ArticlePDF Available

Structure and Biochemical Properties of Fission Yeast Arp2/3 Complex Lacking the Arp2 Subunit

Authors:

Abstract

Arp2/3 (actin-related protein 2/3) complex is a seven-subunit complex that nucleates branched actin filaments in response to cellular signals. Nucleation-promoting factors such as WASp/Scar family proteins activate the complex by facilitating the activating conformational change and recruiting the first actin monomer for the daughter branch. Here we address the role of the Arp2 subunit in the function of Arp2/3 complex by isolating a version of the complex lacking Arp2 (Arp2Δ Arp2/3 complex) from fission yeast. An x-ray crystal structure of the ΔArp2 Arp2/3 complex showed that the rest of the complex is unperturbed by the loss of Arp2. However, the Arp2Δ Arp2/3 complex was inactive in actin nucleation assays, indicating that Arp2 is essential to form a branch. A fluorescence anisotropy assay showed that Arp2 does not contribute to the affinity of the complex for Wsp1-VCA, a Schizosaccharomyces pombe nucleation-promoting factor protein. Fluorescence resonance energy transfer experiments showed that the loss of Arp2 does not prevent VCA from recruiting an actin monomer to the complex. Truncation of the N terminus of ARPC5, the smallest subunit in the complex, increased the yield of Arp2Δ Arp2/3 complex during purification but did not compromise nucleation activity of the full Arp2/3 complex.
Structure and Biochemical Properties of Fission Yeast Arp2/3
Complex Lacking the Arp2 Subunit
*
S
Received for publication, April 3, 2008, and in revised form, June 17, 2008 Published, JBC Papers in Press, July 18, 2008, DOI 10.1074/jbc.M802607200
Brad J. Nolen
1
and Thomas D. Pollard
From the Departments of Molecular, Cellular and Developmental Biology, Cell Biology, and Molecular Biophysics and
Biochemistry, Yale University, New Haven, Connecticut 06520-8103
Arp2/3 (actin-related protein 2/3) complex is a seven-subunit
complex that nucleates branched actin filaments in response to
cellular signals. Nucleation-promoting factors such as WASp/
Scar family proteins activate the complex by facilitating the acti-
vating conformational change and recruiting the first actin
monomer for the daughter branch. Here we address the role of
the Arp2 subunit in the function of Arp2/3 complex by isolating
a version of the complex lacking Arp2 (Arp2Arp2/3 complex)
from fission yeast. An x-ray crystal structure of the Arp2
Arp2/3 complex showed that the rest of the complex is unper-
turbed by the loss of Arp2. However, the Arp2Arp2/3 complex
was inactive in actin nucleation assays, indicating that Arp2 is
essential to form a branch. A fluorescence anisotropy assay
showed that Arp2 does not contribute to the affinity of the com-
plex for Wsp1-VCA, a Schizosaccharomyces pombe nucleation-
promoting factor protein. Fluorescence resonance energy trans-
fer experiments showed that the loss of Arp2 does not prevent
VCA from recruiting an actin monomer to the complex. Trun-
cation of the N terminus of ARPC5, the smallest subunit in the
complex, increased the yield of Arp2Arp2/3 complex during
purification but did not compromise nucleation activity of the
full Arp2/3 complex.
Dynamic rearrangements of the actin cytoskeleton are essen-
tial for cellular responses to the environment, and cells employ
a host of proteins to control nucleation, polymerization, cap-
ping, severing, bundling, cross-linking, and depolymerization
of actin (1). Fission yeast have two well characterized nucleators
of actin filaments, formins (Cdc12 and For3) and Arp2/3 com-
plex (consisting of seven-subunits, Arp3, Arp2, and ARPC1–5)
(2). Formins assemble unbranched filaments present both in
cables that span the length of interphase cells and in the con-
tractile ring, which constricts during cytokinesis (3, 4). Arp2/3
complex nucleates branched filaments in cortical structures
called actin patches, which are sites of endocytosis located at
cell poles during interphase and the cleavage furrow during
cytokinesis (5–7).
Arp2/3 complex nucleates actin filament branches on the
sides of pre-existing (mother) filaments (8). The new (daughter)
filament grows at an angle of 78° on the side of the mother
filament (9). Arp2/3 complex from most species is intrinsically
inactive but is stimulated to form an actin filament branch
through interactions with proteins called nucleation promoting
factors (NPFs),
2
ATP, an actin monomer, and the side of a
mother filament (2). WASp/Scar family proteins, the prototyp-
ical NPFs, contain a C-terminal region termed VCA (verprolin
homology, central, acidic), which is the minimal fragment
required to activate nucleation by Arp2/3 complex. Crystallo-
graphic and electron microscopic data suggest that a conforma-
tional change reorients the two actin-related subunits, Arp2
and Arp3, like two successive actin subunits along the short
pitch helix of an actin filament to create the nucleus for polym-
erization of the daughter filament (9, 10). A model built by
fitting crystal structures into reconstructions of electron tomo-
grams of branch junctions shows that all seven subunits of
Arp2/3 complex contact the mother filament and that the
barbed ends of Arp2 and Arp3 interact with the pointed end of
the daughter filament (9).
Numerous questions remain about the mechanism of
branching nucleation despite many structural and kinetic stud-
ies. Nucleotide binding favors a conformation of Arp3 that may
contribute to activation (11, 12). The V region of NPFs binds an
actin monomer (13, 14), recruiting it to the branch point,
whereas the C and A regions bind to Arp2/3 complex and are
thought to facilitate conformational changes required for
nucleation (14, 15). Cross-linking, radiation footprinting, and
NMR experiments have implicated all but two subunits
(ARPC2 and ARPC4) in interactions with VCA (16 –19), but no
high resolution structural information on VCA binding is avail-
able. A model based on small angle x-ray scattering of Arp2/3
complex bound to actin and VCA has the actin monomer
located at the barbed end of Arp2 (20). Expression of recombi-
nant human Arp2/3 complex subunits in insect cells demon-
strated that ARPC2 and ARPC4 are essential for the integrity of
the complex and that both of these subunits and Arp3 are nec-
essary to assemble a complex that can nucleate actin filaments
*This work was supported, in whole or in part, by National Institutes of Health
Grant GM066311. This work was also supported by National Institutes of
Health Ruth Kirschstein postdoctoral fellowship GM074374-02. The costs
of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked advertisement”in
accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
The atomic coordinates and structure factors (code 3DWL) have been deposited
in the Protein Data Bank, Research Collaboratory for Structural Bioinformat-
ics, Rutgers University, New Brunswick, NJ (http://www.rcsb.org/).
S
The on-line version of this article (available at http://www.jbc.org) contains
supplemental text and Figs. S1–S4.
1
To whom correspondence should be addressed: Tel.: 203-432-3194; Fax:
203-432-6161; E-mail: bradley.nolen@yale.edu.
2
The abbreviations used are: NPF, nucleation promoting factor; DTT, dithio-
threitol; GST, glutathione S-transferase; OG-actin, Oregon Green-actin;
FRET, fluorescence resonance energy transfer; YFP, yellow fluorescent pro-
tein; CFP, cyan fluorescent protein.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 283, NO. 39, pp. 26490–26498, September 26, 2008
© 2008 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.
26490 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 283NUMBER 39SEPTEMBER 26, 2008
by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from by guest on November 7, 2015http://www.jbc.org/Downloaded from
(21). This study did not address the role of Arp2 in the stability
and nucleation activity of Arp2/3 complex.
Here, we show that a complex lacking the Arp2 subunit
(Arp2 Arp2/3 complex) can be isolated from fission yeast.
Except for the absence of Arp2, the loss of Arp2 did not perturb
the crystal structure of Arp2/3 complex. Arp2 Arp2/3 com-
plex bound Wsp1-VCA but did not nucleate actin filaments, so
Arp2 is essential to initiate a branch. Mutational analysis
showed that branching nucleation does not require the N ter-
minus of ARPC5 to anchor Arp2 as proposed in one model of
activation (22).
EXPERIMENTAL PROCEDURES
Purification of S. pombe Arp2/3 Complex—We purified
native Arp2/3 complex from Schizosaccharomyces pombe
strain TM011 typically starting with 500 g of wet cells. Ten
milliliters of a turbid culture of cells was inoculated per liter of
media made from 35 g/liter YES (Q-biogene) and grown with
vigorous shaking overnight at 30 °C. In the morning, an addi-
tional 70 g/liter YES was added, and the cultures were grown to
an optical density of 6.0 at 600 nm. Cells were then pelleted
and washed in lysis buffer (20 mMTris, pH 8.0, 50 mMNaCl, and
1m
MEDTA). Pellets of cells were resuspended in 1 ml of lysis
buffer per gram of wet cells and stored at 80 °C until lysis.
After thawing, an additional 0.4 ml of lysis buffer plus DTT (to
1m
M) and Complete protease inhibitor tablets (Roche Applied
Science, 1 tablet per 50 ml) were added per gram of cells. All
subsequent steps were at 04 °C. Cells were lysed using a
Microfluidizer (Microfluidics, model 110EH). Phenylmethyl-
sulfonyl fluoride was added to 1 mM, and the lysate was centri-
fuged at 30,000 gfor 20 min. The supernatant was centri-
fuged a second time at 125,000 gfor 1 h, filtered through
cheesecloth, and loaded onto a 150-ml Q-Sepharose column
equilibrated with lysis buffer. The column was washed with 1.5
column volumes of lysis buffer containing 1 mMDTT, and pro-
tein eluted with 20 mMTris, pH 8.0, 300 mMNaCl, 1 mMEDTA,
1m
MDTT. Protein was precipitated from the eluted fractions
with 40% ammonium sulfate, pelleted, and resuspended in 50
ml of PKME (25 mMPIPES, pH 7.0, 50 mMKCl, 1 mMEGTA, 3
mMMgCl
2
,1mMDTT, and 0.1 mMATP) and dialyzed over-
night against the same buffer. The sample was then loaded onto
an 8-ml column of glutathione-Sepharose 4B (GE Healthcare)
pre-charged with GST-N-WASp-VCA and equilibrated with
PKME. GST-N-WASp-VCA was purified as described for
GST-WASp-VCA (23). Arp2/3 complex was eluted with a lin-
ear 60-ml gradient of 50–1000 mMNaCl. Peak fractions were
pooled and dialyzed against 20 mMTris, pH 8.0, 100 mMNaCl,
1m
MEGTA, 1 mMDTT, and 1 mMMgCl
2
and fractionated on
a 0.5- 5-cm column of Mono Q (GE Healthcare) equilibrated
with 20 mMTris, pH 8.0, 100 mMNaCl, 1 mMEGTA, 1 mM
DTT, and 1 mMMgCl
2
and eluted with a linear 20-ml gradient
of 100400 mMNaCl. Pooled fractions were concentrated to
1.0 ml with Amicon Ultra-15 concentrators (Millipore) and
loaded onto a Superdex 200 HR16/60 gel-filtration column (GE
Healthcare) equilibrated with 20 mMTris, pH 8.0, 100 mM
NaCl, and 1 mMDTT. Peak fractions were pooled and diluted
2-fold with 20 mMTris, pH 8.0, and 1 mMDTT before concen-
trating and flash freezing in liquid nitrogen. Protein concentra-
tion was determined by measuring absorbance at 280 nm with
extinction coefficients determined by the Von Hippel method
(24).
Purification of S. pombe Wsp1-VCA Constructs—pGV67-
Wsp1-VCA was constructed by amplifying S. pombe Wsp1 res-
idues 497–574 with 5-BamHI and 3-EcoRI restriction sites
and cloning into pGV67, a plasmid derived from p21d (Nova-
gen) and containing an N-terminal glutathione S-transferase
(GST) tag followed by a tobacco etch virus cleavage site.
pGV67-Wsp1-Cys-VCA was constructed using the same pro-
cedure, with a single cysteine inserted 5to the 497–574
fragment.
Escherichia coli strain BL21(DE3) was transformed with
either pGV67-Wsp1-VCA or pGV67-Wsp1-Cys-VCA, grown
to A
600 nm
of 0.8, and induced with 0.4 mMisopropyl 1-thio-
-
D-galactopyranoside for overnight expression at 22 °C. Cells
from an 8-liter culture were harvested and lysed in 300 ml of 20
mMTris, pH 8.0, 25 mMNaCl, 2 mMDTT, and 1 mMphenyl-
methylsulfonyl fluoride containing four protease inhibitor tab-
lets (Roche Applied Science) per 50 ml. Cells were lysed with a
Branson sonicator, and the lysate was cleared by centrifugation
at 100,000 gfor 30 min. Supernatant was loaded onto a 35-ml
column DEAE-Sepharose equilibrated in lysis buffer and eluted
with a linear 300-ml gradient of 25–700 mMNaCl. Fractions
containing recombinant protein were pooled and diluted with
lysis buffer lacking NaCl to reduce the salt to 140 mM. The
sample was then loaded onto a 10-ml column of glutathione-
Sepharose 4B equilibrated with 20 mMTris, pH 8.0, 140 mM
NaCl, 2 mMEDTA, and 2 mMDTT (binding buffer) and washed
with 5 column volumes of the same buffer. The wash buffer was
allowed to drain to the top of the column bed, and 30
lof75
Mtobacco etch virus protease (purified from E. coli by expres-
sion from the pRK1043 tobacco etch virus vector kindly pro-
vided by D. Waugh (25)) was mixed into the glutathione-Sepha-
rose slurry. Protein was released from the column during
incubation with gentle rocking overnight at 4 °C. Eluted protein
was pooled and loaded onto a 0.5- 5-cm Mono Q column
equilibrated with 20 mMTris, pH 8.0, 100 mMNaCl, and 2 mM
DTT. Protein was eluted with a 20-ml linear gradient of 100
800 mMNaCl, concentrated, and purified by size-exclusion
chromatography onto a Superdex 75 HR16/60 column (GE
Healthcare) in 20 mMTris, pH 8.0, 100 mMNaCl, and 1 mM
DTT before re-concentrating and flash freezing in liquid
nitrogen.
Preparation of Rhodamine-labeled Sp-Wsp1-Cys-VCA—Sp-
Wsp1-Cys-VCA was purified as the SpWsp1-VCA construct,
except that DTT was excluded from the final gel filtration
buffer. A fresh 20 mMstock of tetramethylrhodamine-6-male-
imide (Invitrogen) in dimethylformamide was added dropwise
to the pooled peak fractions of Sp-Wsp1-Cys-VCA while stir-
ring on ice. Reaction was stopped after1hbytheaddition of 1
mMDTT. Labeled protein was separated from free dye by ion-
exchange chromatography on a 0.5- 5-cm Mono Q column
and through repeated concentration and dilution in an Amicon
Ultra protein concentrator. Concentration of SpWsp1-Rho-
VCA (Rho-VCA) was determined by measuring absorbance at
552 nm with an extinction coefficient of 44,660 M
1
cm
1
(14).
Arp2/3 Complex Crystal Structures
SEPTEMBER 26, 2008VOLUME 283NUMBER 39 JOURNAL OF BIOLOGICAL CHEMISTRY 26491
by guest on November 7, 2015http://www.jbc.org/Downloaded from
Preparation of Labeled and Unlabeled and Actin Monomers
We purified actin by established procedures: chicken skeletal
muscle actin (26), pyrene-labeled actin (27), and Oregon
Green-actin (OG-actin) (28).
Construction of Mutant Strains of S. pombe—We constructed
mutant strains by PCR-based gene targeting (29). For ARPC5
mutants, cassettes containing a KanMX6 module and the
P3nmt1 thymine-repressible promoter were amplified with
long primers to generate a 5-flanking sequence complimentary
to a region 330 nucleotides upstream from the ARPC5 start
codon. The 3-flanking sequence was complimentary to either
the first 60 nucleotides of the ARPC5 coding region (with codon
3 mutated to change arginine to glutamic acid) or to a region 42
nucleotides downstream from the start codon (for ARPC5N).
Cassettes were transformed into the TM011 strain of S. pombe
using the lithium acetate method (29, 30). Transformed cells
were plated on YES plates and incubated for 18hat3C
before replica plating onto YES plates containing 100 mg/liter
G418/Geneticin (Invitrogen). Replica plates were incubated for
2–3 days, and large colonies were re-streaked on fresh YES
plates. Genomic DNA was isolated from mutant strains (29),
and the junctions and the entire coding region of ARPC5 were
sequenced to verify the presence of the mutations. ARPC3-YFP
and arp2-CFP strains were provided by Chris Beltzner, and
were constructed by replacing the ARPC3 or arp2 stop codon in
FY527 (h
leu1–32 ura4-D18 his3-D1 ade6-M216) or FY528
(h
leu1–32 ura4-D18 his3-D1 ade6-M210) with a pFA6a-
YFP-kanMX6 or pFA6a-CFP-kanMX6 cassette, crossing, and
analyzing tetrads.
Fluorescence Anisotropy Assays—We used an Alphascan flu-
orometer with a T-format for fluorescence measurements
(Photon Technology International). Samples were excited at
552 nm, and anisotropy was measured at 578 nm. The excita-
tion monochromometer bandwidth was 6 nm, and the emission
bandwidth was 4 nm. Fixed concentrations of Rho-VCA were
titrated with SpArp2/3 complex, and the fluorescence was
measured for both filter positions at 0.1 points/s for 40 s. Pho-
ton Technology International Felix software was used for data
collection and to calculate average anisotropy. For FRET meas-
urements fixed concentrations of OG-actin were titrated with
Rho-VCA with excitation at 480 nm, emission at 517 nm, and 3
nm slit widths.
Actin Polymerization Assays—We measured the nucleation
activity of wild-type and mutant Arp2/3 complex from the time
course of actin polymerization. Polymerization reactions of 100
l were assembled as follows: 2
lof10mMEGTA and 1 mM
MgCl
2
were added to 20
lof20
M15% pyrene actin mono-
mers in G-buffer (2 mMTris-HCl, pH 8.0, 0.2 mMATP, 0.1 mM
CaCl
2
, 0.5 mMDTT, and 0.01% NaN
3
) followed immediately by
adding 78
l of a solution containing Arp2/3 complex,
SpWsp1-VCA with buffers, and salts to bring the final reaction
conditions to 10 mMimidazole, pH 7.0, 50 mMKCl, 1 mM
EGTA, 1 mMMgCl
2
. Fluorescence measurements were made at
15-s intervals in a 96-well plate using a Gemini XPS spectroflu-
orometer (Molecular Devices) with an excitation wavelength of
365 nm and an emission wavelength of 407 nM. The rate of
polymerization was determined from the slope of the polymer-
ization curve at 50% polymer formation. The concentration of
barbed ends was calculated by setting the rate of polymer for-
mation equal to k
[ends][actin monomer], where k
10
M
1
s
1
and solving for [ends].
Determination of Equilibrium Binding Constants—Binding
constants were determined by fitting the anisotropy curves to
Equation 1,
rrf rb
rf
KdRL兴兲
KdRL兴兲24R兴关L
2L
(Eq. 1)
where rf is the signal of the free receptor (R), rb is the signal of
the bound receptor, and [L] is the total concentration of the
ligand (species titrated). rb and K
d
were fit using KaleidaGraph
(Synergy Software). Determination of binding constants for
unlabeled SpWsp1-VCA measured using competition assay is
described in the supplemental materials.
Crystal Growth and Structure Determination—Crystals of
Arp2 Arp2/3 complex grew in 750 mMammonium sulfate, 50
mMsodium citrate, and 7% glycerol at 4 °C to an average size of
150 80 80
m. ATP and calcium chloride were present in
the crystallization drop at 0.5 mMeach. The crystals diffracted
weakly and were indexed to the P4
2
22 space group with a large
unit cell (a 219.1 Å, b 219.1 Å, c 315.2 Å) and two
molecules in the asymmetric unit (76% solvent). Attempts to
dehydrate crystals by increasing the precipitant concentration
did not improve diffraction. We used a homology model of the
S. pombe Arp2/3 complex based on the crystal structure of
bovine Arp2/3 complex (31) with the Arp2 subunit removed as
a molecular replacement search model. We also removed
inserts and regions with high B-factors in the bovine complex
structure from the search model. We found one solution using
the program Phaser (32) with a translational Z-score of 25.6.
The initial model was improved through three successive
rounds of rebuilding and restrained refinement carried out
using REFMAC with a weighting parameter of 0.005. B-factors
were refined by TLS refinement using one TLS group for each
subunit (33). The two molecules in the asymmetric unit were
symmetrically constrained throughout the refinement. To
improve density for model building, B-factors were sharpened
using a value of 70 Å
2
. The final model had an R
free
of 34.4% and
an R
work
of 32.4%. Coordinates were deposited in the Protein
Data Bank with accession code 3DWL.
Fluorescence Microscopy—Cells were grown to A
595
0.2–
0.5 in YE5S at 25 °C and washed in Edinburgh minimal
medium (potassium phthalate (3 g/liter), Na
2
SO
4
(0.04 g/li-
ter), ZnSO
4
(0.4 mg/liter), Na
2
HPO
4
(2.2 g/liter), panto-
thenic acid (1 mg/liter), FeCl
2
(0.2 mg/liter), NH
4
Cl (5 g/li-
ter), nicotinic acid (10 mg/liter), molybdic acid (40
g/liter),
dextrose (20 g/liter), myo-inositol (10 mg/liter), potassium
iodide (0.1 mg/liter), MgCl
2
(1.05 g/liter), biotin (1 mg/liter),
CuSO
4
(40
g/liter), CaCl
2
(14.7 mg/liter), boric acid (0.5
mg/liter), citric acid (1 mg/liter), KCl (1 g/liter), MnSO
4
(0.4
mg/liter)) before mounting on 25% gelatin pads (34). Images
were acquired on an UltraView RS (PerkinElmer Life Sciences)
spinning disk confocal system installed on an Olympus IX-71
Arp2/3 Complex Crystal Structures
26492 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 283NUMBER 39SEPTEMBER 26, 2008
by guest on November 7, 2015http://www.jbc.org/Downloaded from
microscope with a 100, 1.4
numerical aperture PlanApo lens
(Olympus), and analyzed using
ImageJ software (rsb.info.nih.gov/
ij/). Z-series were collected in
0.6-
m steps with 200-ms sequen-
tial excitation of YFP and CFP.
RESULTS
Isolation of S. pombe Arp2/3
Complex Lacking Arp2—Arp2/3
complex from S. pombe purified
as a single component during
ammonium sulfate precipitation,
ion-exchangechromatography(Q-
Sepharose) and affinity chromatog-
raphy on GST-N-WASp-VCA on a
glutathione-Sepharose 4B column
but split into two peaks during
anion exchange chromatography on
Mono Q. The first Mono Q peak
eluted at a conductivity of 22
mS/cm, and the second peak eluted
at 29 mS/cm (Fig. 1A). We com-
pleted the purification of the two
fractions of Arp2/3 complex by gel
filtration on a column of Superdex
200. Both peaks contained Arp2/3
complex subunits, but the band on
SDS-PAGE previously shown to
contain both Arp2 and ARPC1 (7)
was less intense in pool A (Fig. 1B).
Two-dimensional PAGE (Fig. 1C)
showed that pool A lacked a cluster
of four spots corresponding to Arp2
(theoretical pI 5.8). Immunoblots
with an S. pombe Arp2 antibody ver-
ified that Arp2 was missing (data
not shown). The fraction of com-
plex lacking Arp2 (Arp2 Arp2/3
complex) ranged from 5 to 35%
(average 13%) of total Arp2/3
complex in five preparations from
wild-type cells.
Arp2 Is Required for Arp2/3 Com-
plex Nucleation Activity—Polymer-
ization assays showed that 200 nM
Arp2 Arp2/3 complex did not
nucleate actin filaments, whereas
200 nMof complete Arp2/3 complex
(pool B) increased the concentra-
tion of barbed ends to a maximum
of 3.5 nMin the presence of 1.6
M
SpWsp1-VCA and 4
Mchicken
skeletal muscle actin (Fig. 2A). Even
1
MArp2Arp2/3 complex had no
detectable nucleation activity (data
not shown).
FIGURE 1. Purification of S. pombe Arp2/3 complex lacking Arp2. A, elution of Arp2/3 complex from a Mono
Q column by a gradient of 100 –400 mMNaCl. Absorbance at 280 nm shows two peaks, both containing Arp2/3
complex subunits. B, SDS-PAGE (10–20% gradient of acrylamide) stained with Coomassie Blue of 6.3 pmol of
pools A and B from the Mono Q column further purified by gel filtration on a Superdex 200 column. ARPC1 and
Arp2 are not resolved under these conditions. C, two-dimensional gel electrophoresis of purified pool A and
pool B, stained with Coomassie Blue. Both pools contained a cluster of four spots at the apparent molecular
weight of Arp3, indicating multiple species with distinct isoelectric points (the theoretical pI of Arp3 is 5.8). Pool
B contained a cluster of four additional spots (arrow) just below Arp3, which we presumed to be Arp2 (theo-
retical pI 5.8). This part of the gel shows the bands corresponding to Arp3, Arp2, ARPC1, and ARPC2. Arp2 is
missing from peak A. Theoretical pI values for each subunit are as follows: Arp3, 5.8; Arp2, 5.8; ARPC1, 8.2; and
ARPC2, 6.1.
FIGURE 2. Biochemical characterization of S. pombe Arp2/3 complex with and without Arp2. A, effect of
native and Arp2 Arp2/3 complex on the time course of polymerization of pyrene-labeled Mg-ATP actin.
Conditions: 4
M15% pyrene-labeled chicken skeletal muscle actin, 0.8
MSpWsp1-VCA, 200
Mcomplete
SpArp2/3 complex (“complete”) or Arp2 Arp2/3 complex (“Arp2”) in 10 mMimidazole, pH 7.0, 50 mMKCl, 1
mMMgCl
2
,1mMEGTA, 0.13 mMATP, 63
MCaCl
2
, 0.3 mMDTT, 0.6 mMNaN
3
at 22 °C. Thick black line shows 4
M
actin and 0.8
MSpWsp1-VCA without Arp2/3 complex. Inset shows the concentration of barbed ends when
50% of the actin was polymerized plotted as a function of SpWsp1-VCA concentration for the complete
SpArp2/3 complex (pool B). High concentrations of VCA decrease the rate of polymer formation by inhibiting
nucleation and slowing pointed end elongation (23, 44). B, equilibrium binding of rhodamine-labeled and
unlabeled SpWsp1-VCA to Arp2 Arp2/3 complex and complete Arp2/3 complex measured by fluorescence
anisotropy. Conditions: 50 mMKCl, 10 mMimidazole, pH 7.0, 1 mMMgCl
2
,1mMEGTA, 0.1 mMATP, 1 mMDTT, and
0.2% thesit. Inset: titration of 100 nMSpWsp1-Rho-VCA with Arp2 (dashed line) and native Arp2/3 complex
(solid line). The K
d
values of SpWsp1-Rho-VCA were 120 13 nMfor the Arp2 and 49 5nMfor complete
Arp2/3 complex. Main plot: titration of 100 nMSpWsp1-Rho-VCA and 300 nMArp2 (dashed line) or native
Arp2/3 complex (solid line) with unlabeled SpWsp1-VCA. Curves were fit as described in the methods giving K
d
values of 0.4 0.1
Mand 0.9 0.1
Mfor unlabeled SpWsp1-VCA binding the Arp2 and complete com-
plexes, respectively. C, fluorescence resonance energy transfer to measure binding of SpWsp1-Rho-VCA to
OG-actin. Emission scans showing the dependence of the quenching of the fluorescence of 100 nMOG-actin on
the concentration of SpWsp1-Rho-VCA in the same buffer as in B.Numbers below the curves indicate Rho-VCA
concentrations, in nanomolar. Samples were excited at 480 nm. D, effect of Arp2 Arp2/3 complex and native
Arp2/3 complex on binding of OG-actin to Rho-VCA measured by FRET as in C. Plots of fraction of OG-actin
emission at 517 nm quenched verses Rho-VCA concentration. Fits of the data gave a K
d
of 15.5 1.7 nMfor
Rho-VCA binding to OG-actin (solid line,open squares). In the presence of 3
Mnative Arp2/3 complex (solid line,
filled circles), the K
d
increased to 23.9 1.3 nM. In the presence of 3
MArp2-less complex (dashed line,filled
triangles) the K
d
was 12.1 1.6 nM.
Arp2/3 Complex Crystal Structures
SEPTEMBER 26, 2008VOLUME 283NUMBER 39 JOURNAL OF BIOLOGICAL CHEMISTRY 26493
by guest on November 7, 2015http://www.jbc.org/Downloaded from
The Arp2 Arp2/3 Complex Binds SpWsp1-VCA and Actin
Monomer—To determine the basis for the inactivity of Arp2
Arp2/3 complex, we used fluorescence anisotropy to measure
the affinity of rhodamine-labeled SpWsp1-VCA (Rho-VCA)
for complete and Arp2 Arp2/3 complex. Rho-VCA bound
complete Arp2/3 complex with a dissociation equilibrium con-
stant (K
d
)of495nMand the Arp2 complex with an affinity
of 120 13 nM(Fig. 2B,inset). Because the rhodamine label can
affect the affinity of VCA for Arp2/3 complex (14), we carried
out competition assays using unlabeled SpWsp1-VCA (VCA)
(Fig. 2B). VCA had a slightly higher affinity (K
d
0.4 0.1
M)
for Arp2 Arp2/3 complex than native Arp2/3 complex (K
d
0.9 0.1
M). Therefore, the Arp2-less complex is not inactive
due to a loss of ability to bind this nucleation-promoting factor.
We next sought to determine if VCA bound to Arp2 Arp2/3
complex could recruit an actin monomer to form the ternary
complex of Arp2/3 complex, NPF, and an actin monomer. This
ternary complex is thought to assemble before binding to an
actin filament to initiate a branch (14). To detect interaction of
the VCA NPF with actin, we used fluorescence energy reso-
nance transfer (FRET) from an Oregon Green 488 label on Cys-
374 of actin (OG-actin) to Rho-VCA (Fig. 2C) (35). This assay
gave a K
d
of 16 2nMfor Rho-VCA binding OG-actin in the
absence of Arp2/3 complex (Fig. 2D). FRET was then measured
in the presence of 3
MArp2/3 complex, so that 98% of the
Rho-VCA was bound to Arp2/3 complex. The presence of 3
M
complete Arp2/3 complex slightly increased the K
d
of Rho-
VCA and OG-actin to 24 1n
M, whereas the presence of 3
M
Arp2Arp2/3 complex had no effect on the binding (K
d
12
2n
M). These results indicate that both complete and Arp2
Arp2/3 complexes can form a ternary complex with a VCA
nucleation-promoting factor and an actin monomer. Because
excess Arp2/3 complex does not increase the affinity of actin for
Rho-VCA, we conclude that actin does not make productive
contacts with Arp2/3 complex in the ternary complex.
Structure of Arp2 Arp2/3 Complex—To determine whether
loss of the Arp2 subunit perturbs the structure of Arp2/3 com-
plex, we solved the crystal structure of Arp2Arp2/3 complex
using molecular replacement with a homology model based on
the crystal structure of bovine Arp2/3 complex (31) to estimate
initial phases. We used this model to generate a solvent-flat-
tened electron density map averaged using the two molecules in
the asymmetric unit. The map showed density for a number of
features not included in the original model, including a section
of 33 residues of random coil inserted into ARPC1 and ATP in
the Arp3 cleft (see “Discussion”). Three successive rounds of
rebuilding and restrained refinement improved the model. Of
the 1615 total residues per complex, 1308 were modeled, but
287 of these were lacking density for side chains and were mod-
eled as alanine. Much of the backbone of ARPC3 was disor-
dered, so only 65% of this subunit could be modeled. The final
R
free
was 34.4%, and the final R
work
was 32.4% (Table 1), high
R-factors typical of this resolution range. Despite the low reso-
lution, omit electron density maps calculated at the start of the
refinement clearly showed not only the secondary structure,
but side-chain density for most residues (supplemental Fig. S1).
This is the first crystal structure of Arp2/3 complex other than
that from cow (10 –12). Given the limited resolution of the data,
we confine our discussion to comparisons of gross structural
features of S. pombe and bovine complexes, such as the overall
arrangement of the subunits and the conformation of long
backbone segments that adopt dramatically different confor-
mations in the structures.
The overall structure of the Arp2 Arp2/3 complex from S.
pombe closely resembles bovine Arp2/3 complex, except for the
absence of Arp2 (Fig. 3A). Arp2/3 complexes from these two
species can be overlaid with an overall root mean square devi-
ation of 1.85 Å for 1166 aligned C
atoms. Therefore, dissoci-
ation of Arp2 does not cause major changes in the rest of the
complex (see supplemental Fig. S2 for detailed analysis).
The complete S. pombe Arp2/3 complex did not crystallize
under the same conditions used to grow crystals of Arp2
Arp2/3 complex. Modeling Arp2 into the Arp2 Arp2/3 com-
plex crystal revealed that Arp2 sterically clashes with the Arp3
subunit from a symmetry-related complex, explaining why the
packing arrangement is not possible if Arp2 is present.
The presence of the ARPC1 insert was the most striking fea-
ture of the electron density map of S. pombe Arp2 Arp2/3
complex (Fig. 3, Aand B, and 4). Electron density for this region
was present in the first F
o
F
c
map and subsequent rounds of
rebuilding/refinement allowed us to build 19 of 43 residues of
the insert. This region forms a random coil that inserts into the
groove between subdomains 2 and 4 of Arp3 from the other
molecule in the asymmetric unit (supplemental Fig. S3A). In
some crystals of bovine Arp2/3 complex, the most conserved
portion of the ARPC1 insert forms an
-helix, which packs
against the barbed end of Arp3 between subdomains 1 and 3 in
a symmetry-related molecule (supplemental Fig. S3B) (10). The
conformations of the ARPC1 inserts differ markedly in crystals
of S. pombe and bovine Arp2/3 complexes (Fig. 3A). The
ordered region of the insert in the structure of Arp2 Arp2/3
complex consists of residues unique to S. pombe (Fig. 3C),
whereas the conserved region of the insert that forms an
-helix
in some bovine Arp2/3 complex structures is disordered in the
TABLE 1
Data collection and refinement statistics
Data collection statistics
Resolution limits (Å) 29.0-3.80
Space group P4
2
22
Cell constants a b218.98
c315.05
90.0
Mosaicity (°) 0.46
Measured reflections 1,179,649
Unique reflections 72,502
Mean I/
9.5 (2.6)
R
sym
(%) 19.8 (49.5)
Completeness (%) 99.4 (95.2)
Refinement statistics
Modeled atoms 18,717
R
free
reflections 3,847 (5%)
Average B-factor
2
)74.7
Root mean square from ideal
Bond lengths (Å) 0.007
Bond angles (°) 1.042
Ramachandran statistics
Most favored 1,751 (76.2%)
Additionally allowed 492 (21.4%)
Generously allowed 54 (2.3%)
Disallowed 0
R
free
(%) 34.4
R
work
(%) 32.4
Arp2/3 Complex Crystal Structures
26494 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 283NUMBER 39SEPTEMBER 26, 2008
by guest on November 7, 2015http://www.jbc.org/Downloaded from
structure of Arp2 Arp2/3 complex (Fig. 3C). The conserved
portion of the ARPC1 insert is proposed to interact with one or
more actin subunits in the mother filament (9, 10). The flexibil-
ity revealed by our new structure may be critical for this pro-
posed function of this part of Arp2/3 complex.
The N Terminus of ARPC5 Is Not Essential for Activity—A
model of Arp2/3 complex in an actin filament branch based on
a three-dimensional reconstruction from electron tomograms
has Arp2 positioned next to Arp3 as the second subunit in the
daughter filament (9). This requires a shift of Arp2 by 31 Å from
its position in the inactive conformations captured in crystals.
The “migration model” for activation (22) proposed that the N
terminus of ARPC5 serves as a tether for Arp2 as it dissociates
from Arp3, ARPC1, and ARPC4 and migrates into the active
conformation (Fig. 5, Aand B).
We tested this hypothesis by
mutating the conserved arginine
from the N terminus of ARPC5 to a
glutamic acid (ARPC5R3E) or delet-
ing the N terminus entirely
(ARPC5N). Both mutant strains
grew normally at 30 °C on YES
plates but slower than wild type at
25 °C and 36 °C on YES 1MNaCl
(data not shown). Additionally, both
mutants were defective in expelling
Phloxin-B at 25 °C and 36 °C, but
not at 30 °C (data not shown). The
ARPC5N strain yielded 50%
more purified complex lacking
Arp2 than wild-type cells (data not
shown). Purified ARPC5N Arp2/3
complex (Fig. 5C) nucleated fila-
ments as efficiently as native
SpArp2/3 complex (Fig. 5D). Thus,
the N terminus of ARPC5 is not
necessary for branching nucle-
ation in vitro.
Arp2 Co-localizes with Other
Arp2/3 Complex Subunits through-
out the Cell Cycle—To determine if
Arp2/3 complex in some parts of
cells lacks Arp2, we imaged haploid
strains of S. pombe expressing both Arp2 C-terminally labeled
with cyan fluorescent protein (CFP) and ARPC3 C-terminally
labeled with yellow fluorescent protein (YFP). Both tagged
genes were expressed from native promoters and provided the
sole copy of the gene. Cells depending on both tagged proteins
were viable but often misshapen. Their actin patches turned
over slower than patches in cells with only one tagged Arp.
3
Arp2 and ARPC3 co-localized in actin patches during all stages
of the cell cycle (Fig. 6A). The relative fluorescence intensities of
Arp2-CFP and ARPC3-YFP in patches were well correlated
(Fig. 6B), showing that subpopulations of patches depleted of
Arp2 were not present. The ratios of Arp2-CFP and ARPC3-
YFP fluorescence were unequal in some patches moving away
from the plasma membrane, owing to the time lapse between
acquisition of CFP and YFP images. These results suggest that,
if Arp2 dissociates from the rest of Arp2/3 complex in S. pombe,
it does not leave the actin patches.
DISCUSSION
Isolation and Activity of the Arp2 Arp2/3 Complex—We
considered multiple hypotheses to explain why some of puri-
fied fission yeast Arp2/3 complex lacks Arp2. Measurements
of fluorescently tagged Arp2/3 complex subunits in fission
yeast indicated that all subunits were present in the cyto-
plasm at near equal concentrations, with Arp2 the second
most abundantly expressed subunit (36). Therefore, low lev-
els of Arp2 expression cannot explain the existence of the
Arp2 Arp2/3 complex.
3
V. Sirotkin, personal communication.
FIGURE 3. Crystal structure of SpArp2/3 complex lacking Arp2. A,C
trace showing overlay of the Arp2
Arp2/3 complex (orange, Arp3; green, ARPC1; cyan, ARPC2; magenta, ARPC3; blue, ARPC4; yellow, ARPC5) onto
the bovine apo-protein complex (1K8K.pdb, gray C
trace). The overall arrangement of the subunits is identical
in both Arp2/3 complexes with 1166 atoms overlaid with an overall root mean square deviation of 1.9 Å
2
. The
left panel shows the standard orientation of the complex, the right panel shows the complex rotated by 90°
about the vertical axis. All but the Arp2 and ARPC1 subunits of the bovine complex are omitted from the right
panel for clarity. B,ribbon diagrams of overlaid S. pombe (green) and bovine (blue) ARPC1. The seven blades of
the propeller are numbered 1–7 and the
-strands from one propeller are labeled A–D. The ARPC1 insert is
located between
D6 and
A7. C, sequence alignment of the ARPC1 insert from six diverse species. Abbrevi-
ations are as follows: Sp,S. pombe;Sc,S. cerevisiae;Bt,Bos taurus;Ce,Caenorhabditis elegans;Dd,Dictyostelium
discoideum; and Dm,Drosophila melanogaster. Secondary structure for S. pombe (green) and bovine complexes
(blue) is indicated above and below the alignment, respectively. Dashed lines indicate regions of disorder in the
structures.
FIGURE 4. Stereo figure of electron density of the ARPC1 insert. 2F
o
F
c
electron density map contoured at 2.0
calculated with phase contributions
for the insert region from ARPC1 (subunits C and H, residues 291–311) omit-
ted. C
trace of ARPC1 is shown in blue.
Arp2/3 Complex Crystal Structures
SEPTEMBER 26, 2008VOLUME 283NUMBER 39 JOURNAL OF BIOLOGICAL CHEMISTRY 26495
by guest on November 7, 2015http://www.jbc.org/Downloaded from
Dissociation of Arp2 during purification is the most likely
explanation for the presence of Arp2 Arp2/3 complex. No
information is available on the affinity of individual subunits for
Arp2/3 complex in S. pombe, but no Arp2 dissociated from the
complete fission yeast complex during gel-filtration chroma-
tography, indicating that the complete complex is stable under
these conditions. A similar experiment with Acanthamoeba
Arp2/3 complex showed that the highest dissociation equilib-
rium constant for any subunit is 70 nM(37). However, 7% of
Arp2 dissociated when the complete complex was subjected to
a second round of Mono Q purification. We conclude that Arp2
is the least tightly associated subunit in the S. pombe complex
and that some Arp2 dissociates during anion-exchange chro-
matography and elution with high salt. Electrospray mass spec-
trometry after high-pressure liquid chromatography showed
no mass differences for the individual subunits of the complete
and Arp2complexes.
4
Therefore, neither proteolysis during
purification nor differences in post-translational modifications
to subunits common to both complexes is responsible for dis-
sociation of Arp2. We attempted to purify His-tagged Arp2
from the rest of the complex for
reconstitution experiments but
could not isolate adequate amounts
of purified Arp2 under either native
or denaturing conditions (data not
shown).
Previous work suggested that
Arp2 may not be essential in some
organisms, but no one had studied
the biochemical properties of
Arp2/3 complex lacking Arp2. Win-
ter et al. (38) found that deletion of
ARP2 from Saccharomyces cerevi-
siae is not lethal, suggesting that
actin might substitute for Arp2 dur-
ing filament branching. However,
the E316K mutation of S. pombe
Arp2 causes dissociation of Arp2
from the complex and septation
defects at 36 °C indicating the
importance of Arp2 (39). Gournier
et al. (21) expressed human Arp2/3
complex without the Arp2 subunit
in insect cells, but could not isolate
the Arp2Arp2/3 complex.
Our preparation of fission yeast
Arp2/3 complex lacking Arp2 did
not nucleate actin filaments in an
assay with SpWsp1-VCA and
chicken skeletal muscle actin (Fig.
2A), so actin cannot substitute for
Arp2 in the branching reaction. S.
pombe Arp2 and chicken skeletal
muscle actin are 45% identical, but
many residues of Arp2 that contact
ARPC4, ARPC1, or ARPC5 are different in actin. For example,
the
G helix in Arp2 (residues 226–236) and the loop immedi-
ately following it (residues 237–242) make extensive contacts to
ARPC4, and only four residues in this region are identical in
actin and Arp2. In addition, both Arp2 and Arp3 contain an
insert in the
K/
15 (Arp2 residues 320 –334) loop not present
in actin. In Arp2, this insert forms a major part of the interac-
tion surface with the N terminus of ARPC5. In the model based
on the reconstruction of branch junctions (9), the DNase-bind-
ing loop of Arp2 (residues 39–51) interacts with an actin sub-
unit in the mother filament. The DNase-binding loop of Arp2
differs in sequence and is two residues longer than in actin.
These differences provide a convincing structural basis for the
failure of actin to substitute for Arp2 in the nucleation reaction.
The presence of Arp2/3 complex lacking Arp2 in extracts of
S. pombe led us to wonder if dissociation of Arp2 plays a role in
regulating Arp2/3 complex activity in vivo. The inactivity of
Arp2 Arp2/3 complex is consistent with the observation that
the E316K mutation in Arp2 causes defects in cell septation, a
process thought to require Arp2/3 complex (39, 40). Fluores-
cence microscopy of wild-type fission yeast strains expressing
Arp2-CFP and ARPC3-YFP showed that the ratio of Arp2 to
ARPC3 was uniform in actin patches throughout the cell cycle
4
S. Almo and W. Zenchek, Albert Einstein College of Medicine, personal
communication.
FIGURE 5. The N-terminal residues of ARPC5 are not required for nucleation activity of purified S. pombe
Arp2/3 complex. A, ribbon diagram of bovine Arp2/3 complex (1K8K) showing the “back side” of the complex
relative to the standard orientation in Fig. 3A. The N terminus of ARPC5 forms a random coil that wraps around
the back of the complex and binds to a groove between subdomains 3 and 4 of Arp2. Disordered residues of
ARPC5 (31–34) are indicated with a dashed yellow line.B, detail of the boxed region in Awith key side chains
shown as sticks. Arp2 is pink, and ARPC5 is yellow.C, SDS-PAGE of purified wild-type (left lane) and ARPC5N
complexes stained with Coomassie Blue. Fourteen residues are deleted from the N terminus of ARPC5N
Arp2/3 complex (see supplemental Fig. S4). The mutated ARPC5 subunit has a higher mobility than the native
subunit (arrow). D, comparison of the effects of native and ARPC5N Arp2/3 complex on the time course of the
polymerization of pyrene-labeled actin. Conditions: same as those in Fig. 2Awith 80 nMwild-type (red)or
ARPC5N(green) Arp2/3 complex, 1.0
MSpWsp1-VCA, 4
M15% pyrene-labeled Mg-ATP actin. The black line
shows actin polymerization in the absence of Arp2/3 complex and activator.
Arp2/3 Complex Crystal Structures
26496 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 283NUMBER 39SEPTEMBER 26, 2008
by guest on November 7, 2015http://www.jbc.org/Downloaded from
(Fig. 6). Previous studies established that all actin patches con-
tain a 1:1:1:1:1 ratio of all five subunits measured (Arp3, Arp2,
ARPC1, ARPC3, and ARPC5) (36).
3
Although we cannot rule
out the possibility that Arp2 dissociates from the rest of the
complex but remains in the patches, this scenario seems
unlikely, because the E316K mutation, which favors dissocia-
tion of Arp2 from the complex in vivo, causes Arp2 to leave the
actin patches and adopt a diffuse localization (39). Therefore,
we have no evidence that cells use Arp2 Arp2/3 complex
selectively in patches or other structures.
Interaction of VCA and Actin Monomers with Arp2/3 Com-
plex Lacking Arp2—We found that the loss of Arp2 does not
decrease the affinity of SpWsp1-Rho-VCA for SpArp2/3 com-
plex. Similarly, GST-Bee1-VCA (a budding yeast WASp hom-
olog construct) pulled down similar amounts of Arp3 from
extracts of wild-type and arp2 strains of budding yeast (41).
This is consistent with experiments that showed that most of
the binding energy of GST-Bee1-VCA for ScArp2/3 complex
comes from its interaction with the ARPC1 (Arc40) subunit
(41). On the other hand, VCA constructs protect Arp2 of
bovine Arp2/3 complex from oxidation in synchrotron radia-
tion footprinting experiments (16) and have been chemically
cross-linked to the Arp2 subunit of Arp2/3 complex from mul-
tiple species (17–19). All of these observations may be recon-
ciled, if VCA contacts Arp2 without contributing strongly to
the binding affinity.
Our FRET assays confirmed that the affinity of OG-actin for
rhodamine-VCA is slightly weaker in the presence of the full
Arp2/3 complex (K
d
16 2nMwithout Arp2/3 complex,
24 1n
Mwith Arp2/3 complex) (35). This indicates that VCA,
actin, and Arp2/3 complex can interact simultaneously, but
that the actin monomer in this ternary complex does not make
productive contacts with Arp2/3 complex. We suggest that
actin in the ternary complex must reorient relative to Arp2/3
complex subunits during the activation step (35) to establish a
nucleus for the polymerization of the daughter filament. The
lower affinity of actin for VCA in the presence of Arp2/3 com-
plex also suggests that actin and Arp2/3 complex compete for a
common binding site on VCA. The C region is the best candi-
date for this common site (14), because it binds weakly to both
actin (K
d
12
M) and Arp2/3 complex (K
d
200
M)ina
mutually exclusive manner (42).
The FRET assay showed that the affinity of VCA for OG-
actin is 2-fold stronger (K
d
12 2nM) with Arp2 Arp2/3
than full Arp2/3 complex (K
d
24 1nM). We speculate that
dissociation of Arp2 either relieves steric inhibition that pre-
vents the C region from interacting with actin and/or that Arp2
subunit completes with actin directly by weakly binding the C
region.
We note that our results differ from the interpretation of a
small angle x-ray scattering model of Arp2/3 with bound acti-
vator and monomeric actin (20), in which the C-region of the
activator makes extensive interactions with Arp2. Although we
do not rule out weak interactions between Arp2 and the activa-
tor, our data show that the loss of Arp2 increases the affinity of
VCA for Arp2/3 complex and does not affect recruitment of
actin to VCA bound to Arp2/3 complex.
Insights into the Mechanism of Arp2/3 Complex Activation
Our structure of the Arp2 Arp2/3 complex shows that disso-
ciation of the Arp2 subunit did not perturb the remaining sub-
units in the complex. This establishes the feasibility that Arp2
partially dissociates during the conformational change that
activates the complex (9). However, the N terminus of ARPC5,
which has been proposed to tether Arp2 to the complex during
the proposed conformational rearrangement (22), is not neces-
sary for Arp2/3 complex activity in vitro, suggesting that Arp2
maintains contacts with other subunits during activation. This
observation supports a model where a twisting motion rotates a
rigid body composed of Arp2, ARPC1, ARPC4, and ARPC5 into
the active conformation (10). The interpretation of small angle
x-ray scattering from Arp2/3 complex with bound activator
and actin monomer also supports the rotation model (20).
Although the N-terminal tether of ARPC5 does not play a
fundamental role in the activation of Arp2/3 complex, it may be
important in stabilizing interactions of Arp2 with the rest of the
complex. Consistent with this hypothesis, we recovered more
Arp2 Arp2/3 complex from the ARPC5N strain than wild-
type cells. The residues involved in the interface between
ARPC5 and Arp2 are conserved in most species (supplemental
Fig. S4), but in most plant species the N terminus of ARPC5 is
FIGURE 6. Arp2-CFP and ARPC3-YFP co-localize to actin patches. A, spin-
ning disk confocal fluorescence micrographs of a haploid strain of S. pombe
with Arp2 tagged on its C terminus with CFP and ARPC3 tagged on its C
terminus with YFP, both expressed from their native promoters. Maximum
projection images created from fourteen 0.6
m Z-sections. Left panel, Arp2-
CFP intensity; center panel, ARPC3-YFP intensity; and right panel, merged
images with Arp2-CFP intensity shown in red and ARPC3-YFP intensity shown
in green.B, correlation of the intensity of CFP and YFP fluorescence in individ-
ual actin patches. Plot of mean CFP and YFP intensity for 163 patches in 3 cells.
The linear correlation coefficient 0.81.
Arp2/3 Complex Crystal Structures
SEPTEMBER 26, 2008VOLUME 283NUMBER 39 JOURNAL OF BIOLOGICAL CHEMISTRY 26497
by guest on November 7, 2015http://www.jbc.org/Downloaded from
too short to reach Arp2 in models based on structures of inac-
tive bovine Arp2/3 complex (10 –12) or the model of the branch
junction based on electron tomography (9). Therefore, the sta-
bilizing function of the N terminus of ARPC5 is unlikely to
occur in plants.
We were surprised that Arp2 does not contribute to the affin-
ity of VCA for Arp2/3 complex and that, in fact, the affinity is
slightly higher without Arp2. This observation suggests that
VCA facilitates the movement of Arp2 next to Arp3 without
strongly interacting with Arp2 Perhaps Arp2 is a passive partic-
ipant in activation, with interactions of other subunits of
Arp2/3 complex with NPFs and a mother filament providing
most of the free energy for the conformational change that cre-
ates a favorable binding site for the first actin subunit of the
daughter filament with Arp2 and Arp3. Alternatively, VCA may
interact strongly with Arp2 only after the complex is bound to
mother filament. Consistent with this hypothesis, kinetic and
thermodynamic data demonstrated multiple modes of NPF
binding to Arp2/3 complex (42), and sequence similarities
between C and V regions suggest that the C region may interact
with Arp2 just as V interacts with actin during activation (43).
Elucidation of this complex activation mechanism will require
much more structural, biophysical, and biochemical informa-
tion than is currently available.
Acknowledgments—We thank Kathleen Gould for the S. pombe Arp2
antibody, Chris Beltzner for yeast strains with Arp2 and Arp3 tagged
with CFP and YFP, Vladimir Sirotkin for help with microscopy, Aaron
Downs for help with 2D gels, Yong Xiong for advice on analyzing low
resolution data, Hongli Chen for assistance with protein preparations,
Aditya Paul for comments on the manuscript, and Shih-Chieh Ti and
Julien Berro for deriving the formula for fitting the competition bind-
ing curve.
REFERENCES
1. Pollard, T. D., Blanchoin, L., and Mullins, R. D. (2000) Annu. Rev. Biophys.
Biomol. Struct. 29, 545–576
2. Pollard, T. D. (2007) Annu. Rev. Biophys. Biomol. Struct. 36, 451– 477
3. Chang, F., Drubin, D., and Nurse, P. (1997) J. Cell Biol. 137, 169 –182
4. Feierbach, B., and Chang, F. (2001) Curr. Biol. 11, 1656 –1665
5. McCollum, D., Feoktistova, A., Morphew, M., Balasubramanian, M., and
Gould, K. L. (1996) EMBO J. 15, 64386446
6. Toshima, J. Y., Toshima, J., Kaksonen, M., Martin, A. C., King, D. S., and
Drubin, D. G. (2006) Proc. Natl. Acad. Sci. U. S. A. 103, 5793–5798
7. Sirotkin, V., Beltzner, C. C., Marchand, J. B., and Pollard, T. D. (2005)
J. Cell Biol. 170, 637–648
8. Amann, K. J., and Pollard, T. D. (2001) Nat. Cell Biol. 3, 306 –310
9. Rouiller, I., Xu, X. P., Amann, K. J., Egile, C., Nickell, S., Nicastro, D., Li, R.,
Pollard, T. D., Volkmann, N., and Hanein, D. (2008) J. Cell Biol. 180,
887–895
10. Robinson, R. C., Turbedsky, K., Kaiser, D. A., Marchand, J. B., Higgs, H. N.,
Choe, S., and Pollard, T. D. (2001) Science 294, 1679–1684
11. Nolen, B. J., Littlefield, R. S., and Pollard, T. D. (2004) Proc. Natl. Acad. Sci.
U. S. A. 101, 15627–15632
12. Nolen, B. J., and Pollard, T. D. (2007) Mol. Cell 26, 449 457
13. Chereau, D., Kerff, F., Graceffa, P., Grabarek, Z., Langsetmo, K., and
Dominguez, R. (2005) Proc. Natl. Acad. Sci. U. S. A. 102, 16644–16649
14. Marchand, J. B., Kaiser, D. A., Pollard, T. D., and Higgs, H. N. (2001) Nat.
Cell Biol. 3, 76–82
15. Panchal, S. C., Kaiser, D. A., Torres, E., Pollard, T. D., and Rosen, M. K.
(2003) Nat. Struct. Biol. 10, 591–598
16. Kiselar, J. G., Mahaffy, R., Pollard, T. D., Almo, S. C., and Chance, M. R.
(2007) Proc. Natl. Acad. Sci. U. S. A. 104, 1552–1557
17. Kreishman-Deitrick, M., Goley, E. D., Burdine, L., Denison, C., Egile, C., Li,
R., Murali, N., Kodadek, T. J., Welch, M. D., and Rosen, M. K. (2005)
Biochemistry 44, 15247–15256
18. Zalevsky, J., Grigorova, I., and Mullins, R. D. (2001) J. Biol. Chem. 276,
3468–3475
19. Zalevsky, J., Lempert, L., Kranitz, H., and Mullins, R. D. (2001) Curr. Biol.
11, 1903–1913
20. Boczkowska, M., Rebowski, G., Petoukhov, M. V., Hayes, D. B., Svergun,
D. I., and Dominguez, R. (2008) Structure 16, 695–704
21. Gournier, H., Goley, E. D., Niederstrasser, H., Trinh, T., and Welch, M. D.
(2001) Mol. Cell 8, 1041–1052
22. Irobi, E., Aguda, A. H., Larsson, M., Guerin, C., Yin, H. L., Burtnick, L. D.,
Blanchoin, L., and Robinson, R. C. (2004) EMBO J. 23, 3599–3608
23. Higgs, H. N., Blanchoin, L., and Pollard, T. D. (1999) Biochemistry 38,
15212–15222
24. Gill, S. C., and von Hippel, P. H. (1989) Anal. Biochem. 182, 319 –326
25. Kapust, R. B., Tozser, J., Fox, J. D., Anderson, D. E., Cherry, S., Copeland,
T. D., and Waugh, D. S. (2001) Protein Eng. 14, 993–1000
26. MacLean-Fletcher, S., and Pollard, T. D. (1980) Biochem. Biophys. Res.
Commun. 96, 18–27
27. Pollard, T. D. (1984) J. Cell Biol. 99, 769 –777
28. Kuhn, J. R., and Pollard, T. D. (2005) Biophys. J. 88, 1387–1402
29. Bahler, J., Wu, J. Q., Longtine, M. S., Shah, N. G., McKenzie, A., 3rd,
Steever, A. B., Wach, A., Philippsen, P., and Pringle, J. R. (1998) Yeast 14,
943–951
30. Keeney, J. B., and Boeke, J. D. (1994) Genetics 136, 849 856
31. Beltzner, C. C., and Pollard, T. D. (2004) J. Mol. Biol. 336, 551–565
32. McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, L. C., Storoni,
L. C., and Read, R. J. (2007) J. Appl. Crystallogr. 40, 658674
33. Winn, M. D., Isupov, M. N., and Murshudov, G. N. (2001) Acta Crystal-
logr. D. Biol. Crystallogr. 57, 122–133
34. Wu, J. Q., Kuhn, J. R., Kovar, D. R., and Pollard, T. D. (2003) Dev. Cell 5,
723–734
35. Beltzner, C. C., and Pollard, T. D. (2007) J. Biol. Chem. 283, 7135–7144
36. Wu, J. Q., and Pollard, T. D. (2005) Science 310, 310 –314
37. Mullins, R. D., Stafford, W. F., and Pollard, T. D. (1997) J. Cell Biol. 136,
331–343
38. Winter, D. C., Choe, E. Y., and Li, R. (1999) Proc. Natl. Acad. Sci. U. S. A.
96, 7288–7293
39. Morrell, J. L., Morphew, M., and Gould, K. L. (1999) Mol. Biol. Cell 10,
4201–4215
40. Welch, M. D., Holtzman, D. A., and Drubin, D. G. (1994) Curr. Opin. Cell
Biol. 6, 110–119
41. Pan, F., Egile, C., Lipkin, T., and Li, R. (2004) J. Biol. Chem. 279,
54629–54636
42. Kelly, A. E., Kranitz, H., Dotsch, V., and Mullins, R. D. (2006) J. Biol. Chem.
281, 10589–10597
43. Aguda, A. H., Burtnick, L. D., and Robinson, R. C. (2005) EMBO Rep. 6,
220–226
44. Boujemaa-Paterski, R., Gouin, E., Hansen, G., Samarin, S., Le Clainche, C.,
Didry, D., Dehoux, P., Cossart, P., Kocks, C., Carlier, M. F., and Pantaloni,
D. (2001) Biochemistry 40, 11390–11404
Arp2/3 Complex Crystal Structures
26498 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 283NUMBER 39SEPTEMBER 26, 2008
by guest on November 7, 2015http://www.jbc.org/Downloaded from
Supplemental Experimental Procedures
Determination of binding affinity of unlabelled VCA in competition assay. The following
equations were used to determine the binding affinity of VCA:
TB
AB
rfrbrfr ][
][
)( +=
)][][][(2
][][][2][][2][2][][][2
][
][
1
2
11212
22
1
2
1
22
1
2
12
2
12
2
1
2
211212
TTT
TTTTTTTTT
TAKKXKAKKK
AKAXKXKAKKXKKKKAKXKAKKK
B
AB
++
++++++
=
where r is the anisotropy of Rho-VCA, rf is the anisotropy of free Rho-VCA, rb is the
anisotropy of Rho-VCA bound to Arp2/3 complex, [AB] is the concentration of bound
Rho-VCA, [B]T is the total concentration of Rho-VCA, K1 is the binding affinity of Rho-
VCA for Arp2/3 complex, [X]T is the total concentration of VCA, [A]T is the total
concentration of Arp2/3 complex and K2 is the affinity of VCA for Arp2/3 complex.
SUPPLEMENTAL FIGURE S1. 2Fo-Fc omit electron density of ARPC4. An omit electron
density map was generated by removing subunit F (ARPC4) from the molecular replacement
solution and improving the structure with one round of rigid body refinement and simulated
annealing with NCS restraints. The resulting pdb file was used to generate an omit map in which
none of the atoms in subunit F contribute to the calculated phases. The map was improved
through NCS averaging and is shown here contoured at 2 σ.
SUPPLEMENTAL FIGURE S2. Rfree values are sensitive to small changes in the orientation
of individual subunits in the ΔArp2 Arp2/3 complex structure. One of the twelve subunits
(Arp3, subunit A) in the asymmetric unit (this subunit includes 2557 of the 18726 atoms in the
asymmetric unit) was rotated by small angle increments and the Rfree value of the resulting
structure was calculated. The axis of rotation goes through the center of mass of the subunit, so
displacements of the atoms in the subunit are small even at relatively large rotation angles. Figure
B shows that rotation by as little as 0.5 degrees causes a detectable increase in the Rfree. Rotation
of 2.5 degrees causes an increase in Rfree of over 2%.
SUPPLEMENTAL FIGURE S3. Interaction of the ARPC1 insert with Arp3 from
symmetryrelated molecules in bovine and fission yeast crystal structures. The insert in
ARPC1 (green) interacts with symmetry-related Arp3 (orange) molecules in both the bovine and
fission yeast structures.
A, the ARPC1 insert (276-311,337-343) forms a random coil that interacts with subdomains 1 and
2 of Arp3 from the second molecule in the asymmetric unit in fission yeast structure.
B, residues (297-309) of the bovine ARPC1 insert form an α-helix which binds to the
groove between subdomains 1 and 3 in a symmetry-related Arp3 molecule (1K8K). In the right
panel the structures are rotated ~90° about the y-axis relative to the left panel. Arrows point to the
ARPC1 insert.
SUPPLEMENTAL FIGURE S4. Alignment of N-terminal sequences of ARPC5 from diverse
species. Species abbreviations are as follows: Bt: B. taurus (gi17943205), Hs: H. sapiens
(NP_005708), Ec: E. caballus (XP_00149039), Ce: C. elegans (NP_491099), Dr: D. rerio
(AAL55526), Xl: X. laevis (NP_001086165), Dm: D. melanogaster (NP_608693), Sp: S. pombe
(NP_593727), Sc: S. cerevisiae (NP_012202), Af: A. funestus (ABI83786), An: A. oryzae
(BAE57777), Vv: V. vinifera (CAN73657), At: A. thaliana (CAB77741), Os: Oryza sativa
(NP_001068095). Alignment was created using the Dialign server (45). Secondary structure from
1K8K.pdb is blue. Lines represent random coil, dashes are disordered regions, and boxes indicate
α-helices. Conserved residues that interact with Arp2 in 1K8K are boxed (hydrophobic residues)
or colored red (charged residues).
Figure S1
Arp3 r otat ion: effect on Rfree
30.0
34.0
38.0
42.0
46.0
-30.0-20.0-10.00.010.020.030.0
angle
Rfree (%
)
Figure S2
~90°
AB
1
1
2
2
3
3
4
4
2
1
4
Figure S3
Bt MSKNTVSSARFRKVDVDEY-------------------DENKFVDEDDGGDGQ-A--------GPDEGEVDSCLRQ-------GNMTAALQAALKNPPIN
Hs MSKNTVSSARFRKVDVDEY-------------------DENKFVDEEDGGDGQ-A--------GPDEGEVDSCLRQ-------GNMTAALQAALKNPPIN
Ec MSKNTVSSARFRKVDVDEY-------------------DENKFVDEEDGGDGQ-A--------GPDEGEVDSRLRQ-------GNMMAALQAALKNPPIN
Ce M-SKNMQNTSYRKLDVDSF-------------------DPEQYDENDETVDTPGL--------GPDERAVQGFLSS-------NRLEDALHAALLSPPLK
Dr M-SKNTVSDRFRKVDVDEY-------------------DENKFVDEEDGGENQ-L--------GPDEAEVDSLIRS-------GNLMGALQAVLKNPPIH
Xl M-AKNTLSSRFRKVDIDEY-------------------DENKFVDDQLQEEPVEP-------QGPDEAEVDSLIRQ-------GDLLRAFQSALINSPVN
Dm M-AKNTSSNAFRKIDVDQY-------------------NEDNFREDDG-VESAAA--------GPDESEITTLLTQ-------GKSVEALLSALQNAPLR
Sp --------MTFRTLDVDSI-------------------TEPVLTEQDIFPIRNET-------AEQVQAAVSQLIPQARSAIQTGNALQGLKTLLSYVPYG
Sc ------MEADWRRIDIDAF-------------------DPESGRLTAADLVPPYETTVTLQELQPRMNQLRSLATS-------GDSLGAVQLLTTDPPYS
Af -MAKNTSSSAFRKIDVDQY-------------------NEDNFKEDDADQASSGM-------IVPDEAEINSLLNQ-------GRNIDALKTVLQNAPLM
An -----MAQINYRTINIDVLDPESSVNFPMETLLPPTLPAPTT--------SSE-A--------ANVAAQVRQLLRS-------GDPEGALRAVLDTAPLG
Vv ------MAGTKEFVEADN--------------------AEAI--------ITR-I--------EHKSRKIESLLKQ-------HKPIEALKTALEGSPPN
At ---------MAEFVEADN--------------------AEAI--------IAR-I--------ETKSRKIESLLKQ-------YKHVEALKTALEGSPPK
Os -----MASSAAAYLDADEN-------------------LEAI--------ISR-I--------EQKSRKIETLLKQ-------SKPVEALKTALEGTPLK
Figure S4
Brad J. Nolen and Thomas D. Pollard
Arp2 Subunit
Fission Yeast Arp2/3 Complex Lacking the
Structure and Biochemical Properties of
Developmental Biology:
Molecular Basis of Cell and
doi: 10.1074/jbc.M802607200 originally published online July 18, 2008
2008, 283:26490-26498.J. Biol. Chem.
10.1074/jbc.M802607200Access the most updated version of this article at doi:
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
Supplemental material:
http://www.jbc.org/content/suppl/2008/07/23/M802607200.DC1.html
http://www.jbc.org/content/283/39/26490.full.html#ref-list-1
This article cites 44 references, 19 of which can be accessed free at
by guest on November 7, 2015http://www.jbc.org/Downloaded from
... For example, in vivo nucleotide-binding test showed that actin patches in yeast arp3-G302Y mutant become almost completely depolarized when shifted to the restrictive temperature of 37 • C, while actin patches in arp2-G302Y mutant are polarized at both 25 and 37 • C. Similarly, the arp3-G302Y mutants are significantly impaired in Lucifer yellow uptake (an indicator of endocytic function), while arp2-G302Y only exhibits moderate reduction, which indicates the importance of ARP3 over ARP2 for endocytosis in yeast. In addition, only ARP3 is necessary to assemble the ARP2/3 complex, since fission yeast that lacks ARP2 can still form the complex and has all of the other subunits in their normal positions (Nolen and Pollard, 2008). The actin patches formed in Arabidopsis-powdery mildew interactions are assembled by the ARP2/3 complex, with the help of Class I formins (Qin et al., 2021). ...
... It has been previously reported that the key subunits of the ARP2/3 complex, ARP2 and ARP3, show some differences in several cellular processes in yeast, although they are indispensable in the initiation of branched filaments. For example, only ARP3 but not ARP2 is necessary for the assembly of the ARP2/3 complex, and arp3 mutants show more severe defects while arp2 mutants display moderate defects during the endocytosis in fission yeast (Nolen and Pollard, 2008). However, in Arabidopsis, the present study identified equal importance and essential non-redundant roles for both ARP2 and ARP3. ...
Article
Full-text available
In plants, the actin cytoskeleton plays a critical role in defense against diverse pathogens. The formation of actin patches is essential for the intracellular transport of organelles and molecules toward pathogen penetration sites and the formation of papillae for an early cellular response to powdery mildew attack in Arabidopsis thaliana. This response process is regulated by the actin-related protein (ARP)2/3 complex and its activator, the WAVE/SCAR complex (W/SRC). The ARP2/3 complex is also required for maintaining steady-state levels of the defense-associated protein, PENETRATION 1 (PEN1), at the plasma membrane and for its deposition into papillae. However, specific ARP2 functionalities in this context remain unresolved, as knockout mutants expressing GFP-PEN1 reporter constructs could not be obtained by conventional crossing approaches. In this study, employing a CRISPR/Cas9 multiplexing-mediated genome editing approach, we produced an ARP2 knockout expressing the GFP-PEN1 marker in Arabidopsis. This study successfully identified diallelic somatic mutations with both ARP2 alleles edited among the primary T1 transgenic plants, and also obtained independent lines with stable arp2/arp2 mutations in the T2 generation. Further analyses on these arp2/arp2 mutants showed similar biological functions of ARP2 to ARP3 in the accumulation of PEN1 against fungal invasion. Together, this CRISPR/Cas9-based approach offers highly efficient simultaneous disruption of the two ARP2 alleles in GFP-PEN1-expressing lines, and a rapid method for performing live-cell imaging to facilitate the investigation of important plant–pathogen interactions using a well-established and widely applied GFP marker system, thus gaining insights and elucidating the contributions of ARP2 upon fungal attack.
... Arp2/3 complex Saccharomyces cerevisiae Arp2/3 complex was purified from commercially purchased baker's yeast (L'Hirondelle) based on a protocol modified from (Nolen & Pollard, 2008;Doolittle et al, 2013;Antkowiak et al, 2019). Yeast powder was prepared by flash freezing droplets of liquid yeast culture in liquid nitrogen and grinding them in a steel blender. ...
... The mixture was centrifuged at 160,000 g for 30 min and the supernatant was fractioned by a 50% ammonium sulfate cut. The insoluble fraction was dissolved, dialyzed in HKME buffer (25 mM of Hepes, pH 7.5, 50 mM of KCl, 1 mM of EGTA, 3 mM of MgCl 2 , 1 mM of DTT, 0.1 mM of ATP) overnight at 4°C and loaded onto a 2-ml Glutathione-Sepharose 4B (GE Healthcare Life Sciences, Piscataway, NJ, USA) column precharged with GST-N-WASp-VCA (Nolen & Pollard, 2008;Doolittle et al, 2013, 3;Antkowiak et al, 2019). The column was washed with HKME buffer and bound Arp2/3 was eluted with 20 mM of Tris-HCl pH 7.5, 25 mM of KCl, 200 mM of MgCl 2 , 1 mM of EGTA, and 1 mM of DTT. ...
Article
Full-text available
A paradox of eukaryotic cells is that while some species assemble a complex actin cytoskeleton from a single ortholog, other species utilize a greater diversity of actin isoforms. The physiological consequences of using different actin isoforms, and the molecular mechanisms by which highly conserved actin isoforms are segregated into distinct networks, are poorly known. Here, we sought to understand how a simple biological system, composed of a unique actin and a limited set of actin-binding proteins, reacts to a switch to heterologous actin expression. Using yeast as a model system and biomimetic assays, we show that such perturbation causes drastic reorganization of the actin cytoskeleton. Our results indicate that defective interaction of a heterologous actin for important regulators of actin assembly limits certain actin assembly pathways while reinforcing others. Expression of two heterologous actin variants, each specialized in assembling a different network, rescues cytoskeletal organization and confers resistance to external perturbation. Hence, while species using a unique actin have homeostatic actin networks, actin assembly pathways in species using several actin isoforms may act more independently.
... Specifically, the loop 184-189 and the first five residues of the helix 190-206 of CH1 are sufficiently close to make contacts at both sides (subdomains 1 and 3) of the hydrophobic cleft of the "i" actin subunit but also with the 44-48 segment of the D-loop (subdomain 2) of the longitudinally adjacent subunit "i+2". Also, in subdomain 1 of the "i+2" subunit, the tip of [79][80][81][82][83][84][85][86][87][88][89][90][91][92][93][94][95] helix is in proximity to residues Gln118, Asn213, and His225 of plastin's CH1. With the exception of Asn213, all these PLS3 residues are different from those determined in a previous model based on a ~ 30-Å resolution reconstruction of ABD1-decorated actin filaments 65 . ...
... Unlabeled ABD2 was added from 0 to 3000 nM. The dissociation constant of the unlabeled protein was determined by fitting the data to the following equation 95 : ...
Preprint
Plastins/fimbrins are conserved actin-bundling proteins contributing to motility, cytokinesis, and other cellular processes by organizing actin assemblies of strikingly different geometries as in aligned bundles and branched networks. We propose that this unique ability stems from an allosteric communication between the two actin-binding domains (ABD1/2) engaged in a tight spatial association. We found that although ABD1 binds actin first, ABD2 can bind to actin three orders of magnitude stronger if not inhibited by an equally strong allosteric engagement with ABD1. Binding of ABD1 to actin lessened the inhibition, enabling weak bundling within aligned bundles. A mutation mimicking physiologically relevant phosphorylation at the ABD1-ABD2 interface strongly reduced their association, dramatically potentiating actin cross-linking. Cryo-EM reconstruction revealed the ABD1-actin interface and enabled modeling of the plastin bridge to confirm domain separation in parallel bundles. The characteristic domain organization with a strong allosteric inhibition imposed by ABD1 on ABD2 allows plastins to tune cross-linking, contributing to the assembly and remodeling of actin assemblies with different morphological and functional properties defining the unique place of plastins in actin organization.
... Actin (Act1), the actin filament depolymerization factor Cofilin (Adf1), and Pan1 and Ede1 of the complex that links actin to endocytic vesicles were identified as a Pef1 neighbors in both exponential and autophagic cells (Supplementary Table S7). Arp2 and Arc2, part of a complex that allows for actin filament branching (49)(50)(51)(52), were p. 25 of 45 identified in autophagic cells. Pef1 regulates autophagy, and actin filament dynamics are important for various stages of autophagy (53)(54)(55)(56), and these neighbors may reveal how Pef1 regulates this process. ...
... Actin (Act1), the actin filament depolymerization factor Cofilin (Adf1), and Pan1 and Ede1 of the complex that links actin to endocytic vesicles were identified as a Pef1 neighbors in both exponential and autophagic cells (Supplementary Table S7). Arp2 and Arc2, part of a complex that allows for actin filament branching (49)(50)(51)(52), were . CC-BY-NC 4.0 International license available under a (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. ...
Preprint
Full-text available
Cdk5 is a highly-conserved, noncanonical cell division kinase important to the terminal differentiation of mammalian cells in multiple organ systems. We previously identified Pef1, the Schizosaccharomyces pombe ortholog of cdk5, as regulator of chronological lifespan. To reveal the processes impacted by Pef1, we developed APEX2-biotin phenol-mediated proximity labeling in S. pombe. Efficient labeling required a short period of cell wall digestion and eliminating glucose and nitrogen sources from the medium. We identified 255 high-confidence Pef1 neighbors in growing cells and a novel Pef1-interacting partner, the DNA damage response protein Rad24. The Pef1-Rad24 interaction was validated by reciprocal proximity labeling and co-immunoprecipitation. Eliminating Pef1 partially rescued the DNA damage sensitivity of cells lacking Rad24. To monitor how Pef1 neighbors change under different conditions, cells induced for autophagy were labeled and 177 high-confidence Pef1 neighbors were identified. Gene ontology (GO) analysis of the Pef1 neighbors identified proteins participating in processes required for autophagosome expansion including regulation of actin dynamics and vesicle-mediated transport. Some of these proteins were identified in both exponentially growing and autophagic cells. Pef1-APEX2 proximity labeling therefore identified a new Pef1 function in modulating the DNA damage response and candidate processes that Pef1 and other cdk5 orthologs may regulate.
... The relative fluorescence anisotropy changes were normalized to the calculated bound fractions (0.60 for actin and 0.52 for modified actin) and fitted to the following equation [83] using Origin software (OriginLab, Northampton, MA, USA) to estimate the binding affinity of actin to the TMSB4: ...
Article
Full-text available
Due to its essential role in cellular processes, actin is a common target for bacterial toxins. One such toxin, TccC3, is an effector domain of the ABC-toxin produced by entomopathogenic bacteria of Photorhabdus spp. Unlike other actin-targeting toxins, TccC3 uniquely ADP-ribosylates actin at Thr-148, resulting in the formation of actin aggregates and inhibition of phagocytosis. It has been shown that the fully modified F-actin is resistant to depolymerization by cofilin and gelsolin, but their effects on partially modified actin were not explored. We found that only F-actin unprotected by tropomyosin is the physiological TccC3 substrate. Yet, ADP-ribosylated G-actin can be produced upon cofilin-accelerated F-actin depolymerization, which was only mildly inhibited in partially modified actin. The affinity of TccC3-ADP-ribosylated G-actin for profilin and thymosin-β4 was weakened moderately but sufficiently to potentiate spontaneous polymerization in their presence. Interestingly, the Arp2/3-mediated nucleation was also potentiated by T148-ADP-ribosylation. Notably, even partially modified actin showed reduced bundling by plastins and α-actinin. In agreement with the role of these and other tandem calponin-homology domain actin organizers in the assembly of the cortical actin network, TccC3 induced intense membrane blebbing in cultured cells. Overall, our data suggest that TccC3 imposes a complex action on the cytoskeleton by affecting F-actin nucleation, recycling, and interaction with actin-binding proteins involved in the integration of actin filaments with each other and cellular elements.
... Unlabeled ABD2 was added from 0 to 3,000 nM. The dissociation constant of the unlabeled protein was determined by fitting the data to the following equation 83 : ...
Article
Full-text available
Plastins/fimbrins are conserved actin-bundling proteins contributing to motility, cytokinesis and other cellular processes by organizing strikingly different actin assemblies as in aligned bundles and branched networks. We propose that this ability of human plastins stems from an allosteric communication between their actin-binding domains (ABD1/2) engaged in a tight spatial association. Here we show that ABD2 can bind actin three orders of magnitude stronger than ABD1, unless the domains are involved in an equally strong inhibitory engagement. A mutation mimicking physiologically relevant phosphorylation at the ABD1–ABD2 interface greatly weakened their association, dramatically potentiating actin cross-linking. Cryo-EM reconstruction revealed the ABD1–actin interface and enabled modeling of the plastin bridge and domain separation in parallel bundles. We predict that a strong and tunable allosteric inhibition between the domains allows plastins to modulate the cross-linking strength, contributing to remodeling of actin assemblies of different morphologies defining the unique place of plastins in actin organization.
Article
Full-text available
Disruption of intercellular communication within tumors is emerging as a novel potential strategy for cancer-directed therapy. Tumor-Treating Fields (TTFields) therapy is a treatment modality that has itself emerged over the past decade in active clinical use for patients with glioblastoma and malignant mesothelioma, based on the principle of using low-intensity alternating electric fields to disrupt microtubules in cancer cells undergoing mitosis. There is a need to identify other cellular and molecular effects of this treatment approach that could explain reported increased overall survival when TTFields are added to standard systemic agents. Tunneling nanotube (TNTs) are cell-contact-dependent filamentous-actin-based cellular protrusions that can connect two or more cells at long-range. They are upregulated in cancer, facilitating cell growth, differentiation, and in the case of invasive cancer phenotypes, a more chemoresistant phenotype. To determine whether TNTs present a potential therapeutic target for TTFields, we applied TTFields to malignant pleural mesothelioma (MPM) cells forming TNTs in vitro. TTFields at 1.0 V/cm significantly suppressed TNT formation in biphasic subtype MPM, but not sarcomatoid MPM, independent of effects on cell number. TTFields did not significantly affect function of TNTs assessed by measuring intercellular transport of mitochondrial cargo via intact TNTs. We further leveraged a spatial transcriptomic approach to characterize TTFields-induced changes to molecular profiles in vivo using an animal model of MPM. We discovered TTFields induced upregulation of immuno-oncologic biomarkers with simultaneous downregulation of pathways associated with cell hyperproliferation, invasion, and other critical regulators of oncogenic growth. Several molecular classes and pathways coincide with markers that we and others have found to be differentially expressed in cancer cell TNTs, including MPM specifically. We visualized short TNTs in the dense stromatous tumor material selected as regions of interest for spatial genomic assessment. Superimposing these regions of interest from spatial genomics over the plane of TNT clusters imaged in intact tissue is a new method that we designate Spatial Profiling of Tunneling nanoTubes (SPOTT). In sum, these results position TNTs as potential therapeutic targets for TTFields-directed cancer treatment strategies. We also identified the ability of TTFields to remodel the tumor microenvironment landscape at the molecular level, thereby presenting a potential novel strategy for converting tumors at the cellular level from ‘cold’ to ‘hot’ for potential response to immunotherapeutic drugs.
Article
Full-text available
We reconstructed the structure of actin filament branch junctions formed by fission yeast Arp2/3 complex at 3.5 Å resolution from images collected by electron cryo-microscopy. During specimen preparation, all of the actin subunits and Arp3 hydrolyzed their bound adenosine triphosphate (ATP) and dissociated the γ-phosphate, but Arp2 retained the γ-phosphate. Binding tightly to the side of the mother filament and nucleating the daughter filament growing as a branch requires Arp2/3 complex to undergo a dramatic conformational change where two blocks of structure rotate relative to each other about 25° to align Arp2 and Arp3 as the first two subunits in the branch. During branch formation, Arp2/3 complex acquires more than 8,000 Å2 of new buried surface, accounting for the stability of the branch. Inactive Arp2/3 complex binds only transiently to the side of an actin filament, because its conformation allows only a subset of the interactions found in the branch junction.
Preprint
We used fluorescence spectroscopy and electron microscopy to determine how binding of ATP, nucleation-promoting factors (NPF), actin monomers and actin filaments change the conformation of Arp2/3 complex during the process that nucleates an actin filament branch. We mutated subunits of Schizosaccharomyces pombe Arp2/3 complex for labeling with fluorescent dyes at either the C-termini of Arp2 and Arp3 or ArpC1 and ArpC3. We measured Förster resonance energy transfer (FRET) efficiency (ET eff ) between the dyes in the presence of the various ligands. We also computed class averages from electron micrographs of negatively stained specimens. ATP binding made small conformational changes of the nucleotide binding clefts of the Arp subunits. WASp-VCA, WASp-CA, and WASp-actin-VCA changed the ET eff between the dyes on the Arp2 and Arp3 subunits much more than between dyes on ArpC1 and ArpC3. Ensemble FRET detected a different structural change that involves bringing ArpC1 and ArpC3 closer together when Arp2/3 complex bound actin filaments. Each of the ligands that activates Arp2/3 complex changes the structure in different ways, each leading progressively to fully activated Arp2/3 complex on the side of a filament.
Article
Full-text available
Phaser is a program for phasing macromolecular crystal structures by both molecular replacement and experimental phasing methods. The novel phasing algorithms implemented in Phaser have been developed using maximum likelihood and multivariate statistics. For molecular replacement, the new algorithms have proved to be significantly better than traditional methods in discriminating correct solutions from noise, and for single-wavelength anomalous dispersion experimental phasing, the new algorithms, which account for correlations between F + and F −, give better phases (lower mean phase error with respect to the phases given by the refined structure) than those that use mean F and anomalous differences ΔF. One of the design concepts of Phaser was that it be capable of a high degree of automation. To this end, Phaser (written in C++) can be called directly from Python, although it can also be called using traditional CCP4 keyword-style input. Phaser is a platform for future development of improved phasing methods and their release, including source code, to the crystallographic community.
Article
Full-text available
Using hexokinase, glucose, and ATP to vary reversibly the concentrations of ADP and ATP in solution and bound to Acanthamoeba actin, I measured the relative critical concentrations and elongation rate constants for ATP-actin and ADP-actin in 50 mM KCl, 1 mM MgCl2, 1 mM EGTA, 0.1 mM nucleotide, 0.1 mM CaCl2, 10 mM imidazole, pH 7. By both steady-state and elongation rate methods, the critical concentrations are 0.1 microM for ATP-actin and 5 microM for ADP-actin. Consequently, a 5 microM solution of actin can be polymerized, depolymerized, and repolymerized by simply cycling from ATP to ADP and back to ATP. The critical concentrations differ, because the association rate constant is 10 times higher and the dissociation rate constant is five times lower for ATP-actin than ADP-actin. These results show that ATP-actin occupies both ends of actin filaments growing in ATP. The bound ATP must be split on internal subunits and the number of terminal subunits with bound ATP probably depends on the rate of growth.
Article
Full-text available
Homologous integration into the fission yeast Schizosaccharomyces pombe has not been well characterized. In this study, we have examined integration of plasmids carrying the leu1+ and ura4+ genes into their chromosomal loci. Genomic DNA blot analysis demonstrated that the majority of transformants have one or more copies of the plasmid vector integrated via homologous recombination with a much smaller fraction of gene conversion to leu1+ or ura4+. Non-homologous recombination events were not observed for either gene. We describe the construction of generally useful leu1+ and ura4+ plasmids for targeted integration at the leu1-32 and ura4-294 loci of S. pombe.
Article
Full-text available
The gene encoding the actin-related protein Arp3 was first identified in the fission yeast Schizosaccharomyces pombe and is a member of an evolutionarily conserved family of actin-related proteins. Here we present several key findings that define an essential role for Arp3p in the functioning of the cortical actin cytoskeleton. First, mutants in arp3 interact specifically with profilin and actin mutants. Second, Arp3 localizes to cortical actin patches which are required for polarized cell growth. Third, the arp3 gene is required for the reorganization of the actin cytoskeleton during the cell cycle. Finally, the Arp3 protein is present in a large protein complex. We believe that this complex may mediate the cortical functions of profilin at actin patches in S. pombe.
Article
Full-text available
The Arp2/3 complex, first isolated from Acanthamoeba castellani by affinity chromatography on profilin, consists of seven polypeptides; two actin-related proteins, Arp2 and Arp3; and five apparently novel proteins, p40, p35, p19, p18, and p14 (Machesky et al., 1994). The complex is homogeneous by hydrodynamic criteria with a Stokes' radius of 5.3 nm by gel filtration, sedimentation coefficient of 8.7 S, and molecular mass of 197 kD by analytical ultracentrifugation. The stoichiometry of the subunits is 1:1:1:1:1:1:1, indicating the purified complex contains one copy each of seven polypeptides. In electron micrographs, the complex has a bilobed or horseshoe shape with outer dimensions of approximately 13 x 10 nm, and mathematical models of such a shape and size are consistent with the measured hydrodynamic properties. Chemical cross-linking with a battery of cross-linkers of different spacer arm lengths and chemical reactivities identify the following nearest neighbors within the complex: Arp2 and p40; Arp2 and p35; Arp3 and p35; Arp3 and either p18 or p19; and p19 and p14. By fluorescent antibody staining with anti-p40 and -p35, the complex is concentrated in the cortex of the ameba, especially in linear structures, possibly actin filament bundles, that lie perpendicular to the leading edge. Purified Arp2/3 complex binds actin filaments with a Kd of 2.3 microM and a stoichiometry of approximately one complex molecule per actin monomer. In electron micrographs of negatively stained samples, Arp2/3 complex decorates the sides of actin filaments. EDC/NHS cross-links actin to Arp3, p35, and a low molecular weight subunit, p19, p18, or p14. We propose structural and topological models for the Arp2/3 complex and suggest that affinity for actin filaments accounts for the localization of complex subunits to actin-rich regions of Acanthamoeba.
Article
Full-text available
As in many other eukaryotic cells, cell division in fission yeast depends on the assembly of an actin ring that circumscribes the middle of the cell. Schizosaccharomyces pombe cdc12 is an essential gene necessary for actin ring assembly and septum formation. Here we show that cdc12p is a member of a family of proteins including Drosophila diaphanous, Saccharomyces cerevisiae BNI1, and S. pombe fus1, which are involved in cytokinesis or other actin-mediated processes. Using indirect immunofluorescence, we show that cdc12p is located in the cell division ring and not in other actin structures. When overexpressed, cdc12p is located at a medial spot in interphase that anticipates the future ring site. cdc12p localization is altered in actin ring mutants. cdc8 (tropomyosin homologue), cdc3 (profilin homologue), and cdc15 mutants exhibit no specific cdc12p staining during mitosis. cdc4 mutant cells exhibit a medial cortical cdc12p spot in place of a ring. mid1 mutant cells generally exhibit a cdc12p spot with a single cdc12p strand extending in a random direction. Based on these patterns, we present a model in which ring assembly originates from a single point on the cortex and in which a molecular pathway for the functions of cytokinesis proteins is suggested. Finally, we found that cdc12 and cdc3 mutants show a synthetic-lethal genetic interaction, and a proline-rich domain of cdc12p binds directly to profilin cdc3p in vitro, suggesting that one function of cdc12p in ring assembly is to bind profilin.
Article
We determined a crystal structure of bovine Arp2/3 complex, an assembly of seven proteins that initiates actin polymerization in eukaryotic cells, at 2.0 angstrom resolution. Actin-related protein 2 (Arp2) and Arp3 are folded like actin, with distinctive surface features. Subunits ARPC2 p34 and ARPC4 p20 in the core of the complex associate through long carboxyl-terminal α helices and have similarly folded amino-terminal α/β domains. ARPC1 p40 is a seven-blade β propeller with an insertion that may associate with the side of an actin filament. ARPC3 p21 and ARPC5 p16 are globular α-helical subunits. We predict that WASp/Scar proteins activate Arp2/3 complex by bringing Arp2 into proximity with Arp3 for nucleation of a branch on the side of a preexisting actin filament.
Article
Gel filtration of depolymerized conventionally purified muscle actin separates from the actin monomers a fraction of minor contaminants with a Stokes' radius of 4.7 nm which has the ability to block actin filament network formation. On the basis of heat and trypsin sensitivity, this inhibitory activity appears to be a protein. The inhibitory activity binds to actin filaments and reduces their low shear viscosity by up to 99% in a concentration dependent fashion while reducing polymerization to only a minor extent.
Article
Budding and fission yeast present significant advantages for studies of the actin cytoskeleton. The application of classical and molecular genetic techniques provides a facile route for the analysis of structure/function relationships, for the isolation of novel proteins involved in cytoskeletal function, and for deciphering the signals that regulate actin assembly in vivo. This review focuses on the budding yeast Saccharomyces cerevisiae and also identifies some recent advances from studies on the fission yeast Schizosaccharomyces pombe, for which studies on the actin cytoskeleton are still in their infancy.
Article
An important recent advance in the functional analysis of Saccharomyces cerevisiae genes is the development of the one-step PCR-mediated technique for deletion and modification of chromosomal genes. This method allows very rapid gene manipulations without requiring plasmid clones of the gene of interest. We describe here a new set of plasmids that serve as templates for the PCR synthesis of fragments that allow a variety of gene modifications. Using as selectable marker the S. cerevisiae TRP1 gene or modules containing the heterologous Schizosaccharomyces pombe his5+ or Escherichia coli kan(r) gene, these plasmids allow gene deletion, gene overexpression (using the regulatable GAL1 promoter), C- or N-terminal protein tagging [with GFP(S65T), GST, or the 3HA or 13Myc epitope], and partial N- or C-terminal deletions (with or without concomitant protein tagging). Because of the modular nature of the plasmids, they allow efficient and economical use of a small number of PCR primers for a wide variety of gene manipulations. Thus, these plasmids should further facilitate the rapid analysis of gene function in S. cerevisiae.