions of the contractile ring components including mDia2.
Recently, anillin has been reported to bind myosin, and
RNAi of anillin induces a phenotype similar to that we have
found in mDia2-depleted cells (Straight et al., 2005).
Finally, in this study, we also noted that the mitotic cells
depleted of mDia2 exhibited mild oscillatory contractions
during prometaphase and metaphase and that the cortical
localization of myosin was not uniform as typically seen in
control cells (Movies S4, S7, and S8). Mitotic cell rounding is
the process in which a flat interphase cell becomes spherical
and is associated with rearrangement of the actin cytoskel-
eton, de-adhesion, and an increase in cortical rigidity (Mad-
dox and Burridge, 2003; Thery and Bornens, 2006). Maddox
and Burridge (2003) reported that mitotic cell rounding re-
quires activation of RhoA. In this study, we observed that
mDia2 localized to the cell margin in rounding NIH 3T3 cells
and mDia2-depleted cells exhibit impaired mitotic rounding
(Figures 2 and 3B). The oscillatory contractions of mDia2-
depleted cells mentioned above may be caused by impaired
rigidity of mitotic cells in the absence of mDia2. Eisenmann
et al. (2007) showed that expression of Dip, which they
claimed as an inhibitory binding protein for mDia2, induced
nonapoptotic blebbing in HeLa cells, which is thought to be
caused by breaks in cortical rigidity. These results suggest
that, in addition to its action in cytokinesis, mDia2 also
function in maintenance of cortical rigidity and rounding of
mitotic cells. Given that cytokinesis is now recognized as a
consequence of many events occurring globally in the cell
cortex through cell division (Maddox and Burridge, 2003;
Wang, 2005; Mukhina et al., 2007), these results may suggest
that mDia2 functions not only by inducing F-actin in the
cleavage furrow but also by regulating the cortical rigidity
globally. It may shift the balance of the contractility of
mitotic cells by shifting its accumulation in the cell depen-
dent on the phase of cell division. Elucidation of a mecha-
nism how functions of mDia2 in different phases of cell
division is regulated may unravel how cells execute mitosis
and cytokinesis properly through adjusting cell morphogen-
esis with chromosome separation.
ACKNOWLEDGMENTS
We thank K. Nonomura, Y. Kitagawa, and T. Arai for assistance and A. Fujita
(of our department) for N1 antibody to mDia2. We also thank C. Higashida,
T. Miki, F. Oceguera-Yanez, and J. Monypenny for helpful suggestions and
comments. This work was supported by a Grant-in-Aid for Specially Pro-
moted Research from the Ministry for Education, Culture, Sports, Science,
and Technology.
REFERENCES
Alberts, A. S., Bouquin, N., Johnston, L. H., and Treisman, R. (1998). Analysis
of RhoA-binding proteins reveals an interaction domain conserved in hetero-
trimeric G protein beta subunits and the yeast response regulator protein
Skn7. J. Biol. Chem. 273, 8616–8622.
Ando, Y., Yasuda, S., Oceguera-Yanez, F., and Narumiya, S. (2007). Inactiva-
tion of Rho GTPases with Clostridium difficile toxin B impairs centrosomal
activation of Aurora-A in G2/M transition of HeLa cells. Mol. Biol. Cell 18,
3752–3763.
Arakawa, Y., Bito, H., Furuyashiki, T., Tsuji, T., Takemoto-Kimura, S.,
Kimura, K., Nozaki, K., Hashimoto, N., and Narumiya, S. (2003). Control of
axon elongation via an SDF-1
/Rho/mDia pathway in cultured cerebellar
granule neurons. J. Cell Biol. 161, 381–391.
Balasubramanian, M. K., Bi, E., and Glotzer, M. (2004). Comparative analysis
of cytokinesis in budding yeast, fission yeast and animal cells. Curr. Biol. 14,
R806–R818.
Canman, J. C., Cameron, L. A., Maddox, P. S., Straight, A., Tirnauer, J. S.,
Mitchison, T. J., Fang, G., Kapoor, T. M., and Salmon, E. D. (2003). Determin-
ing the position of the cell division plane. Nature 424, 1074–1078.
Castrillon, D. H., and Wasserman, S. A. (1994). Diaphanous is required for
cytokinesis in Drosophila and shares domains of similarity with the products
of the limb deformity gene. Development 120, 3367–3377.
Dean, S. O., Rogers, S. L., Stuurman, N., Vale, R. D., and Spudich, J. A. (2005).
Distinct pathways control recruitment and maintenance of myosin II at the
cleavage furrow during cytokinesis. Proc. Natl. Acad. Sci. USA 102, 13473–
13478.
Eggert, U. S., Mitchison, T. J., and Field, C. M. (2006). Animal cytokinesis:
from parts list to mechanisms. Annu. Rev. Biochem. 75, 543–566.
Eisenmann, K. M., Harris, E. S., Kitchen, S. M., Holman, H. A., Higgs, H. N.,
and Alberts, A. S. (2007). Dia-interacting protein modulates formin-mediated
actin assembly at the cell cortex. Curr. Biol. 17, 579–591.
Goode, B. L., and Eck, M. J. (2007). Mechanism and function of formins in the
control of actin assembly. Annu. Rev. Biochem. 76, 593–627.
Harris, E. S., Rouiller, I., Hanein, D., and Higgs, H. N. (2006). Mechanistic
differences in actin bundling activity of two mammalian formins, FRL1 and
mDia2. J. Biol. Chem. 281, 14383–14392.
Higgs, H. N. (2005). Formin proteins: a domain-based approach. Trends
Biochem. Sci. 30, 342–353.
Ishizaki, T., Morishima, Y., Okamoto, M., Furuyashiki, T., Kato, T., and
Narumiya, S. (2001). Coordination of microtubules and the actin cytoskeleton
by the Rho effector mDia1. Nat. Cell Biol. 3, 8–14.
Kamasaki, T., Osumi, M., and Mabuchi, I. (2007). Three-dimensional arrange-
ment of F-actin in the contractile ring of fission yeast. J. Cell Biol. 178, 765–771.
Kato, T., Watanabe, N., Morishima, Y., Fujita A., Ishizaki T., and Narumiya,
S. (2001). Localization of a mammalian homolog of disphanous, mDia1, to the
mitotic spindle in HeLa cells. J Cell Sci. 114, 775–784.
Kosako, H., Yoshida, T., Matsumura, F., Ishizaki, T., Narumiya, S., and
Inagaki, M. (2000). Rho-kinase/ROCK is involved in cytokinesis through the
phosphorylation of myosin light chain and not ezrin/radixin/moesin pro-
teins at the cleavage furrow. Oncogene 19, 6059–6064.
Kovar, D. R., Harris, E. S., Mahaffy, R., Higgs, H. N., and Pollard, T. D. (2006).
Control of the assembly of ATP- and ADP-actin by formins and profilin. Cell
124, 423–435.
Mabuchi, I., Hamaguchi, Y., Fujimoto, H., Morii, N., Mishima, M., and Na-
rumiya, S. (1993). A rho-like protein is involved in the organisation of the
contractile ring in dividing sand dollar eggs. Zygote 1, 325–331.
Mabuchi, I., Tsukita, S., and Sawai, T. (1988). Cleavage furrow isolated from
newt eggs: contraction, organization of the actin filaments, and protein com-
ponents of the furrow. Proc. Natl. Acad. Sci. USA 85, 5966–5970.
Madaule, P., Eda, M., Watanabe, N., Fujisawa, K., Matsuoka, T., Bito, H.,
Ishizaki, T., and Narumiya, S. (1998). Role of citron kinase as a target of the
small GTPase Rho in cytokinesis. Nature 394, 491–494.
Maddox, A. S., and Burridge, K. (2003). RhoA is required for cortical retrac-
tion and rigidity during mitotic cell rounding. J. Cell Biol. 160, 255–265.
Miyauchi, K., Yamamoto, Y., Kosaka, T., and Hosoya, H. (2006). Myosin II
activity is not essential for recruitment of myosin II to the furrow in dividing
HeLa cells. Biochem. Biophys. Res. Commun. 350, 543–548.
Mukhina, S., Wang, Y. L., and Murata-Hori, M. (2007).
-Actinin is required
for tightly regulated remodeling of the actin cortical network during cytoki-
nesis. Dev. Cell 13, 554–565.
Narumiya, S., and Yasuda, S. (2006). Rho GTPases in animal cell mitosis. Curr.
Opin. Cell Biol. 18, 199–205.
Oceguera-Yanez, F., Kimura, K., Yasuda, S., Higashida, C., Kitamura, T.,
Hiraoka, Y., Haraguchi, T., and Narumiya, S. (2005). Ect2 and MgcRacGAP
regulate the activation and function of Cdc42 in mitosis. J. Cell Biol. 168,
221–232.
Palazzo, A. F., Cook, T. A., Alberts, A. S., and Gundersen, G. G. (2001). mDia
mediates Rho-regulated formation and orientation of stable microtubules.
Nat. Cell Biol. 3, 723–729.
Pelham, R. J., and Chang, F. (2002). Actin dynamics in the contractile ring
during cytokinesis in fission yeast. Nature 419, 82–86.
Piekny, A., Werner, M., and Glotzer, M. (2005). Cytokinesis: welcome to the
Rho zone. Trends Cell Biol. 15, 651–658.
Romero, S., Le Clainche, C., Didry, D., Egile, C., Pantaloni, D., and Carlier,
M. F. (2004). Formin is a processive motor that requires profilin to accelerate
actin assembly and associated ATP hydrolysis. Cell 119, 419–429.
mDia2 in Cytokinesis
Vol. 19, May 2008 2337
Rose, R., Weyand, M., Lammers, M., Ishizaki, T., Ahmadian, M. R., and
Wittinghofer, A. (2005). Structural and mechanistic insights into the interac-
tion between Rho and mammalian Dia. Nature 435, 513–518.
Sagot, I., Klee, S. K., and Pellman, D. (2002). Yeast formins regulate cell
polarity by controlling the assembly of actin cables. Nat. Cell Biol. 4, 42–50.
Straight, A. F., Cheung, A., Limouze, J., Chen, I., Westwood, N. J., Sellers, J. R.,
and Mitchison, T. J. (2003). Dissecting temporal and spatial control of cytoki-
nesis with a myosin II Inhibitor. Science 299, 1743–1747.
Straight, A. F., Field, C. M., and Mitchison, T. J. (2005). Anillin binds non-
muscle myosin II and regulates the contractile ring. Mol. Biol. Cell 16, 193–
201.
Thery, M., and Bornens, M. (2006). Cell shape and cell division. Curr. Opin.
Cell Biol. 18, 648–657.
Tominaga, T., Sahai, E., Chardin, P., McCormick, F., Courtneidge, S. A., and
Alberts, A. S. (2000). Diaphanous-related formins bridge Rho GTPase and Src
tyrosine kinase signaling. Mol. Cell 5, 13–25.
Ueda, K., Murata-Hori, M., Tatsuka, M., and Hosoya, H. (2002). Rho-kinase
contributes to diphosphorylation of myosin II regulatory light chain in non-
muscle cells. Oncogene 21, 5852–5860.
Wang, Y. L. (2005). The mechanism of cortical ingression during early cyto-
kinesis: thinking beyond the contractile ring hypothesis. Trends Cell Biol. 15,
581–588.
Watanabe, N., Madaule, P., Reid, T., Ishizaki, T., Watanabe, G., Kakizuka, A.,
Saito, Y., Nakao, K., Jockusch, B. M., and Narumiya, S. (1997). p140mDia, a
mammalian homolog of Drosophila diaphanous, is a target protein for Rho
small GTPase and is a ligand for profilin. EMBO J. 16, 3044–3056.
Xu, Y., Moseley, J. B., Sagot, I., Poy, F., Pellman, D., Goode, B. L., and Eck, M. J.
(2004). Crystal structures of a Formin Homology-2 domain reveal a tethered
dimer architecture. Cell 116, 711–723.
Yamana, N. et al. (2006). The Rho-mDia1 pathway regulates cell polarity and
focal adhesion turnover in migrating cells through mobilizing Apc and c-Src.
Mol. Cell. Biol. 26, 6844–6858.
Yamashiro, S., Totsukawa, G., Yamakita, Y., Sasaki, Y., Madaule, P., Ishizaki,
T., Narumiya, S., and Matsumura, F. (2003). Citron kinase, a Rho-dependent
kinase, induces di-phosphorylation of regulatory light chain of myosin II.
Mol. Biol. Cell 14, 1745–1756.
Yasuda, S., Oceguera-Yanez, F., Kato, T., Okamoto, M., Yonemura, S., Terada,
Y., Ishizaki, T., and Narumiya, S. (2004). Cdc42 and mDia3 regulate microtu-
bule attachment to kinetochores. Nature 428, 767–771.
Yokoyama, T., Goto, H., Izawa, I., Mizutani, H., and Inagaki, M. (2005).
Aurora-B and Rho-kinase/ROCK, the two cleavage furrow kinases, indepen-
dently regulate the progression of cytokinesis: possible existence of a novel
cleavage furrow kinase phosphorylates ezrin/radixin/moesin (ERM). Genes
Cells 10, 127–137.
Yonemura, S., Hirao-Minakuchi, K., and Nishimura, Y. (2004). Rho localiza-
tion in cells and tissues. Exp. Cell Res. 295, 300–314.
S. Watanabe et al.
Molecular Biology of the Cell2338
... Arp2/3 is strongly implicated in cell motility, as it is present and essential in lamellipodia of migrating cells (Buracco et al., 2019). Formin are required in numerous species for cytokinesis in particular it elongates long Actin filaments that form the bundles which compose the contractile ring Chang et al., 1997;Pelham and Chang, 2002;Severson et al., 2002;Watanabe et al., 2008), while Arp2/3 -despite its role regulating Formin activity (Chan et al., 2019)-does not seem an essential factor for this process, as the cytokinesis still occurs when Arp2/3 is depleted by RNAi (Severson et al., 2002). ...
... Even though active polymerization of Actin filaments is not require for constriction of the contractile ring (Mishra et al., 2013), proper Actomyosin contractility requires Actin turnover during cytokinesis in order to preserve Actin filaments homeostasis (Chew et al., 2017). In mouse, in addition to the activation of RhoA GTPase, the binding of Anillin is necessary for the proper localization and function of Formin mDia2 at the cleavage furrow location (Watanabe et al., 2008;. ...
Thesis
Actin network architecture and dynamics play a central role in cell contractility and tissue morphogenesis. Local modulations of Actomyosin network dynamics depend largely on the activation of the RhoA activation cascade. In my thesis, I combined quantitative microscopy using TIRFM, single-molecule imaging, numerical simulations and simple mathematical modeling, to explore the dynamic network architecture underlying pulsed contractions in a simple model, the C. elegans early embryo. Focusing on the Actin elongator Formin, we observed that F-Actin elongation was catalyzed by a specific subpopulation of cortical Formins – termed elongating Formins – that displayed a characteristic ballistic mobility. My results also showed that Formin-mediated F-Actin elongation rate was dependent on the phase of the cell cycle and embryonic stage. We subsequently showed that elongating Formins saturate available barbed ends of Actin filaments, converting a local biochemical gradient of RhoA activity into a polar network architecture. In second study, focusing on the kinetics of the RhoA activation cascade, we developed and functionally challenged a simple numerical model. This model takes advantage of the measurements of the dynamical parameters of the Myosin, downstream effector of the RhoA activation cascade, to predict the temporal evolution of this cascade. I propose that this simple and generic model – which can in essence fit any activation cascade – offers a simple mathematical framework to understand the temporal dynamics of signaling cascades, and the delay and change in the shape of the response which can be observed between the input and the output of a cascade.
... DIAPH3 is essential for cell division, and several studies have emphasized its role in cytokinesis (7)(8)(9). More recently, DIAPH3 has also been shown to be crucial for karyokinesis, specifically for mitotic spindle organization (10) and activation of the spindle assembly checkpoint (11). ...
Article
Full-text available
Background Glioblastoma is one of the most aggressive primary brain tumors, with a poor outcome despite multimodal treatment. Methylation of the MGMT promoter, which predicts the response to temozolomide, is a well-established prognostic marker for glioblastoma. However, a difference in survival can still be detected within the MGMT methylated group, with some patients exhibiting a shorter survival than others, emphasizing the need for additional predictive factors. Methods We analyzed DIAPH3 expression in glioblastoma samples from the cancer genome atlas (TCGA). We also retrospectively analyzed one hundred seventeen histological glioblastomas from patients operated on at Saint-Luc University Hospital between May 2013 and August 2019. We analyzed the DIAPH3 expression, explored the relationship between mRNA levels and Patient’s survival after the surgical resection. Finally, we assessed the methylation pattern of the DIAPH3 promoter using a targeted deep bisulfite sequencing approach. Results We found that 36% and 1% of the TCGA glioblastoma samples exhibit copy number alterations and mutations in DIAPH3, respectively. We scrutinized the expression of DIAPH3 at single cell level and detected an overlap with MKI67 expression in glioblastoma proliferating cells, including neural progenitor-like, oligodendrocyte progenitor-like and astrocyte-like states. We quantitatively analyzed DIAPH3 expression in our cohort and uncovered a positive correlation between DIAPH3 mRNA level and patient’s survival. The effect of DIAPH3 was prominent in MGMT-methylated glioblastoma. Finally, we report that the expression of DIAPH3 is at least partially regulated by the methylation of three CpG sites in the promoter region. Conclusion We propose that combining the DIAPH3 expression with MGMT methylation could offer a better prediction of survival and more adapted postsurgical treatment for patients with MGMT-methylated glioblastoma.
... In addition, ubiquitylation of mDia2 is largely dependent on the cell cycle [44]. Dysregulation of its expression or activity has been demonstrated to cause cytokinesis failure in tumor cells [44], fibroblasts [45], and fetal and adult erythroblasts [23,25]. Specifically, By modulating the incorporation of the centromere-specific histone H3 variant, CENP-A, and nuclear actin polymerization, mDia2 governs the centromere movement [46,47]. ...
Article
Full-text available
mDia formin proteins regulate the dynamics and organization of the cytoskeleton through their linear actin nucleation and polymerization activities. We previously showed that mDia1 deficiency leads to aberrant innate immune activation and induces myelodysplasia in a mouse model, and mDia2 regulates enucleation and cytokinesis of erythroblasts and the engraftment of hematopoietic stem and progenitor cells (HSPCs). However, whether and how mDia formins interplay and regulate hematopoiesis under physiological and stress conditions remains unknown. Here, we found that both mDia1 and mDia2 are required for HSPC regeneration under stress, such as serial plating, aging, and reconstitution after myeloid ablation. We showed that mDia1 and mDia2 form hetero-oligomers through the interactions between mDia1 GBD-DID and mDia2 DAD domains. Double knockout of mDia1 and mDia2 in hematopoietic cells synergistically impaired the filamentous actin network and serum response factor-involved transcriptional signaling, which led to declined HSPCs, severe anemia, and significant mortality in neonates and newborn mice. Our data demonstrate the potential roles of mDia hetero-oligomerization and their non-rodent functions in the regulation of HSPCs activity and orchestration of hematopoiesis.
... The central spindle activates RhoA at the equatorial cortex (Bement et al., 2005), while astral microtubules inhibit its activation at the peripheral cortex (Dechant and Glotzer, 2003;Zanin et al., 2013). Activated RhoA promotes actomyosin assembly at the equatorial cortex and subsequent ring contraction (Castrillon and Wasserman, 1994;Matsumura, 2005;Watanabe et al., 2008). ...
Article
Cell division involves separating the genetic material and cytoplasm of a mother cell into two daughter cells. The last step of cell division, abscission, consists of cutting the cytoplasmic bridge, a microtubule-rich membranous tube connecting the two cells, which contains the midbody, a dense proteinaceous structure. Canonically, abscission occurs 1-3 h after anaphase. However, in certain cases, abscission can be severely delayed or incomplete. Abscission delays can be caused by mitotic defects that activate the abscission 'NoCut' checkpoint in tumor cells, as well as when cells exert abnormally strong pulling forces on the bridge. Delayed abscission can also occur during normal organism development. Here, we compare the mechanisms triggering delayed and incomplete abscission in healthy and disease scenarios. We propose that NoCut is not a bona fide cell cycle checkpoint, but a general mechanism that can control the dynamics of abscission in multiple contexts.
... The reason for this partly different localization pattern is currently unclear, but could be due to the different fixation methods used. The function of mDia2 in cytokinesis is further aided by an abnormal configuration of the spindle in mDia2-depleted cells [78] as well as the analysis of mice with global knockout of mDia2, which proved to be embryonically lethal [77]. Thus, although we cannot formally exclude the presence of minute amounts of mDia2 on filopodia tips that are below the detection limit of antibody staining, collectively our data do not support a contribution of this formin to filopodia formation. ...
Article
Full-text available
Background: Filopodia are dynamic, finger-like actin-filament bundles that overcome membrane tension by forces generated through actin polymerization at their tips to allow extension of these structures a few microns beyond the cell periphery. Actin assembly of these protrusions is regulated by accessory proteins including heterodimeric capping protein (CP) or Ena/VASP actin polymerases to either terminate or promote filament growth. Accordingly, the depletion of CP in B16-F1 melanoma cells was previously shown to cause an explosive formation of filopodia. In Ena/VASP-deficient cells, CP depletion appeared to result in ruffling instead of inducing filopodia, implying that Ena/VASP proteins are absolutely essential for filopodia formation. However, this hypothesis was not yet experimentally confirmed. Methods: Here, we used B16-F1 cells and CRISPR/Cas9 technology to eliminate CP either alone or in combination with Ena/VASP or other factors residing at filopodia tips, followed by quantifications of filopodia length and number. Results: Unexpectedly, we find massive formations of filopodia even in the absence of CP and Ena/VASP proteins. Notably, combined inactivation of Ena/VASP, unconventional myosin-X and the formin FMNL3 was required to markedly impair filopodia formation in CP-deficient cells. Conclusions: Taken together, our results reveal that, besides Ena/VASP proteins, numerous other factors contribute to filopodia formation.
... Furthermore, both Rac and Cdc42, upstream regulators of Arp2/3 (ref. 28 ), are enriched in the posterior cortex (Extended Data Fig. 3a-d), but not Diaphanous, which is the main Formin known to nucleate Actin at the cell cortex and to be involved during cytokinesis 21,[29][30][31] (Extended Data Fig. 3e,f). Additionally, only Filamin (a large, flexible crosslinker arranging Actin into a meshwork) 32,33 is found enriched at the cortical poles, but not Fascin, α-Actinin or Fimbrin, all of which organize F-Actin into bundles [34][35][36] (Extended Data Figs. ...
Article
Full-text available
The control of cell shape during cytokinesis requires a precise regulation of mechanical properties of the cell cortex. Only few studies have addressed the mechanisms underlying the robust production of unequal-sized daughters during asymmetric cell division. Here we report that unequal daughter-cell sizes resulting from asymmetric sensory organ precursor divisions in Drosophila are controlled by the relative amount of cortical branched Actin between the two cell poles. We demonstrate this by mistargeting the machinery for branched Actin dynamics using nanobodies and optogenetics. We can thereby engineer the cell shape with temporal precision and thus the daughter-cell size at different stages of cytokinesis. Most strikingly, inverting cortical Actin asymmetry causes an inversion of daughter-cell sizes. Our findings uncover the physical mechanism by which the sensory organ precursor mother cell controls relative daughter-cell size: polarized cortical Actin modulates the cortical bending rigidity to set the cell surface curvature, stabilize the division and ultimately lead to unequal daughter-cell size.
... In addition to the above two structural domains, some formin protein sequences contain some or all of the GBD, Diaphanous Inhibitory Domain (DID) and Diaphanous Autoregulatory Domain (DAD). The role of Formins in regulating the formation of unbranched actin filaments affects activities such as formation of cellular protrusions (7,142,143) and cytokinesis (144). In addition, Formins can regulate microtubule dynamics and link them to actin dynamics to coordinate cytoskeletal activities (145-147). ...
Article
Full-text available
Actin is an important cytoskeletal protein involved in signal transduction, cell structure and motility. Actin regulators include actin-monomer-binding proteins, Wiskott-Aldrich syndrome (WAS) family of proteins, nucleation proteins, actin filament polymerases and severing proteins. This group of proteins regulate the dynamic changes in actin assembly/disassembly, thus playing an important role in cell motility, intracellular transport, cell division and other basic cellular activities. Lymphocytes are important components of the human immune system, consisting of T-lymphocytes (T cells), B-lymphocytes (B cells) and natural killer cells (NK cells). Lymphocytes are indispensable for both innate and adaptive immunity and cannot function normally without various actin regulators. In this review, we first briefly introduce the structure and fundamental functions of a variety of well-known and newly discovered actin regulators, then we highlight the role of actin regulators in T cell, B cell and NK cell, and finally provide a landscape of various diseases associated with them. This review provides new directions in exploring actin regulators and promotes more precise and effective treatments for related diseases.
... Tight spatiotemporal regulation and activation of RHOA by ECT2 at the cleavage furrow are essential for the initial stages of cytokinesis (171). Other RHOA GEFs ensure efficient division, such as GEF-H1, MYOGEF, and VAV3 (196), while RhoA coordinates mDia2-mediated formation of linear filamentous actin with Myosin II to promote the assembly and constriction of a contractile ring (197). RHOA drives contraction of the actin ring through the stimulation of MLC2 phosphorylation by ROCK in glioblastoma cells (198), while contribution to MLC2 phosphorylation by myosin light chain kinase (MLCK), Citron, and through MLCP inactivation (by ROCK and Aurora-B) have also been reported (199). ...
Article
Full-text available
Rho GTPases are a family of small G proteins that regulate a wide array of cellular processes related to their key roles controlling the cytoskeleton. On the other hand, cancer is a multi-step disease caused by the accumulation of genetic mutations and epigenetic alterations, from the initial stages of cancer development when cells in normal tissues undergo transformation, to the acquisition of invasive and metastatic traits, responsible for a large number of cancer related deaths. In this review, we discuss the role of Rho GTPase signalling in cancer in every step of disease progression. Rho GTPases contribute to tumour initiation and progression, by regulating proliferation and apoptosis, but also metabolism, senescence and cell stemness. Rho GTPases play a major role in cell migration, and in the metastatic process. They are also involved in interactions with the tumour microenvironment and regulate inflammation, contributing to cancer progression. After years of intensive research, we highlight the importance of relevant models in the Rho GTPase field, and we reflect on the therapeutic opportunities arising for cancer patients.
Article
During cytokinesis, a contractile ring consisting of unbranched filamentous actin (F-actin) and myosin II constricts at the cell equator. Unbranched F-actin is generated by formin, and without formin no cleavage furrow forms. In Caenorhabditis elegans, depletion of septin restores furrow ingression in formin mutants. How the cleavage furrow ingresses without a detectable unbranched F-actin ring is unknown. We report that, in this setting, anillin (ANI-1) forms a meshwork of circumferentially aligned linear structures decorated by non-muscle myosin II (NMY-2). Analysis of ANI-1 deletion mutants reveals that its disordered N-terminal half is required for linear structure formation and sufficient for furrow ingression. NMY-2 promotes the circumferential alignment of the linear ANI-1 structures and interacts with various lipids, suggesting that NMY-2 links the ANI-1 network with the plasma membrane. Collectively, our data reveal a compensatory mechanism, mediated by ANI-1 linear structures and membrane-bound NMY-2, that promotes furrowing when unbranched F-actin polymerization is compromised.
Article
Tetraploidy is a hallmark of broad cancer types, but it remains largely unknown which aspects of cellular processes are influenced by tetraploidization in human cells. Here, we found that tetraploid HCT116 cells manifested severe cell shape instability during cytokinesis, unlike their diploid counterparts. The cell shape instability accompanied the formation of protrusive deformation at the cell poles, indicating ectopic contractile activity of the cell cortex. While cytokinesis regulators such as RhoA and anillin correctly accumulated at the equatorial cortex, myosin II was over-accumulated at the cell poles, specifically in tetraploid cells. Suppression of myosin II activity by Y27632 treatment restored smooth cell shape in tetraploids during cytokinesis, indicating dysregulation of myosin II as a primary cause of the cell shape instability in the tetraploid state. Our results demonstrate a new aspect of the dynamic cellular process profoundly affected by tetraploidization in human cells, which provides a clue to molecular mechanisms of tetraploidy-driven pathogenic processes.
Article
Full-text available
Mammalian Diaphanous (mDia)-related formins and the N-WASP-activated Arp2/3 complex initiate the assembly of filamentous actin. Dia-interacting protein (DIP) binds via its amino-terminal SH3 domain to the proline-rich formin homology 1 (FH1) domain of mDia1 and mDia2 and to the N-WASp proline-rich region. Here, we investigated an interaction between a conserved leucine-rich region (LRR) in DIP and the mDia FH2 domain that nucleates, processively elongates, and bundles actin filaments. DIP binding to mDia2 was regulated by the same Rho-GTPase-controlled autoinhibitory mechanism modulating formin-mediated actin assembly. DIP was previously shown to interact with and stimulate N-WASp-dependent branched filament assembly via Arp2/3. Despite direct binding to both mDia1 and mDia2 FH2 domains, DIP LRR inhibited only mDia2-dependent filament assembly and bundling in vitro. DIP expression interfered with filopodia formation, consistent with a role for mDia2 in assembly of these structures. After filopodia retraction into the cell body, DIP expression induced excessive nonapoptotic membrane blebbing, a physiological process involved in both cytokinesis and amoeboid cell movement. DIP-induced blebbing was dependent on mDia2 but did not require the activities of either mDia1 or Arp2/3. These observations point to a pivotal role for DIP in the control of nonbranched and branched actin-filament assembly that is mediated by Diaphanous-related formins and activators of Arp2/3, respectively. The ability of DIP to trigger blebbing also suggests a role for mDia2 in the assembly of cortical actin necessary for maintaining plasma-membrane integrity.
Article
Full-text available
To identify potential RhoA effector proteins, we conducted a two-hybrid screen for cDNAs encoding proteins that interact with a Gal4-RhoA.V14 fusion protein. In addition to the RhoA effector ROCK-I we identified cDNAs encoding Kinectin, mDia2 (a p140 mDia-related protein), and the guanine nucleotide exchange factor, mNET1. ROCK-I, Kinectin, and mDia2 can bind the wild type forms of both RhoA and Cdc42 in a GTP-dependent manner in vitro. Comparison of the ROCK-I and Kinectin sequences revealed a short region of sequence homology that is both required for interaction in the two-hybrid assay and sufficient for weak interaction in vitro. Sequences related to the ROCK-I/Kinectin sequence homology are present in heterotrimeric G protein β subunits and in theSaccharomyces cerevisiae Skn7 protein. We show that β2 and Skn7 can interact with mammalian RhoA and Cdc42 and yeast Rho1, both in vivo and in vitro. Functional assays in yeast suggest that the Skn7 ROCK-I/Kinectin homology region is required for its function in vivo.
Article
Full-text available
We show that the Drosophila gene diaphanous is required for cytokinesis. Males homozygous for the dia1 mutation are sterile due to a defect in cytokinesis in the germline. Females trans-heterozygous for dia1 and a deficiency are sterile and lay eggs with defective eggshells; failure of cytokinesis is observed in the follicle cell layer. Null alleles are lethal. Death occurs at the onset of pupation due to the absence of imaginal discs. Mitotic figures in larval neuroblasts were found to be polyploid, apparently due to a defect in cytokinesis. The predicted 123 x 10(3) M(r) protein contains two domains shared by the formin proteins, encoded by the limb deformity gene in the mouse. These formin homology domains, which we have termed FH1 and FH2, are also found in Bni1p, the product of a Saccharomyces cerevisiae gene required for normal cytokinesis in diploid yeast cells.
Article
Full-text available
Rho small GTPase regulates cell morphology, adhesion and cytokinesis through the actin cytoskeleton. We have identified a protein, p140mDia, as a downstream effector of Rho. It is a mammalian homolog of Drosophila diaphanous, a protein required for cytokinesis, and belongs to a family of formin-related proteins containing repetitive polyproline stretches. p140mDia binds selectively to the GTP-bound form of Rho and also binds to profilin. p140mDia, profilin and RhoA are co-localized in the spreading lamellae of cultured fibroblasts. They are also co-localized in membrane ruffles of phorbol ester-stimulated sMDCK2 cells, which extend these structures in a Rho-dependent manner. The three proteins are recruited around phagocytic cups induced by fibronectin-coated beads. Their recruitment is not induced after Rho is inactivated by microinjection of botulinum C3 exoenzyme. Overexpression of p140mDia in COS-7 cells induced homogeneous actin filament formation. These results suggest that Rho regulates actin polymerization by targeting profilin via p140mDia beneath the specific plasma membranes.
Article
Kyoto University (京都大学) 0048 新制・課程博士 博士(医学) 甲第12878号 医博第3038号 新制/医/938 UT51-2007-H151 2007-03-23 京都大学大学院医学研究科脳統御医科学系専攻 (主査)教授 影山 龍一郎, 教授 鍋島 陽一, 教授 大森 治紀 学位規則第4条第1項該当
Article
The cleavage-furrow region was isolated surgically from newt eggs at the early stage of the first cleavage. The isolated furrow contracted in the presence of ATP at a Ca2+ concentration of 10 or 100 nM, although the speed was less than that of the furrow in vivo. Cytochalasin B, cytochalasin D, phalloidin, p-chloromercuribenzoate, and N-ethyl-maleimide interfered with the contraction, but colchicine did not. The furrow contained bundles of actin filaments of opposite polarities oriented parallel to the long axis of the furrow; these bundles may be the main component of the contractile arc. From electron microscopic observation of thin sections of the furrow, it was suggested that the actin bundles of the contractile arc were organized from preexisting cortical filaments that were connected to the plasma membrane by granular materials at their barbed ends. Contractile-arc actin filaments were revealed to be crosslinked by thin strands by the rapid freezing/deep etching-replication technique. Two-dimensional polyacrylamide gel electrophoresis showed that several proteins found in the furrow cortex are absent from the cortical layer before the cleavage furrow is formed.
Article
Sand dollar eggs were microinjected with botulinum C3 exoenzyme, an ADP-ribosyltransferase from Clostridium botulinum that specifically ADP-ribosylates and inactivates rho proteins. C3 exoenzyme microinjected during nuclear division interfered with subsequent cleavage furrow formation. No actin filaments were detected in the equatorial cortical layer of these eggs by rhodamine-phalloidin staining. When microinjected into furrowing eggs, C3 exoenzyme rapidly disrupted the contractile ring actin filaments and caused regression of the cleavage furrows. C3 exoenzyme had no apparent effect on nuclear division, however, and multinucleated embryos developed from the microinjected eggs. By contrast, C3 exoenzyme did not affect the organisation of cortical actin filaments immediately after fertilisation. Only one protein (molecular weight 22,000) was ADP-ribosylated by C3 exoenzyme in the isolated cleavage furrow. This protein co-migrated with ADP-ribosylated rhoA derived from human platelets when analysed by two-dimensional gel electrophoresis. These results strongly suggest that a rho-like, small GTP-binding protein is selectively involved in the organisation and maintenance of the contractile ring.
Article
During mitosis, a ring containing actin and myosin appears beneath the equatorial surface of animal cells. This ring then contracts, forms a cleavage furrow and divides the cell, a step known as cytokinesis. The two daughter cells often remain connected by an intercellular bridge which contains a refringent structure known as the midbody. How the appearance of this ring is regulated is unclear, although the small GTPase Rho, which controls the formation of actin structures, is known to be essential. Protein kinases are also thought to participate in cytokinesis. We now show that a splice variant of a Rho target protein, named citron, contains a protein kinase domain that is related to the Rho-associated kinases ROCK14 and ROK, which regulate myosin-based contractility. Citron kinase localizes to the cleavage furrow and midbody of HeLa cells; Rho is also localized in the midbody. We find that overexpression of citron mutants results in the production of multinucleate cells and that a kinase-active mutant causes abnormal contraction during cytokinesis. We propose that citron kinase regulates cytokinesis at a step after Rho in the contractile process.
Article
We have examined the role of the mouse Diaphanous-related formin (DRF) Rho GTPase binding proteins, mDia1 and mDia2, in cell regulation. The DRFs are required for cytokinesis, stress fiber formation, and transcriptional activation of the serum response factor (SRF). 'Activated' mDia1 and mDia2 variants, lacking their GTPase binding domains, cooperated with Rho-kinase or ROCK to form stress fibers but independently activated SRF. Src tyrosine kinase associated and co-localized with the DRFs in endosomes and in mid-bodies of dividing cells. Inhibition of Src also blocked cytokinesis, SRF induction by activated DRFs, and cooperative stress fiber formation with active ROCK. Our results show that the DRF proteins couple Rho and Src during signaling and the regulation of actin dynamics.