ArticlePDF Available

Dectin-2 Is a Pattern Recognition Receptor for Fungi That Couples with the Fc Receptor Chain to Induce Innate Immune Responses

Authors:
  • National Veterinary Assay Laboratory

Abstract and Figures

Antigen presenting cells recognize pathogens via pattern recognition receptors (PRR), which upon ligation transduce intracellular signals that can induce innate immune responses. Because some C-type lectin-like receptors (e.g. dectin-1 and DCSIGN) were shown to act as PRR for particular microbes, we considered a similar role for dectin-2. Binding assays using soluble dectin-2 receptors showed the extracellular domain to bind preferentially to hyphal (rather than yeast/conidial) components of Candida albicans, Microsporum audouinii, and Trichophyton rubrum. Selective binding for hyphae was also observed using RAW macrophages expressing dectin-2, the ligation of which by hyphae or cross-linking with dectin-2-specific antibody led to protein tyrosine phosphorylation. Because dectin-2 lacks an intracellular signaling motif, we searched for a signal adaptor that permits it to transduce intracellular signals. First, we found that the Fc receptor γ (FcRγ) chain can bind to dectin-2. Second, ligation of dectin-2 on RAW cells induced tyrosine phosphorylation of FcRγ, activation of NF-κB, internalization of a surrogate ligand, and up-regulated secretion of tumor necrosis factor α and interleukin-1 receptor antagonist. Finally, these dectin-2-induced events were blocked by PP2, an inhibitor of Src kinases that are mediators for FcRγ chain-dependent signaling. We conclude that dectin-2 is a PRR for fungi that employs signaling through FcRγ to induce innate immune responses.
Dectin-2 on RAW cells recognizes C. albicans hyphae. A, protein expression of dectin-1 and dectin-2 in RAW macrophage transfectants. Whole cell extracts were prepared from parental RAW macrophages (M) or those transfected with dectin-1 or dectin-2 tagged with the C-terminal V5 epitope (Dec1V5 or Dec2V5). Aliquots (each 1 10 5 cells eq) were examined for protein expression of Dec1V5 or Dec2V5 by Western blotting using anti-V5 Ab. Two arrows (31 and 42 kDa) indicate bands corresponding to Dec1V5 and Dec2V5, respectively. B, RAW cells were pretreated with Fc block before staining with FITC/anti-V5 Ab. Surface expression of Dec1V5 (open histogram) or Dec2V5 (closed histogram) was assayed by flow cytometry and compared with parental cells (dashed lines). C, binding to yeast. RAW cells were cocultured with fluorescence-labeled live yeasts at varying m.o.i. values. Frequency (%) of fluorescently labeled RAW cells was determined by flow cytometry. D, binding to hyphae. RAW cells metabolically labeled with [ 3 H]thymidine were incubated in the 96-well plate covered with/without hyphae (4 10 5 cells/well) at 37 °C for 30 min. After washing, cells were lysed and measured for 3 H counts, and cell numbers adhered to a well were computed. Binding of RAW cells to hyphae is expressed as fold difference (hyphae versus culture well). E, COS-1 cells were transfected with an empty vector (open triangles) or expression vector for dectin-1-V5 (open circles) or dectin- 2-V5 (closed circles). [ 3 H]Thymidine-labeled COS-1 cells at increasing cell numbers were incubated with a constant number of hyphae (2 10 5 cells/well in triplicate). After washing, hyphae-adhered COS-1 cells were lysed and measured for 3 H counts/min. The number of cells that bound to hyphae was calculated as well bound cpm/specific activity of 3 H-labeled cells (0.24 – 0.32 cpm/cell). F, dectin-2 protein on RAW cells was fluorescently labeled with FITC/anti-V5 Ab and then cocultured with pseudohyphae labeled with rhodamine (red fluorescence). Green (anti-V5 Ab) and red (hyphae) fluorescence images were separately taken and merged (Merge). Surface distribution of dectin-2 before coculturing with hyphae was also shown (Before). All data shown are representative of three independent experiments.
… 
Content may be subject to copyright.
Dectin-2 Is a Pattern Recognition Receptor for Fungi That
Couples with the Fc Receptor
Chain to Induce Innate
Immune Responses
*
Received for publication, July 10, 2006, and in revised form, October 12, 2006 Published, JBC Papers in Press, October 18, 2006, DOI 10.1074/jbc.M606542200
Kota Sato
‡1
, Xiao-li Yang
, Tatsuo Yudate
‡2
, Jin-Sung Chung
, Jianming Wu
§
, Kate Luby-Phelps
,
Robert P. Kimberly
§
, David Underhill
, Ponciano D. Cruz, Jr.
, and Kiyoshi Ariizumi
‡3
From the
Department of Dermatology, the University of Texas Southwestern Medical Center and Dermatology Section (Medical
Service), Dallas Veterans Affairs Medical Center, Dallas, Texas 75390,
Institute for Systems Biology, Seattle, Washington 98103,
the
Department of Cell Biology, the University of Texas Southwestern Medical Center, Dallas, Texas 75390, and
§
Division of Clinical
Immunology and Rheumatology, Department of Medicine, University of Alabama at Birmingham, Birmingham, Alabama 35294
Antigen presenting cells recognize pathogens via pattern rec-
ognition receptors (PRR), which upon ligation transduce intra-
cellular signals that can induce innate immune responses.
Because some C-type lectin-like receptors (e.g. dectin-1 and DC-
SIGN) were shown to act as PRR for particular microbes, we
considered a similar role for dectin-2. Binding assays using sol-
uble dectin-2 receptors showed the extracellular domain to bind
preferentially to hyphal (rather than yeast/conidial) compo-
nents of Candida albicans, Microsporum audouinii, and
Trichophyton rubrum. Selective binding for hyphae was also
observed using RAW macrophages expressing dectin-2, the
ligation of which by hyphae or cross-linking with dectin-2-spe-
cific antibody led to protein tyrosine phosphorylation. Because
dectin-2 lacks an intracellular signaling motif, we searched for a
signal adaptor that permits it to transduce intracellular signals.
First, we found that the Fc receptor
(FcR
) chain can bind to
dectin-2. Second, ligation of dectin-2 on RAW cells induced
tyrosine phosphorylation of FcR
, activation of NF-
B, inter-
nalization of a surrogate ligand, and up-regulated secretion of
tumor necrosis factor
and interleukin-1 receptor antagonist.
Finally, these dectin-2-induced events were blocked by PP2, an
inhibitor of Src kinases that are mediators for FcR
chain-de-
pendent signaling. We conclude that dectin-2 is a PRR for fungi
that employs signaling through FcR
to induce innate immune
responses.
To initiate immune responses against infection, antigen pre-
senting cells (APC)
4
must recognize and react to microbes. Rec-
ognition is achieved by interaction of particular surface recep-
tors on APC with corresponding surface molecules on
infectious agents (1). Complement and Fc receptors bind
microbes coated with opsonin (1). By contrast, pattern recog-
nition receptors (PRR) recognize and interact with pathogens
directly (1, 2). PRR include the following: (a) scavenger recep-
tors that bind low density lipoproteins or lipid A on some bac-
teria (3); (b) toll-like receptors (TLR) that bind zymosan, Staph-
ylococcus aureus, lipopolysaccharide (LPS), bacterial flagellin,
or CpG bacterial DNA (4 6); and (c) C-type lectin-like recep-
tors (CLR) that bind carbohydrate moieties of many pathogens
(1, 7). CLR include the following: (a) mannose receptors for
mannose or its polymers (8); (b) mannose-binding lectins for
encapsulated group B or C meningococci (9); (c) DC-SIGN and
structurally related receptors (DC-SIGNR) for mannose on
human immunodeficiency virus, Leishmania, and Mycobacte-
ria (9 –14); and (d) dectin-1 for
-glucan on yeasts (15, 16).
Binding of pathogens to particular PRR transduce intracellu-
lar signals and biologic consequences that may overlap, even
synergize, with those of other PRR. For example, ligation of
TLR2 alone on macrophages by zymosan (containing
-glucan)
led to secretion of IL-12 and TNF
, and ligation of dectin-1
alone by zymosan resulted in production of reactive oxygen
species (but not of IL-12 nor TNF
), whereas coligation of
TLR-2 and dectin-1 by zymosan enhanced secretion of IL-12
and TNF
at levels higher than those induced by TLR-2 alone
(17). On the other hand, ligation of DC-SIGN on dendritic cells
(DC) inhibited TLR-induced IL-12 expression, while stimulat-
ing IL-10 expression (12).
Subtractive cDNA cloning of the XS52 line of epidermal
Langerhans cell-like DC (18) minus J774 macrophages led us to
discover dectin-1 (19) and dectin-2 (20). Both are type II-con-
figured transmembrane proteins with extracellular domains
containing a carbohydrate recognition domain highly con-
served among C-type lectins (19, 20). Dectin-1 is expressed
widely by APC (21) and is a PRR for
-glucan in yeasts (15).
* The costs of publication of this article were defrayed in part by the payment
of page charges. This article must therefore be hereby marked advertise-
ment in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
1
Present address: Dept. of Pathology, Graduate School of Veterinary Medi-
cine, Hokkaido University, Kita 18 Nishi 9, Kita-ku, Sapporo 060-0818,
Japan.
2
Present address: Dept. of Dermatology, Kinki University School of Medicine,
377-2, Ohno-Higashi, Osaka-Sayama, Osaka 589-8511, Japan.
3
To whom correspondence should be addressed: Dept. of Dermatology, Uni-
versity of Texas Southwestern Medical Center, 5323 Harry Hines Blvd.,
Dallas, TX 75390-9069. Tel.: 214-648-7552; Fax: 214-648-0280; E-mail:
Kiyoshi.Ariizumi@UTSouthwestern.edu.
4
The abbreviations used are: APC, antigen presenting cells; Ab, antibody;
mAb, monoclonal antibody; CLR, C-type lectin-like receptors; DC, dendritic
cells; EMSA, electromobility shift assay; FcR
, Fc receptor
; IL-1ra, interleu-
kin-1 receptor antagonist; ITAM, immunoreceptor tyrosine-based activa-
tion motif; m.o.i., multiplication of infection; PRR, pattern recognition
receptors; TLR, toll-like receptor; TNF
, tumor necrosis factor
; FITC, fluo-
rescein isothiocyanate; TRITC, tetramethylrhodamine isothiocyanate; BSA,
bovine serum albumin; PBS, phosphate-buffered saline; DPBS, Dulbecco’s
PBS; FCS, fetal calf serum; FACS, fluorescence-activated cell sorter; LPS,
lipopolysaccharide; HBSS, Hanks’ balanced salt solution.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 281, NO. 50, pp. 38854 –38866, December 15, 2006
Printed in the U.S.A.
38854 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
Dectin-2 is constitutively expressed at very high levels by
mature DC and can be inducibly expressed on macrophages
after activation (20, 22). Here we report that dectin-2 is a PRR
for fungi that employ Fc receptor
(FcR
) chain signaling to
induce internalization, activate NF-
B, and up-regulate pro-
duction of TNF
and IL-1ra.
EXPERIMENTAL PROCEDURES
Microbial Cell Cultures—We obtained Candida albicans
(ATCC 10231 and 14053), Microsporum audouinii (ATCC
10008), Trichophyton rubrum (ATCC 14001), and Pseudomo-
nas aeruginosa (ATCC 10145) from the American Type Cul-
ture Collection; Escherichia coli DH5
from Invitrogen; Staph-
ylococcus aureus without protein A from Molecular Probes Inc.
(Eugene, OR); Saccharomyces cerevisiae Y187 from Clontech;
and group A Streptococci from the Section of Infectious Dis-
ease, Department of Pediatrics, the University of Texas South-
western Medical Center (Dallas, TX). Each microbial strain was
grown in media recommended by the ATCC. C. albicans yeast
transformed to pseudohyphae (herein referred to as hyphae) as
follows. Freshly prepared yeast was resuspended in Hanks’ bal-
anced salt solution (HBSS) containing 1.25 m
M CaCl
2
,1mM
MgCl
2
,10mM HEPES, pH 7.2, and 10% heat-inactivated FCS,
seeded on 96-well plates or ELISA plates (2– 4 10
5
cells/well),
and then incubated at 37 °C for 90 min.
Construction of Expression Vectors—To produce soluble dec-
tin-1 and dectin-2 receptors, we inserted a nucleotide fragment
encoding the extracellular domain of either molecule into an
expression vector, pSTB-Fc (23), that allows secretion of the Fc
portion of human IgG
1
into the culture supernatant of mam
-
malian cells. Respective nucleotide fragments encoding for
extracellular domains of dectin-1 and dectin-2 were obtained
by PCR amplification of the full-length cDNA with primers
containing BamHI (forward primer) and XbaI (reverse primer)
restriction enzyme sites at the 5-end for dectin-1 or containing
HindIII and XbaI sites for dectin-2. PCR fragments remaining
after digestion with restriction enzymes were linked separately
in-frame to the 5-end of a nucleotide for the Fc in pSTB-Fc
(pSTB-Dec1-Fc or pSTB-Dec2-Fc).
Lentiviral vectors encoding dectin-2 or dectin-1 tagged with
the C-terminal V5 epitope were also constructed. Full-length dec-
tin-2- or dectin-1-coding sequence was excised from an original
cDNA clone (20) by PCR amplification with the forward primer
containing a HindIII (or BamHI) restriction site and the reverse
primer containing an ApaI site linked to a sequence (TACCCCT-
ACGACGTGCCCGACTACGCC) encoding for a V5 epitope
(GKPIPNPLLGLDST) at the 5-end. Using these restriction sites,
the PCR product was inserted into a mammalian expression vec-
tor, pcDNA3.1 (Invitrogen) (pcDNA-Dec2V5 or pcDNA-
Dec1V5). The nucleotide sequence for dectin-2-V5 (or dectin-
1-V5) was excised from pcDNA-Dec2V5 (or pcDNA-Dec1V5)
by restriction enzyme digestion with PmeI (a blunt end cutter)
and NotI. The lentiviral vector plasmid, pHR-SIN-CSGW
dlNotI (24) (gift from Y. Ikeda, Mayo Clinic, Rochester, MN),
was digested with BamHI and NotI restriction enzymes to
remove a nucleotide encoding enhanced green fluorescent pro-
tein. After end-filling the BamHI site with Klenow fragments,
the lentiviral vector was ligated to the nucleotide for dectin-
2-V5 (or dectin-1-V5) using the blunt end and the NotI site.
Preparation of infectious particles and their titration were per-
formed according to established protocols (25).
Mouse FcR
chain expression vector (pcDNA-m
chain) was
constructed as follows. Total RNA prepared from RAW264.7
macrophages was reverse-transcribed to the cDNA form and
amplified using upper (5-ATCGGATCCATGATCTCAGCC-
GTGATCTTG-3, where boldface letters indicate EcoRI site)
and lower (5-GAATTCCTACTGGGGTGGTTTTTCATGC-
3, BamHI) primers. The resulting PCR product (260 bp) was
inserted into pcDNA3.1 using EcoRI and BamHI sites.
To determine how dectin-2 associates with the FcR
chain,
we constructed dectin-2 mutants as follows. Mutant R17V with
arginine replaced by valine (point mutation) at amino acid 17 of
the transmembrane domain was generated following instruc-
tions from the QuikChange site-directed mutagenesis kit
(Stratagene, La Jolla, CA) using the forward primer (5-GGAG-
TCTGCTGGACCCTGGTACTCTGGTCAGCTGCTGTG-
3, boldface letter indicates the mutated nucleotide), and the
reverse primer (5-CACAGCAGCTGACCAGAGTACCAGG-
GTCCAGCAGACTCC-3). Mutant ICD lacking the entire
intracellular domain (amino acids 1–14) was generated by PCR
amplification using the forward primer (5-CGAAGCTTGCC-
ACCATGACCCTGAGACTCTGGTCA-3) containing the
HindIII site (in boldface) and the reverse primer (5-TGTGT-
CCTCGAGTAGGTAAATCTTCTTCATTTC-3) containing
the XhoI site. The resulting PCR fragment was ligated to the
HindIII and XhoI sites of pcDNA-V5 vector that encodes the
C-terminal V5 epitope. The same strategy using a different for-
ward primer (5-CCCAAGCTT (HindIII) GCCACCATGCA-
AGGGAAGGGAGTC-3) was used to generate mutant
1/2ICD in which half the N-terminal intracellular domain
(amino acids 1–7) is deleted. Finally, the chimeric mutant
40LECD was generated by fusing the intracellular and trans-
membrane domains of dectin-2 to the extracellular domain of
CD40 ligand (CD40L). A nucleotide fragment coding for the two
domains of dectin-2 was extracted from dectin-2 cDNA by PCR
amplification using the forward primer (5-CGGCTAGC(NheI
site)GCCACCATGGTGCAGGAAAGACAA-3) and the reverse
primer (5-CGAAGCTT(HindIII)TTGGTAAGTCACCACAC-
AGCT-3). A fragment for the extracellular domain of CD40L was
prepared using the forward primer (5-CGAAGCTT(HindIII)
ATAGAAGATTGGATAAGGTC-3) and the reverse primer
(5-TGTGTCCTCGAG(XhoI)GAGTTTGAGTAAGCC-
3). The two fragments were then subcloned in the NheI-
XhoI sites of pcDNA-V5 vector. Nucleotide sequences of all
mutants were confirmed by sequencing.
Gene Delivery to Mammalian Cells—COS-1 cells (5 10
5
cells/dish) were treated with an expression vector DNA (2
g)
and 6
g of FuGENE 6 (Roche Applied Science) and then cul-
tured for 2–3 days.
RAW264.7 cells (5 10
5
) were infected with lentivirus
encoding dectin-2-V5 (or dectin-1-V5) at a multiplication of
infection (m.o.i.) of 20. The next day, the infected cells were
enriched for surface expression of dectin-2-V5 (or dectin-1-V5)
using immuno-magnetic beads. After blocking Fc receptors
with 5
g/ml Fc block (Pharmingen), infected RAW cells (5
10
5
) were incubated with mouse anti-V5 Ab (2
g/ml; Serotec,
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38855
by guest on December 26, 2015http://www.jbc.org/Downloaded from
Raleigh, NC) and biotinylated goat anti-mouse IgG (5
g/ml,
Jackson ImmunoResearch, West Grove, PA) on ice for 60 min
and treated with streptavidin-coated magnetic beads (Miltenyi,
Auburn, CA). Bead-bound cells were collected and cultured in
RPMI 1640 supplemented with 10% FCS. This enrichment was
repeated 3– 4 times, followed by analysis of the purity of the cell
suspension by FACS. Greater than 90% of RAW cells expressed
dectin-2-V5 (or dectin-1-V5) on their surfaces (Fig. 3B).
Purification of Fc Fusion Proteins—Three days after trans-
fecting COS-1 cells with expression vectors for Fc fusion pro-
teins, the culture supernatant was recovered, and Fc fusion pro-
teins were purified by affinity chromatography as described
previously (23). The protein concentrations of Fc fusion prep-
arations were measured using the Bradford method and purity
assessed by SDS-PAGE/Coomassie Brilliant Blue staining (sin-
gle band) and by Western blotting (reactivity for anti-dectin-2
Ab).
Binding Assays for Microbes—Aliquots of freshly cultured
bacteria (0.1 OD
600
), S. cerevisiae and C. albicans yeasts (0.5–
1 10
6
cells each), or hyphae (4 10
5
cells) were washed with
Dulbecco’s PBS (DPBS) and incubated with staining buffer
(0.1% BSA, 2 m
M CaCl
2
, DPBS) containing 20
g/ml Fc proteins
on ice for 1 h. After extensive washing with buffer, cells were
resuspended in 5
g/ml of biotinylated goat anti-human IgG
F(ab)
2
Ab (Jackson ImmunoResearch) on ice for 30 min, fol
-
lowed by incubation with 1:200-diluted FITC-avidin (Vector
Laboratories Inc., Burlingame, CA). We also stained filamen-
tous fungi (M. audouinii, and T. rubrum) as follows. Single col-
onies of fungi were grown on Sabouroud’s agar plates, har-
vested, and suspended in DPBS. After washing with DPBS and
with water, small aliquots were spotted on slide glass, air-dried,
and stained with Fc proteins as before. Binding of Fc proteins to
microbes was examined using a Zeiss LSM510 laser scanning
confocal microscope with 488 nm excitation and transmitted
light detection (Carl Zeiss Microimaging, Thornwood, NY).
Quantitative Binding Assays—Fc protein (20 or 40
g) was
iodinated with 200 or 400
Ci of Na
125
I (ICN Biomedicals,
Aurora, OH) at room temperature for 10 min in the presence of
a rehydrated IODO-BEADs (Pierce). The reaction was stopped
by removing the beads and diluting with 0.1% BSA/DPBS, fol-
lowed by dialysis with CaCl
2
/DPBS until background levels of
radioactivity were detected in the dialysis buffer. Radioactivity
incorporated into Fc protein was measured by
125
I cpm in the
trichloroacetic acid-insoluble fraction. Specific activity was
expressed as incorporated cpm/total input/
g (typically 1–2
10
6
cpm/
g).
Iodinated Fc proteins were used to quantitate binding of Fc
proteins to C. albicans. Freshly cultured yeasts (5 10
5
cells) or
hyphae (4 10
5
cells) were incubated with different doses of
125
I-labeled Fc protein on ice for 1 h (two sets in triplicate).
After extensive washing with CaCl
2
/DPBS, one set was left
untreated, air-dried, and measured for radioactivity bound to
C. albicans using a
-counter. The other set was incubated with
acidic buffer (0.15
M NaCl, 0.1 M glycine-HCl buffer, pH 2.3) or
10 m
M EDTA (for Ca
2
-dependent binding) on ice for 5 min,
followed by washing. Residual radioactivity was regarded as
background. Specific binding was expressed as the counts/min
left after subtracting average background counts/min from
untreated counts/min. The amount of Fc proteins bound to
C. albicans was calculated as specific binding cpm/specific
activity of
125
I-Fc protein.
For experiments measuring specific binding of Dec2-Fc to
hyphae, hyphae (2 10
5
cells) were pretreated with various
concentrations of Fc protein on ice for 1 h (triplicate). After
removing unbound Fc proteins by washing, 1
g/ml of
125
I-
Dec2-Fc was added to pretreated and untreated hyphae and
then incubated on ice for another 1 h. A set of tubes was incu-
bated with
125
I-Dec2-Fc in the presence of 10 mM EDTA, and
Ca
2
-dependent binding activity was calculated as before. The
ability of polysaccharide to inhibit Dec1-Fc or Dec2-Fc binding
to hyphae or yeasts was assayed as follows: yeast or hyphae (1
10
6
or 2 10
5
cells/ELISA well) were washed with 0.1% BSA/
DPBS/CaCl
2
and incubated with
125
I-Dec1-Fc or
125
I-Dec2-Fc
(1
g/ml) in the presence of laminarin or mannan (both from
Sigma) on ice for 1 h. After extensive washing, C. albicans-
bound and background radioactivities were measured as
before.
Binding of Transfectants to C. albicans—The following pro-
cedures were followed for binding of COS-1 transfectants to
C. albicans hyphae. A day after transfecting COS-1 cells with
expression vectors for full-length dectin-1-V5 or dectin-2-V5,
or an empty vector, cells were re-seeded on 60-mm culture
dishes (5 10
5
cells/dish) and metabolically labeled with
[
3
H]thymidine (ICN Biochemicals, 1
Ci/dish) for 16 h. Cells
were then harvested by pipetting in 0.02% EDTA/DPBS. After
washing with 10% FCS/RPMI (cRPMI), specific activity of
labeled cells (cpm/cell) was determined. Cells in increasing
numbers were added to hyphae grown in 96-well plates (2 10
5
cells/well, in triplicate) and cultured in a CO
2
incubator at 37 °C
for 1 h. Amphotericin B (Sigma) was added to block fungal
growth (final concentration of 2.5
g/ml). Unbound COS-1
cells were removed by washing with cRPMI 10 times; cells
bound to hyphae were lysed by incubation with 0.3% Triton
X-100/PBS (200
l/well) at room temperature for 20 min.
For binding of RAW cells to C. albicans hyphae, the RAW
parental cells or those expressing dectin-1-V5 or dectin-2-V5
were metabolically labeled with [
3
H]thymidine (1
Ci/culture)
by overnight incubation. After measuring specific radioactivity
(cpm/cell), labeled cells (3 10
4
cells/well) were incubated in
ELISA wells just treated with 0.1% BSA/PBS or where hyphae
were grown (10
4
cells/well). After culturing at 37 °C for 30 min,
wells were washed with 0.1% BSA/PBS 10 times and lysed with
100
l of 0.3% Triton X-100/PBS, and
3
H counts were deter
-
mined. The number of cells adherent to a well was computed by
dividing
3
H counts/min from a well by specific activity.
For binding of RAW cells to C. albicans yeasts (26), freshly
grown yeasts were washed twice with PBS and resuspended in
0.1 mg/ml FITC (Sigma) at room temperature for 1 h. After
extensive washing, FITC-labeled yeasts were resuspended in
10% FCS-HBSS. RAW cells (5 10
5
) were incubated with
FITC-labeled yeasts at indicated m.o.i. values for 30 min at
room temperature. After removing unbound yeasts by exten-
sive washing, cells were fixed with 1% paraformaldehyde for 1 h
at 4 °C, washed, and then analyzed using FACSCalibur (BD Bio-
sciences). Histograms were made from fluorescent signals after
Recognition of Hyphae by Dectin-2
38856 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
removal of free FITC-yeasts by gating out the small sized pop-
ulation using forward/side scatter analysis.
Immunoprecipitation and Western Blotting—To measure
protein expression of dectin molecules in RAW cells, whole cell
extracts were prepared from cells by lysis in RIPA buffer (0.05
M
Tris-HCl, pH 7.5, 0.15 M NaCl, 1% Triton X-100, 1% sodium
deoxycholate, 0.1% SDS, 20 m
M EDTA) and subsequent centrif-
ugation at 16,000 g for 20 min at 4 °C. Small aliquots were
applied to 4 –20% SDS-PAGE, then transferred to a polyvinyli-
dene difluoride membrane (Hybond P; Amersham Bio-
sciences), followed by immunoblotting using mouse anti-V5
Ab (0.5
g/ml), affinity-purified rabbit anti-dectin-1 oligopep-
tide (1
g/ml) (19), or rat anti-dectin-2 mAb (0.5
g/ml) (20)
diluted with TTBS (20 m
M Tris-HCl, pH 7.6, 137 mM NaCl,
0.1% Tween 20). After washing, the membrane was blotted fur-
ther with horseradish peroxidase-conjugated secondary Ab and
then developed using the ECL Plus system (Amersham
Biosciences).
For protein tyrosine phosphorylation, RAW cells (2.5 10
6
)
were starved by culturing for1hinserum-free DMEM, incu-
bated at 37 °C for 1 h, and cocultured with yeast or hyphae
(7.5 10
6
each) in 24-well plates. At
different time points after incuba-
tion at 37 °C, cells were chilled on
ice and lysed by addition of 10
lysis buffer (20 m
M Tris-HCl, pH
7.6, 10% Triton X-100, 10 m
M
sodium orthovanadate, 10 mM
EDTA) to terminate phosphoryla-
tion. The clear lysate was prepared
by centrifugation at 14,000 rpm for
20 min and subjected to Western
blot analysis using 1:1,000-diluted
horseradish peroxidase anti-phos-
photyrosine Ab (PY-plus, Zymed
Laboratories Inc.).
To examine association of dec-
tin-2 with the FcR
chain, whole cell
extracts were prepared from
Dec2V5-RAW or parental macro-
phages (1 10
6
cells) using a lysis
buffer (1% Brij 55, 50 m
M Tris-HCl,
pH 7.6, 1 m
M Na
2
VO
4
,50mM NaF,
proteinase inhibitor mixture (Sigma))
and incubated with mouse anti-V5 (2
g) or mouse anti-human FcR
chain 7D3.5 mAb (Note: the mAb
we originally developed has cross-
reactivity to mouse FcR
)(3
g) at
4 °C for 16 h, followed by precipita-
tion with 10
l of 50% slurry protein
G-agarose (Roche Applied Science).
After washing the agarose beads, the
immunoprecipitates were dissoci-
ated from the beads by boiling and
then subjected to Western blotting
using anti-FcR
Ab or rat anti-dec-
tin-2 mAb (each 2
g/ml). The
interaction was also examined in COS-1 cells (1 10
6
cells)
cotransfected with two expression vectors encoding for dectin-
2-V5 and FcR
(pcDNA-m
chain), respectively.
To measure phosphorylation of the FcR
chain, Dec2V5-
RAW or RAW parental cells (1 10
6
cells in 100
l of PBS)
were incubated with anti-V5 Ab (5
g/ml) at 4 °C for 40 min.
After extensive washing, cells were treated with goat anti-
mouse IgG (20
g/ml) at 37 °C at various time periods and lysed
using 100
lof2 lysis buffer (1% Triton X-100, 50 mM Tris-
HCl, pH 7.6, 1 m
M Na
2
VO
4
,50mM NaF, proteinase inhibitor
mix (Sigma)). In some experiments, RAW cells were pretreated
with PP2 or PP3 kinase inhibitor at 37 °C for 2 h. Protein
extracts were prepared, immunoprecipitated with anti-FcR
chain Ab, and then blotted with anti-phosphotyrosine Ab 4G10
(1
g/ml) (Upstate Cell Signaling Solutions, Lake Placid, NY) or
anti-FcR
chain Ab. RAW cells (2.5 10
6
) were also treated
with C. albicans hyphae or yeasts (3 10
6
) at 37 °C at different
time periods. Tyrosine phosphorylation was examined as
described previously.
Immunofluorescence Staining—Binding of Dec2V5-RAW
cells to hyphae was also studied using microscopy. Hyphae (3
FIGURE 1. Soluble dectin-2 receptor binds to the cell wall of filamentous fungi. A, freshly cultured C.
albicans yeasts (5 10
5
cells) and pseudohyphae (4 10
5
cells) were incubated separately with Dec2-Fc or Fc
(20
g/ml), followed by staining with biotinylated anti-human IgG Ab and FITC-streptavidin. Morphology
under light microscopy (Phase) is included to distinguish yeast/conidial (round) from hyphal (filamentous)
forms. Fluorescence microscopy (FITC) was used to identify binding of Fc proteins. A black line is a 10-
m scale
bar. B, the dermatophytes, M. audouinii and T. rubrum, were each incubated with Fc, Dec1-Fc, or Dec2-Fc (each
20
g/ml) and then stained for immunofluorescence.
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38857
by guest on December 26, 2015http://www.jbc.org/Downloaded from
10
5
cells/well) grown in 2-well chamber slides (Lab-Tek Prod
-
ucts, Naperville, IL) were labeled with 100
g/ml TRITC
(Sigma) in 0.1
M sodium bicarbonate, pH 8.3, at room temper-
ature for 30 min. Free TRITC was removed completely by
extensive washing with PBS. Dec2V5-RAW cells (5 10
6
cells/
ml) were pretreated with 1% mouse serum (Jackson Immu-
noResearch) in 10% FCS/HBSS on ice for 10 min and surface-
labeled with 2
g/ml FITC-anti-V5 Ab on ice for 1 h. After
washing three times with 10% FCS/HBSS, surface-labeled
RAW cells (3 10
5
cells/ml) were resuspended in complete
DMEM containing 2.5
g/ml amphotericin B and then cul-
tured with TRITC-labeled hyphae (3 10
5
RAW cells/well). At
different time points after incubating at 37 °C, media were
removed, and the cells fixed immediately with 10% formalde-
hyde/PBS. Finally, immunofluorescence microscopy was per-
formed at the Live Cell Imaging Facility at the University of
Texas Southwestern Medical School. Fluorescence images
were taken under confocal microscopy and analyzed using 488
nm excitation for FITC and 543 nm for TRITC.
To determine subcellular localization of dectin-2 and FcR
,
RAW cells (5 10
5
) were incubated with polyclonal rabbit
anti-V5 Ab (10
g/ml) (Chemicon International, Temecula,
CA) at 4 °C for 30 min. After washing with PBS, cells were fixed
with 4% paraformaldehyde/PBS for 20 min at room tempera-
ture, cytospun to a slide glass, and permeabilized with 0.2%
Triton X-100/PBS for 2 min. The slide glass was incubated with
mouse anti-FcR
chain Ab (1
g/ml)
at room temperature for1hand
stained with Alexa488-conjugated
goat anti-mouse or 594-conjugated
goat anti-rabbit IgG (each 1:1,000
dilution) (Molecular Probes). Fluo-
rescence images were taken under
confocal microscopy using 488 nm
(for FcR
) or 594 nm excitation (for
dectin-2). In the case of COS-1 cells,
cells (1 10
4
) were seeded on a cov
-
erslip (12 mm diameter) in 24-well
plates. Two days after transfection,
cells were treated and analyzed in a
similar manner.
Internalization and Its Inhibition
by Tyrosine Kinase Inhibitor
Dec2V5-RAW cells were seeded on a
2-well chamber slide (LabTek) (1
10
5
cells/well) and cultured over
-
night. After washing cell layers once
with PBS, RAW cells were pre-
treated with 2.5
g/ml Fc block in
10% FCS/PBS on ice for 10 min and
processed for surface labeling with
2
g/ml FITC-anti-V5 or isotypic
control Ab (Invitrogen). After elim-
inating unbound Ab, FITC-labeled
dectin-2 was cross-linked with 10
g/ml anti-mouse IgG F(ab)
2
(Jackson ImmunoResearch) on ice
for 30 min, washed, and labeled with
200 n
M LysoTracker Red (Molecular Probes) for1hat37°C.
Cells were washed three times with 1% FCS/PBS and fixed with
10% formaldehyde. Optical sections were acquired using a
Leica TCS SP1 laser scanning confocal microscope (Leica
Micro-systems, Bannockburn, IL) as described previously.
For inhibition of endocytosis (27), RAW cells (1 10
6
) were
pretreated with fresh complete DMEM containing 0.5% Me
2
SO
(control) or the indicated concentrations of PP2 or PP3 (Cal-
biochem) at 37 °C for 30 min. After removing medium, cells
were incubated with 2
g/ml FITC-anti-V5 Ab (Invitrogen) on
ice for 1 h, followed by staining with 20
g/ml goat anti-mouse
IgG F(ab)
2
. After surface labeling with the Ab, cells were
allowed to internalize cross-linked Ab by incubating at 37 °C for
1 h in the continuous presence of an inhibitor at the same con-
centration. Treated cell samples were then examined for FITC
intensity by flow cytometry before and after treating with 0.2%
trypan blue/PBS for 1–2 min to quench the surface FITC. The
value of internalized FITC was computed by subtracting back-
ground fluorescence (surface staining of cells treated with
trypan blue but without incubation) from mean fluorescence of
cells treated with trypan blue. Finally, the effect of a tyrosine
kinase inhibitor on internalization was evaluated by the inter-
nalization value of a sample treated with an inhibitor relative to
untreated control (set at 100%).
Electromobility Shift Assay (EMSA)—RAW cells (3 10
7
cells/dish) were infected with C. albicans yeast or hyphae at a
FIGURE 2. Quantitative analyses of dectin-2 binding to C. albicans hyphae versus yeast. A, binding to C.
albicans. At increasing doses (nanograms of protein/well),
125
I-Dec2-Fc was incubated with 4 10
5
C. albicans
pseudohyphae (closed circles) or yeast (open circles) (triplicate sets for each dose point). Protein bound to C.
albicans was calculated as specific cpm/specific activity of
125
I-Fc protein. B, Ca
2
-dependence.
125
I-Labeled
Dec2-Fc (100 ng/well) was allowed to bind to C. albicans hyphae (4 10
5
) untreated (None) or treated with acid
(gray bar) or EDTA (black bar). After washing, cell-bound radioactivity was measured and binding expressed as
% of control. C, saturation curve of Dec2-Fc binding to hyphae.
125
I-Labeled Dec2-Fc (
g/ml) was incubated
with hyphae (2 10
5
) and Ca
2
-dependent binding expressed as hyphae-bound protein (nanograms).
D, yeasts bind poorly to Dec2-Fc. Hyphae or yeasts in increasing numbers were incubated with a constant
amount of
125
I-Dec2-Fc (100 ng/well), and Ca
2
-dependent binding was examined. E, hyphae (2 10
5
) were
pretreated with indicated concentrations (
g/ml) of cold Dec2-Fc or Fc before assaying binding to hyphae
using 1
g/ml
125
I-Dec2-Fc. F, Dec2-Fc binding is blocked by mannan. Hyphae (3 10
5
) or yeasts (1 10
6
) were
incubated with
125
I-Dec1-Fc or
125
I-Dec2-Fc (1
g/ml), respectively, in the absence (None) or presence of
laminarin (Lam) or mannan (Man). Relative binding (%) to control (no saccharide added) is shown. All data are
representative of at least three independent experiments.
Recognition of Hyphae by Dectin-2
38858 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
m.o.i. of 3. After incubating at 37 °C for 1 h, cells were washed
twice with ice-cold PBS and then lysed by incubating in ice-cold
0.6% Nonidet P-40-containing buffer A (10 m
M HEPES, pH 7.9,
10 m
M KCl, 0.1 mM EDTA, 1 mM EGTA, 1 mM phenylmethyl-
sulfonyl fluoride, 10
g/ml leupeptin, 1 mM dithiothreitol) for 5
min. Cell lysates were collected and centrifuged at 14,000 rpm
at 4 °C for 20 s. The supernatant containing cytosolic proteins
was aspirated, and pellets were resuspended in 1 ml of ice-cold
buffer A without Nonidet P-40. Following centrifugation at
14,000 rpm at 4 °C for 20 s, the pellet was resuspended in 50
l
of ice-cold buffer C (20 m
M HEPES, pH 7.9, 0.4 M NaCl, 1 mM
EDTA, 1 mM EGTA, 1 mM phenylmethylsulfonyl fluoride, 10
g/ml leupeptin, 1 mM dithiothreitol) and incubated on ice for
1 h. Clear lysates (nuclear extracts) were prepared by centrifu-
gation at 14,000 rpm for 30 min. Protein concentration was
determined by the Bradford method (normally, 2– 4 mg/ml),
snap-frozen in liquid nitrogen, and stored at 85 °C until
needed. Activation of NF-
B was examined by EMSA using an
aliquot (4
g) of prepared nuclear extract and a gel-shift assay
kit (Promega, Madison, WI).
Cytokine Expression Analysis—Cytokine gene expression by
RAW cells was examined using RNase protection assay per-
formed according to the manufacturer’s recommended proto-
cols (RiboQuant multiprobe ribonuclease protection assay sys-
tem, Pharmingen). Briefly,
32
P-labeled RNA probes were
generated using the mCK-2b multiprobe template set and
FIGURE 3. Dectin-2 on RAW cells recognizes C. albicans hyphae. A, protein expression of dectin-1 and dectin-2 in RAW macrophage transfectants.
Whole cell extracts were prepared from parental RAW macrophages (M) or those transfected with dectin-1 or dectin-2 tagged with the C-terminal V5
epitope (Dec1V5 or Dec2V5). Aliquots (each 1 10
5
cells eq) were examined for protein expression of Dec1V5 or Dec2V5 by Western blotting using
anti-V5 Ab. Two arrows (31 and 42 kDa) indicate bands corresponding to Dec1V5 and Dec2V5, respectively. B, RAW cells were pretreated with Fc block
before staining with FITC/anti-V5 Ab. Surface expression of Dec1V5 (open histogram) or Dec2V5 (closed histogram) was assayed by flow cytometry and
compared with parental cells (dashed lines). C, binding to yeast. RAW cells were cocultured with fluorescence-labeled live yeasts at varying m.o.i. values.
Frequency (%) of fluorescently labeled RAW cells was determined by flow cytometry. D, binding to hyphae. RAW cells metabolically labeled with
[
3
H]thymidine were incubated in the 96-well plate covered with/without hyphae (4 10
5
cells/well) at 37 °C for 30 min. After washing, cells were lysed
and measured for
3
H counts, and cell numbers adhered to a well were computed. Binding of RAW cells to hyphae is expressed as fold difference (hyphae
versus culture well). E, COS-1 cells were transfected with an empty vector (open triangles) or expression vector for dectin-1-V5 (open circles) or dectin-
2-V5 (closed circles). [
3
H]Thymidine-labeled COS-1 cells at increasing cell numbers were incubated with a constant number of hyphae (2 10
5
cells/well
in triplicate). After washing, hyphae-adhered COS-1 cells were lysed and measured for
3
H counts/min. The number of cells that bound to hyphae was
calculated as well bound cpm/specific activity of
3
H-labeled cells (0.24 0.32 cpm/cell). F, dectin-2 protein on RAW cells was fluorescently labeled with
FITC/anti-V5 Ab and then cocultured with pseudohyphae labeled with rhodamine (red fluorescence). Green (anti-V5 Ab) and red (hyphae) fluorescence
images were separately taken and merged (Merge). Surface distribution of dectin-2 before coculturing with hyphae was also shown (Before). All data
shown are representative of three independent experiments.
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38859
by guest on December 26, 2015http://www.jbc.org/Downloaded from
mixed with 20
g of total RNA isolated from RAW cells
infected with/without yeasts or hyphae at a m.o.i. of 0.3. RAW
cells (10
7
) were incubated with yeasts or hyphae (3.3 10
6
)in
complete DMEM containing 2.5
g/ml amphotericin B. After
culturing for 3 h, total RNA was isolated from treated cells using
RNA-STAT60 (Tel-Test B, Friendswood, TX) (28), hybridized
with probe, and then digested with RNase. mRNA-protected
probes were size-fractionated on 8
M urea, 5% polyacrylamide
gel, and radioactivity was measured using a PhosphorImager
analyzer, STORM820 (Amersham Biosciences).
To measure secretion of cytokines (29), parental or Dec2V5-
RAW cells (set in triplicate) were cocultured using previous
settings but for longer time periods (6 and 16 h). Protein
amount of cytokine secreted in the culture media was measured
using respective ELISA kits as follows: IL-1ra kit purchased
fromR&DSystems (Minneapolis, MN) and other cytokines
from eBioscience (San Diego, CA).
RESULTS
Dectin-2 Binds Hyphal Components of C. albicans in a
Calcium-dependent Manner—Because dectin-1 is a PRR for
-glucan on yeasts (30), we questioned whether dectin-2 also
recognized microbial organisms. We created soluble receptors
of dectin-2 and dectin-1, in which the respective extracellular
domain was fused to the Fc portion of human IgG
1
(Dec2-Fc;
Dec1-Fc). We then performed binding assays to assess the abil-
ity of fluorescence-labeled Dec2-Fc, Dec1-Fc, or Fc alone (con-
trol) to recognize microbes. None of the probes bound to S.
aureus, group A streptococci, P. aeruginosa,orE. coli (data not
shown). As reported previously (26), Dec1-Fc bound to C. albi-
cans yeasts especially at budding sites (data not shown). By con-
trast, Dec2-Fc bound to hyphal (but not yeast) components of C.
albicans (Fig. 1A). We next questioned whether differences in the
ability of dectin-1 and dectin-2 to recognize yeast versus hyphal
forms extended to other fungi (Fig. 1B). Dec2-Fc bound to the
filamentous (hyphal) but not conidial (yeast) form of the dermato-
phytes, M. audouinii and T. rubrum, whereas Dec1-Fc bound
preferentially or predominantly to the conidial form (Fig. 1B).
To quantify binding activity, Candida yeast or hyphae were
incubated with
125
I-labeled Dec2-Fc in increasing doses (Fig. 2
).
After washing, Candida-bound
125
I radioactivity (counts/min)
was determined and nonspecific binding regarded as radioac-
tivity left after treatment with acid buffer. Specific binding was
expressed as counts/min after subtracting nonspecific binding
from untreated counts/min. Using Candida-bound Dec2-Fc pro-
tein, calculated from specific activity of
125
I-Fc proteins (cpm/
g),
we observed binding of dectin-2 to hyphal components in a dose-
dependent manner, whereas binding to yeast components was
minimal, even at the highest dose tested (Fig. 2A).
Because dectin-2 contains an EPN motif required for Ca
2
-
dependent carbohydrate binding by C-type lectins (31), we
examined the effect of the calcium inhibitor, EDTA, on binding
of Dec2-Fc to C. albicans hyphae (Fig. 2B). EDTA treatment
(10 m
M) abrogated such binding as strongly as did acid treat-
ment. Moreover, incubation of a constant number of hyphae
with increasing doses of
125
I-Dec2-Fc in the presence of cal
-
cium revealed saturation of binding at a range of 30 –100
g/ml
(Fig. 2C). These results suggest that putative ligands of dectin-2
are expressed abundantly on hyphae.
Because hyphae are larger than yeasts, we controlled for fun-
gal size by culturing C. albicans yeast or hyphae (increasing
numbers) with
125
I-Dec2-Fc (constant dose) (Fig. 2D). At a dose
range of less than 1 10
6
cells, hyphae bound Dec2-Fc in a
dose-dependent manner. By contrast, yeast bound to Dec2-Fc
only minimally, if at all (Fig. 2D). To more rigorously evaluate
specificity of Dec2-Fc binding to hyphae, we saturated putative
ligands for dectin-2 on hyphae by pretreatment with cold
Dec2-Fc or Fc control at increasing doses before measuring
binding of
125
I-labeled Dec2-Fc (Fig. 2E). Pretreatment with
Dec2-Fc, but not Fc control, blocked binding in a dose-depend-
ent manner, up to 80% at the highest dose tested (100
g/ml) in
which putative ligands of dectin-2 were presumed to be satu-
rated with cold Dec2-Fc (Fig. 2C).
Because
-glucan is a ligand of dectin-1, we examined
whether dectin-2 also recognizes
-glucan or its structurally
related polysaccharide. Consistent with a previous report (26),
laminarin almost completely blocked binding of Dec1-Fc to
FIGURE 4. Hyphae trigger protein tyrosine phosphorylation in dectin-2-
expressing RAW cells. RAW cells were incubated with medium alone (None)
or with C. albicans yeast or hyphae for various times (A). Cells were also treated
for 30 min with medium (None), control IgG (Ctrl Ig), or anti-V5 Ab plus sec-
ondary Ab (cross-linking) (B). Whole cell extracts were prepared, and expres-
sion of tyrosine-phosphorylated proteins was assayed by Western blotting
using horseradish peroxidase-coupled anti-phosphotyrosine mAb. The data
are representative of two experiments.
Recognition of Hyphae by Dectin-2
38860 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
yeast but had almost no effect on binding of Dec2-Fc to hyphae
(Fig. 2F). By contrast, mannan, a polysaccharide purified from
S. cerevisiae, blocked binding of Dec2-Fc to hyphae in a dose-
dependent manner while only minimally blocking binding of
Dec1-Fc to yeast (Fig. 2F). These results indicate that dectin-2
and dectin-1 have disparate ligands.
To confirm that full-length dectin-2 (as expressed on APC
surfaces) can recognize hyphae, we transfected RAW264.7
macrophages with an expression vector that encodes dectin-2
or dectin-1 tagged with a C-terminal V5 epitope (Dec1V5 or
Dec2V5, respectively). Note that parental RAW cells constitu-
tively express dectin-2 and dectin-1 but at markedly lower lev-
els compared with bone marrow-derived DC or the XS52 DC
line (19, 20). Western blotting of Dec1V5 and Dec2V5 proteins
was performed using anti-V5 antibody (Ab) to label dectin-2
(31 kDa) or dectin-1 (42 kDa) (Fig. 3A), and their identities were
also confirmed by immunoreactivity to anti-dectin-2 and anti-
dectin-1 Ab, respectively (data not shown). FACS analysis using
anti-V5 Ab revealed cell surface expression of Dec1V5 or
Dec2V5 at similar high levels (Fig. 3B). We next incubated
transfected RAW cells with FITC-labeled C. albicans yeast at
three different m.o.i. values and measured the percentage of
FITC-labeled RAW cells (Fig. 3C). As reported previously (16),
Dec1V5-RAW cells bound yeast to a greater degree than did
parental cells or Dec2V5-RAW cells (Fig. 3C). Because single
hyphae suspensions are hard to prepare, we measured hyphal
binding based on the ability of RAW cells to adhere to hyphae
stuck to the bottom of culture wells (Fig. 3D). RAW cells were
labeled with [
3
H]thymidine and incubated with/without
hyphae (constant number). The numbers of cells bound to
hyphae versus culture wells were determined (hyphae-bound
3
H cpm/specific activity of
3
H-labeled cells), and fold difference
was calculated. Dec2V5-RAW cells bound hyphae a level 100-
fold greater than did parental cells or Dec1V5-RAW cells (Fig.
3D). To exclude the possibility that adhesiveness of Dec1V5-
RAW cells to culture wells may have
masked binding to hyphae, we
transfected COS-1 cells with
expression vectors for Dec2V5,
Dec1V5, or control and determined
their ability to bind hyphae (Fig. 3E).
Prior to binding assays, we con-
firmed expression of Dec2V5 and
Dec1V5 proteins by Western blot-
ting and FACS. Surface expression
levels were much lower levels than
RAW transfectants (data not shown).
Dec2V5-COS-1 cells bound hyphae
markedly and in a dose-dependent
manner, whereas Dec1V5-COS-1
cells displayed minimal binding (Fig.
3E). Finally, we performed confocal
microscopy of FITC-anti-V5 Ab-
treated Dec2V5-RAW cells incubat-
ed with rhodamine-labeled C. albi-
cans pseudohyphae that contain yeast
and hyphal components (Fig. 3F).
Before incubating with Candida, dec-
tin-2 was distributed evenly on the cell surface. After coculture,
many RAW cells bound to hyphal components to the point of even
engulfing these fungal parts (Fig. 3F). By contrast, we did not
observe Dec2V5-RAW cells to bind yeast components. These
results also indicate that dectin-2 preferentially recognizes hyphal
rather than yeast components of C. albicans.
Binding of Hyphae to Dectin-2 Leads to Protein Tyrosine
Phosphorylation—Because dectin-2 lacks a tyrosine-based sig-
nal motif in its intracellular domain, we questioned whether
ligand-bound dectin-2 receptor was capable of transducing
intracellular signals. We subjected whole cell extracts of RAW
cells cultured with C. albicans yeast or hyphae to Western blot-
ting using anti-phosphotyrosine Ab to detect tyrosine-phos-
phorylated proteins (Fig. 4). Compared with correspondingly
treated parental RAW cells, hyphae (but not yeast)-treated
Dec2V5-RAW cells yielded increased amounts of tyrosine-
phosphorylated proteins as early as 10 min after incubation
(Fig. 4A). To evaluate specificity for dectin-2, we cross-linked
dectin-2 with anti-V5 Ab plus secondary Ab (Fig. 4B); this treat-
ment also induced tyrosine phosphorylation, albeit to a lesser
degree than was achieved by hyphae. These results indicate that
ligation of dectin-2 can transduce tyrosine-based signals in the
absence of an intracellular signal motif.
Dectin-2 Associates with the Fc Receptor
Chain—DCAR is a
C-type lectin shown recently to associate with the FcR
chain
via an arginine in its transmembrane domain (32). Because dec-
tin-2 shows 96% amino acid identity to the transmembrane of
DCAR (25 of 26 amino acids, including the arginine connector
(32)), we posited that dectin-2 also associates with FcR
.We
used anti-V5 Ab to immunoprecipitate Dec2V5 protein from
extracts of Dec2V5-RAW (Fig. 5A) and then blotted it with
anti-dectin-2 or anti-FcR
Ab. We found dectin-2 and FcR
proteins in precipitates from anti-V5 Ab (but not control Ab)-
treated Dec2V5-RAW cells (Fig. 5A); dectin-2 was not detected
in precipitates from RAW parental cells. We also used reverse
FIGURE 5. Dectin-2 associates with FcR
chain. A, Dec2V5 protein was immunoprecipitated (IP) from whole
cell extracts of parental RAW or Dec2V5-RAW cells using anti-V5 or control IgG (Ctrl IgG). The precipitates were
then immunoblotted (IB) with anti-FcR
or anti-dectin-2 (Dec2) mAb. Reverse precipitation was also per-
formed. B, COS-1 cells were cotransfected with expression vectors for Dec2V5 or FcR
and then subjected
similarly to immunoprecipitation analysis. C, localization of Dec2V5 and FcR
proteins. Dec2V5 protein on
parental or Dec2V5-RAW cells or COS-1 cells cotransfected previously was surface-labeled with rabbit anti-V5
plus Alexa594-conjugated anti-rabbit Ab (shown in red fluorescence), fixed, permeabilized, and stained with
mouse anti-FcR
plus Alexa488-anti-mouse Ab (green). Colocalization was examined using confocal
microscopy.
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38861
by guest on December 26, 2015http://www.jbc.org/Downloaded from
immunoprecipitation to show that anti-FcR
Ab coprecipi-
tated Dec2V5 protein. In addition, we employed COS-1 cells
cotransfected with Dec2V5 and FcR
genes to confirm that
dectin-2 associates with FcR
(Fig. 5B). Finally, we used confocal
microscopic analysis to locate dectin-2 and FcR
proteins within
Dec2V5-RAW and cotransfected
COS-1 cells (Fig. 5C). These cells
were surface-labeled with phyco-
erythrin-anti-V5 Ab (Fig. 5C, red flu-
orescence), fixed, and then stained
with FITC-anti-FcR
Ab (Fig. 5C,
green). In RAW cells, the majority of
endogenous FcR
protein resided on
the cell surface colocalizing with sur-
face-labeled dectin-2 (Fig. 5C, yellow).
In cotransfected COS-1 cells, similar
colocalization was observed, although
the majority of FcR
resided intracel-
lularly (Fig. 5C).
To determine whether the trans-
membrane arginine (Arg-17) in
dectin-2 is required to associate
with the FcR
chain, we assayed the
binding of an R17V mutant, in which the positively charged
arginine was replaced by the neutrally charged valine (Fig. 6).
FcR
protein was immunoprecipitated from extracts of COS-1
cells cotransfected with the R17V mutant (tagged with the
C-terminal V5) and FcR
, and then immunoblotted with
anti-V5 to detect dectin-2 (Fig. 6B). Wild-type dectin-2 and the
R17V mutant each coprecipitated FcR
efficiently, indicating
that transmembrane arginine is not essential for the associa-
tion. We next examined the importance of the intracellular
and extracellular domains of dectin-2 by constructing three
other mutants as follows: ICD mutant with the entire intra-
cellular domain (amino acids 1–14) deleted; 1/2 ICD lack-
ing the N-terminal half of the intracellular domain (amino
acids 1–7); and 40LECD in which the extracellular domain is
replaced by the CD40 ligand (CD40L) (33), a type II trans-
membrane receptor that does not associate with FcR
.
Immunoprecipitation revealed binding of 1/2 ICD or
40LECD with FcR
as avidly as that of the wild type, whereas
ICD mutant bound poorly, indicating that a short stretch of
the intracellular domain of dectin-2 (amino acids 8 –14)
proximal to the transmembrane domain is required for asso-
ciating with FcR
.
Dectin-2 Transduces Tyrosine Phosphorylation of Fc Receptor
Chain—We next questioned whether ligation of dectin-2
leads to tyrosine phosphorylation of the FcR
chain (Fig. 7). At
different time points after cross-linking dectin-2 on RAW cells
with anti-V5 Ab or control IgG, whole cell extracts were pre-
pared from treated RAW cells; FcR
protein was immunopre-
cipitated, and tyrosine phosphorylation of FcR
was examined
by immunoblotting with anti-phosphotyrosine Ab (to detect
phosphorylation levels) or anti-FcR
Ab (to measure precipi-
tated FcR
). A single band immunoreactive to anti-phosphoty-
rosine Ab was detected as early as 2 min, and it peaked at 5 min
followed by a rapid decrement (Fig. 7A). The phosphorylated
form, which migrated slower than the unphosphorylated form
(34), was also detected in immunoblots with anti-FcR
(Fig.
6A). Absence of phosphorylation in parental cells treated with
anti-V5 Ab and in Dec2V5-RAW cells treated with control Ab
confirmed specificity for dectin-2. We next determined
whether ligation of dectin-2 by hyphae (versus yeast as control)
FIGURE 6. Intracellular region of dectin-2, proximal to the transmembrane, is required for the associa-
tion. A, amino acid structures of dectin-2 mutants are schematically depicted and aligned with the wild type
(WT), consisting of an intracellular (ICD), a transmembrane (TM), and an extracellular domain (ECD). An inverted
closed triangle represents the location of a point mutation (arginine to valine). B, COS-1 cells were transfected
with an expression vector coding for WT or a mutant with () or without () vector for FcR
chain. Two days
post-transfection, whole cell extracts were prepared and subjected to immunoprecipitation and immunoblot-
ting with the indicated Ab. Representative blotting data of two independent experiments are shown.
FIGURE 7. Ligation of Dec2V5 on RAW cells transduces phosphorylation
of FcR
. A, phosphorylation of FcR
by cross-linking. At different time points
after cross-linking of Dec2V5 on parental RAW or Dec2V5-RAW cells using
anti-V5 Ab or control IgG (Ctrl IgG), whole cell extracts were prepared and
FcR
protein immunoprecipitated. Levels of FcR
protein and of its phospho-
rylation were determined by immunoblotting with anti-FcR
Ab or anti-phos-
photyrosine (p-Tyr). B, phosphorylation by C. albicans. RAW cells were cocul-
tured without (No) or with C. albicans hyphae (Hy;4 10
5
) or yeast (Y;1 10
6
)
and examined by Western blotting for phosphorylated and for total FcR
protein. C, inhibition of phosphorylation by Src kinase inhibitor. RAW cells
were pretreated with PP2 (an inhibitor for Src family kinases) or PP3 (a control
derivative) (
M) prior to coculture with C. albicans. Hyphae-induced tyrosine
phosphorylation was determined as before. Two bands immunoreactive to
anti-FcR
Ab (indicated by arrows) represent the phosphorylated and
unphosphorylated forms.
Recognition of Hyphae by Dectin-2
38862 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
leads to FcR
phosphorylation (Fig. 7B). Rapid phosphorylation
was observed in hyphae (but not in yeast)-treated Dec2V5-
RAW cells.
Because FcR
is phosphorylated by the Src kinases, Lyn and
Fyn (35), we examined whether PP2, an inhibitor of Src kinases,
can block dectin-2-induced FcR
phosphorylation; the biolog-
ically inert derivative, PP3, was used as control. Pretreatment of
RAW cells with PP2 (but not PP3) blocked FcR
phospho-
rylation by 60% (Fig. 7C). Altogether, our results indicate that dec-
tin-2 can associate with FcR
and that
such association is likely to transduce
Src-dependent phosphorylation.
Ligation of Dectin-2 Triggers
Internalization Likely through Src
Family Kinases—We next used con-
focal microscopy to study internal-
ization and intracellular trafficking
after cross-linking of dectin-2 on
Dec2V5-RAW cells with FITC-an-
ti-V5 Ab (as a surrogate ligand) plus
a secondary Ab (Fig. 8A). As early as
15 min after cross-linking, ligand-
loaded dectin-2 was internalized
and formed endosomes, most of
which were not fused to lysosomes
(stained by LysoTracker). We then
examined whether the Src family
kinases are involved in the internal-
ization (Fig. 8B). PP2 (50
M) pre-
treatment blocked internalization
by 70%, whereas PP3 had little
effect. Thus, internalization of
ligated dectin-2 is achieved through
activation of Src family kinases.
Ligation of Dectin-2 Activates
NF-kB—NF-
B is a major transcription pathway utilized by
many immunoregulatory receptors to mediate their down-
stream biologic effects (36). Because FcR
can activate NF-
B,
we examined the effects of ligated dectin-2 on NF-
B activa-
tion using the EMSA on nuclear extracts from transfected or
parental RAW cells infected with C. albicans (hyphae versus
yeasts) or LPS as a nonspecific control. Parental RAW cells
failed to induce NF-
B activation beyond steady-state levels
(Fig. 9A). By contrast, Dec2V5-RAW cells activated NF-
B
in response to hyphal infection but not yeasts. Moreover,
specificity for hyphae-ligated dectin-2 was confirmed by the
finding of close to equal nuclear translocation of NF-
Bin
parental and Dec2V5 RAW cells treated with LPS (Fig. 9B).
Hyphae-bound Dectin-2 Up-regulates IL-1ra and TNF
Expression—To determine whether ligated dectin-2 stimulates
RAW cells to produce cytokines, we again cocultured parental
and Dec2V5-RAW cells with C. albicans hyphae or yeast, and
we examined cytokine gene expression by multiple RNase pro-
tection assay (Fig. 10, A and B). Among the cytokine genes
tested (TNF
was unintentionally not included), IL-1ra was
most markedly up-regulated, 7-fold increase in Dec2V5-RAW
cells treated with hyphae versus 2-fold increase induced by
yeast (Fig. 10, A and B). IL-6 and IL-18 gene expression was
up-regulated minimally in hyphae-treated cells. We next meas-
ured production of five cytokines by RAW cells at 6 and 16 h
after infection with C. albicans (Fig. 10C). Consistent with
mRNA results, hyphae induced considerable secretion of
IL-1ra protein, whereas yeast did so only minimally. Hyphae-
induced TNF
production was even more greatly induced. A
time course study revealed hyphae-induced augmentation as
early as 2 h for TNF
and 6 h for IL-1ra (Fig. 10, D and E).
FIGURE 8. Dectin-2 rapidly internalizes a surrogate ligand (anti-V5 Ab) through activation of Src kinases.
A, Dec2V5-RAW cells were surface-labeled with FITC-anti-V5 Ab (0 min) and then cross-linked with secondary
Ab. At various time points after incubation, cells were fixed and stained with Red-LysoTracker (red fluorescence).
Confocal images of doubly stained cells are shown. B, internalizing capacity of Dec2V5-RAW cells was quanti-
fied by FITC fluorescent intensity of internalized anti-V5 Ab in the absence (100%) or the presence of PP2 or PP3
at varying concentrations (
M). Data shown are representative of two (A) and three (B) experiments.
FIGURE 9. Hyphae stimulate NF-
B activation in dectin-2-expressing
RAW cells. Nuclear extracts (NE) were prepared from RAW cells treated
without (None) or with yeast or hyphae (A) or with LPS (1
g/ml) (B) and
assayed for NF-
B(A and B) activation by EMSA. Specific (NF-
B) and non-
specific (NS) bands are shown by arrows. Second experiment showed sim-
ilar results.
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38863
by guest on December 26, 2015http://www.jbc.org/Downloaded from
Finally, to determine whether Src kinase played a role, we
assayed the inhibitory effect of PP2 on TNF
and IL-1ra secre-
tion (Fig. 10F). PP2 blocked hyphae-induced TNF
production
completely, whereas it inhibited IL-1ra secretion by 80%. These
results indicate that hyphae-ligated dectin-2 stimulates RAW
cells to produce IL-1ra and TNF
likely through activation of Src
kinases.
DISCUSSION
We showed that dectin-2 (soluble
extracellular domain or full-length
form on RAW cells) preferentially
binds hyphal forms rather than
yeast forms of C. albicans, M.
audouinii, and T. rubrum. Dectin-2
ligated by hyphae or cross-linked by
Ab induces phosphorylation of pro-
tein tyrosine, internalizes a surro-
gate ligand, activates NF-
B, and
up-regulates expression of TNF
and IL-1ra. Transduction of these
events after recognition of hyphae is
achieved by coupling of dectin-2
with the signal adaptor, FcR
, which
bears an immunoreceptor tyrosine-
based activation motif (ITAM).
Because ITAM-dependent signal-
ing in leukocytes appears critical to
the differentiation, proliferation,
regulation, and survival of several
immune effector cells (27), we spec-
ulate that dectin-2 on APC contrib-
utes to the initiation and modula-
tion of anti-fungal immunity.
Selective binding of dectin-2 to
hyphae led us to screen carbohy-
drates unique to hyphae (not found
in yeasts) as candidates for the dec-
tin-2 ligand, including chitin (37,
38);
-(1–3)- and
-(1–2)-linked
glucans (39, 40); high molecular
weight mannoproteins like CaCYC3
(41, 42); other mannoproteins (43,
44); and other lipids (45). However,
neither chitin, its carbohydrate unit
(N-acetylglucosamine), nor any of
the glucans blocked binding of dec-
tin-2 to hyphae (data not shown).
We also tested simple hexose carbo-
hydrates for their ability to bind to
dectin-2 and found none to do so
significantly (data not shown).
Rather, we discovered that high
dose mannan blocks binding of dec-
tin-2 to hyphae, a result consistent
with those from a glycan array anal-
ysis that employed synthetic carbo-
hydrates to show that dectin-2 can recognize high mannose
structure (46).
However, several caveats prevent us from declaring mannan
as the dectin-2 ligand. Mannan is a polymer of mannose con-
sisting of various oligomannosides, and the dectin-2 ligand may
FIGURE 10. Infection of dectin-2-RAW cells with hyphae induces expression of IL-1ra and TNF
likely
through activation of Src kinases. A, total RNA was extracted from RAW cells after coculture with yeast or
hyphae, and mRNA expression of cytokine genes was determined by RNase protection assay using multi-
probes to cytokines indicated at left. B, changes in mRNA expression after coculture are indicated as fold
increase over levels in untreated RAW cells. C, production of cytokines by parental and Dec2V5-RAW cells at 6
or 16 h following treatment similar to C. albicans was measured by ELISA. Up-regulated expression for each
cytokine is assessed as fold increase to untreated. D and E, kinetics of C. albicans-induced expression of IL-1ra
and TNF
by RAW cells was determined. F, Dec2V5-RAW cells were pretreated with PP2 or PP3 at different
doses prior to infection with C. albicans, and then IL-1ra and TNF
secretion (ng/ml) was assayed. Data shown
are representative of two (RNase protection assay) and three (ELISA) independent experiments.
Recognition of Hyphae by Dectin-2
38864 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
be a minor oligomannoside of this polysaccharide preparation.
C. albicans yeasts and hyphae both contain mannan in their cell
walls, yet dectin-2 binds preferentially to the latter. Thus, it is
possible that one of the minor oligomannosides is synthesized
more abundantly by hyphae (versus yeast) or that its presence in
hyphae (versus yeast) is more accessible for binding to dectin-2.
For example, dectin-1 binds preferentially to yeasts at budding
sites, where
-glucan is more accessible (26). Transformation
of yeasts to pseudohyphae may alter the three-dimensional
structure of the cell wall (44) that better displays the putative
dectin-2 ligand.
Type II-configured CLR on APC can be sorted into the fol-
lowing two groups based on the presence/absence of signaling
motifs in the intracellular domain: CLR having the motif
include dectin-1 which carries a YXXL (an ITAM-like
sequence) (19, 47), and DCIR which has an immunoreceptor
tyrosine-based inhibitory motif (48). CLR without the motif
include DCAR and dectin-2. Recently, it has been reported that
DCAR associates with the FcR
chain, enabling it to induce
signals leading to Ca
2
influx (32). The same authors claimed
that dectin-2 was unable to couple with the FcR
in COS-1 cells
cotransfected with FcR
and dectin-2 genes (32). Our results
are at odds with this report; not only is dectin-2 capable of
binding with the FcR
chain (coprecipitation of endogenous or
genetically engineered FcR
from RAW cells or from cotrans-
fected COS-1 cells using anti-V5 Ab) (Fig. 5, A and B) but also
dectin-2 and FcR
chain colocalize within these cells (Fig. 5C).
Furthermore, ligation of dectin-2 by hyphae or V5-cross-linked
Ab induces phosphorylation of the FcR
chain (Fig. 7). Note
that both dectin-2 and DCAR possess transmembrane domains
with almost identical amino acid sequence (one miss-match
among 26 amino acids), including a positively charged arginine
residue essential for interaction of many Ig superfamily mem-
bers with the FcR
chain (49). In contrast to DCAR (32) and
other Ig-like receptors (49), the association of dectin-2 with
FcR
was achieved via the intracellular domain proximal to the
transmembrane and not through transmembrane arginine.
Relevant to this finding is platelet receptor GPVI, which was
also shown to associate with FcR
through its intracellular
domain (50).
To study the function of dectin-2 in innate immunity, we
used dectin-2-overexpressing RAW cells as a model of inflam-
matory macrophages and DC expressing high levels of dectin-2
(22). Expression levels by the RAW cells are likely to be more
abundant than levels physiologically expressed by those inflam-
matory cells. Thus, some of our data may not reflect precisely
the real significance of dectin-2 on DC. Recognition of patho-
gens by DC is not achieved by a single receptor. Rather, DC
employ concurrently multiple receptors. In this regard, we
speculate that inflammatory macrophages and DC employ dec-
tin-2 to recognize hyphae, with the dectin-2-induced down-
stream events we found contributing in part to overall changes
induced by DC.
Interaction between particular microbes and PRR on APC
leads to intracellular and secretory events that may govern
whether effector responses generated against infection are pro-
tective or promiscuous. Ligation of dectin-1 by zymosan (con-
taining
-glucan) led to phosphorylation of the ITAM-like
motif of dectin-1, activation of Syk tyrosine kinase, and up-reg-
ulated secretion of IL-2 and IL-10 (47). By contrast, we showed
that ligation of dectin-2 by hyphae led to phosphorylation of
FcR
and up-regulated secretion of TNF
and IL-1ra. This dis-
parity between cytokines produced by each pathway may
account at least partially for differences in the biologic outcome
of infection by dimorphic fungi, with yeast-dominant infections
fostering protective immunity and hyphae-dominant infec-
tions engendering greater tissue invasion.
Acknowledgments—We thank Didier Trono (Ecole Polytechnique
Fédérale de Lausanne (EPEL), Switzerland) and Yasuhiro Ikeda
(Mayo Clinic, Rochester, MN) for providing the lentiviral vectors,
pCMVR8.91, pMD-G, and pHR-SIN-CSGW dlNotI. We are also
grateful to Irene Dougherty for technical expertise and to Susan
Milberger for administrative assistance.
REFERENCES
1. Gordon, S. (2002) Cell 111, 927–930
2. Barton, G. M., and Medzhitov, R. (2002) Curr. Opin. Immunol. 14,
380–383
3. Gough, P. J., and Gordon, S. (2000) Microbes Infect. 2, 305–311
4. Janssens, S., and Beyaert, R. (2003) Clin. Microbiol. Rev. 16, 637–646
5. Schnare, M., Barton, G. M., Holt, A. C., Takeda, K., Akira, S., and Medzhi-
tov, R. (2001) Nat. Immun. 2, 947–950
6. Underhill, D. M. (2003) Eur. J. Immunol. 33, 1767–1775
7. Weis, W. I., Taylor, M. E., and Drickamer, K. (1998) Immunol. Rev. 163,
19–34
8. Stahl, P. D., and Ezekowitz, R. A. (1998) Curr. Opin. Immunol. 10, 50 –55
9. Jack, D. L., Klein, N. J., and Turner, M. W. (2001) Immunol. Rev. 180,
86–99
10. Engering, A., Geijtenbeek, T. B., Van Vliet, S. J., Wijers, M., van Liempt, E.,
Demaurex, N., Lanzavecchia, A., Fransen, J., Figdor, C. G., Piguet, V., and
van Kooyk, Y. (2002) J. Immunol. 168, 2118–2126
11. Geijtenbeek, T. B., Torensma, R., van Vliet, S. J., van Duijnhoven, G. C.,
Adema, G. J., van Kooyk, Y., and Figdor, C. G. (2000) Cell 100, 575–585
12. Geijtenbeek, T. B., Van Vliet, S. J., Koppel, E. A., Sanchez-Hernandez, M.,
Vandenbroucke-Grauls, C. M., Appelmelk, B., and van Kooyk, Y. (2003) J.
Exp. Med. 197, 7–17
13. Colmenares, M., Corbi, A. L., Turco, S. J., and Rivas, L. (2004) J. Immunol.
172, 1186 –1190
14. Geijtenbeek, T. B., Engering, A., and Van, K. Y. (2002) J. Leukocyte Biol. 71,
921–931
15. Brown, G. D., Herre, J., Williams, D. L., Willment, J. A., Marshall, A. S., and
Gordon, S. (2003) J. Exp. Med. 197, 1119–1124
16. Brown, G. D., and Gordon, S. (2001) Nature 413, 36–37
17. Gantner, B. N., Simmons, R. M., Canavera, S. J., Akira, S., and Underhill,
D. M. (2003) J. Exp. Med. 197, 1107–1117
18. Xu, S., Ariizumi, K., Caceres-Dittmar, G., Edelbaum, D., Hashimoto, K.,
Bergstresser, P. R., and Takashima, A. (1995) J. Immunol. 154, 2697–2705
19. Ariizumi, K., Shen, G.-L., Ritter, R., III, Kumamoto, T., Edelbaum, D.,
Morita, A., Bergstresser, P. R., and Takashima, A. (2000) J. Biol. Chem.
275, 20157–20167
20. Ariizumi, K., Shen, G.-L., Shikano, S., Ritter, R., III, Zukas, P., Edelbaum,
D., Morita, A., and Takashima, A. (2000) J. Biol. Chem. 275, 11957–11963
21. Taylor, P. R., Brown, G. D., Reid, D. M., Willment, J. A., Martinez-Po-
mares, L., Gordon, S., and Wong, S. Y. (2002) J. Immunol. 169, 3876–3882
22. Taylor, P. R., Reid, D. M., Heinsbroek, S. E., Brown, G. D., Gordon, S., and
Wong, S. Y. (2005) Eur. J. Immunol. 35, 2163–2174
23. Shikano, S., Bonkobara, M., Zukas, P. K., and Ariizumi, K. (2001) J. Biol.
Chem. 276, 8125– 8134
24. Palmowski, M. J., Lopes, L., Ikeda, Y., Salio, M., Cerundolo, V., and Collins,
M. K. (2004) J. Immunol. 172, 1582–1587
25. Zufferey, R., Nagy, D., Mandel, R. J., Naldini, L., and Trono, D. (1997) Nat.
Recognition of Hyphae by Dectin-2
DECEMBER 15, 2006 VOLUME 281 NUMBER 50 JOURNAL OF BIOLOGICAL CHEMISTRY 38865
by guest on December 26, 2015http://www.jbc.org/Downloaded from
Biotechnol. 15, 871– 875
26. Gantner, B. N., Simmons, R. M., and Underhill, D. M. (2005) EMBO J. 24,
1277–1286
27. Humphrey, M. B., Lanier, L. L., and Nakamura, M. C. (2005) Immunol.
Rev. 208, 50–65
28. Ariizumi, K., Meng, Y., Bergstresser, P. R., and Takashima, A. (1995) J. Im-
munol. 154, 6031– 6039
29. Lutz, M. B., Kukutsch, N., Ogilvie, A. L., Rossner, S., Koch, F., Romani, N.,
and Schuler, G. (1999) J. Immunol. Methods 223, 77–92
30. Brown, G. D., Taylor, P. R., Reid, D. M., Willment, J. A., Williams, D. L.,
Martinez-Pomares, L., Wong, S. Y., and Gordon, S. (2002) J. Exp. Med.
196, 407– 412
31. Iobst, S. T., and Drickamer, K. (1994) J. Biol. Chem. 269, 15512–15519
32. Kanazawa, N., Tashiro, K., Inaba, K., and Miyachi, Y. (2003) J. Biol. Chem.
278, 32645–32652
33. van Kooten, C., and Banchereau, J. (1996) Adv. Immunol. 61, 1–77
34. Gibbins, J., Asselin, J., Farndale, R., Barnes, M., Law, C. L., and Watson,
S. P. (1996) J. Biol. Chem. 271, 18095–18099
35. Chan, A. C., Dalton, M., Johnson, R., Kong, G. H., Wang, T., Thoma, R.,
and Kurosaki, T. (1995) EMBO J. 14, 2499–2508
36. Ghosh, S., May, M. J., and Kopp, E. B. (2000) Annu. Rev. Immunol. 16,
225–260
37. Bernard, M., and Latge, J. P. (2001) Med. Mycol. 39, Suppl. 1, 9 –17
38. Fontaine, T., Mouyna, I., Hartland, R. P., Paris, S., and Latge, J. P. (1997)
Biochem. Soc. Trans. 25, 194 –199
39. Douglas, C. M. (2001) Med. Mycol. 39, Suppl. 1, 55–66
40. Trinel, P. A., Jouault, T., Cutler, J. E., and Poulain, D. (2002) Infect. Immun.
70, 5274 –5278
41. Cervera, A. M., Gozalbo, D., McCreath, K. J., Gow, N. A., Martinez, J. P.,
and Casanova, M. (1998) Mol. Microbiol. 30, 67–81
42. Casanova, M., Gil, M. L., Cardenoso, L., Martinez, J. P., and Sentandreu, R.
(1989) Infect. Immun. 57, 262–271
43. Li, R. K., and Cutler, J. E. (1993) J. Biol. Chem. 268, 18293–18299
44. Chaffin, W. L., Lopez-Ribot, J. L., Casanova, M., Gozalbo, D., and Mar-
tinez, J. P. (1998) Microbiol. Mol. Biol. Rev. 62, 130 –180
45. Trinel, P. A., Maes, E., Zanetta, J. P., Delplace, F., Coddeville, B., Jouault, T.,
Strecker, G., and Poulain, D. (2002) J. Biol. Chem. 277, 37260 –37271
46. McGreal, E. P., Rosas, M., Brown, G. D., Zamze, S., Wong, S. Y., Gordon,
S., Martinez-Pomares, L., and Taylor, P. R. (2006) Glycobiology 16,
422–430
47. Underhill, D. M., Rossnagle, E., Lowell, C. A., and Simmons, R. M. (2005)
Blood 106, 2543–2550
48. Bates, E. E., Fournier, N., Garcia, E., Valladeau, J., Durand, I., Pin, J.-J.,
Zurawski, S. M., Patel, S., Abrams, J. S., Lebecque, S., Garrone, P., and
Saeland, S. (1999) J. Immunol. 163, 1973–1983
49. Cosson, P., Lankford, S. P., Bonifacino, J. S., and Klausner, R. D. (1991)
Nature 351, 414 416
50. Zheng, Y. M., Liu, C., Chen, H., Locke, D., Ryan, J. C., and Kahn, M. L.
(2001) J. Biol. Chem. 276, 12999 –13006
Recognition of Hyphae by Dectin-2
38866 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 281 NUMBER 50 DECEMBER 15, 2006
by guest on December 26, 2015http://www.jbc.org/Downloaded from
Ariizumi
Underhill, Ponciano D. Cruz, Jr. and Kiyoshi
Luby-Phelps, Robert P. Kimberly, David
Jin-Sung Chung, Jianming Wu, Kate
Kota Sato, Xiao-li Yang, Tatsuo Yudate,
Responses
Chain to Induce Innate ImmuneγReceptor
for Fungi That Couples with the Fc
Dectin-2 Is a Pattern Recognition Receptor
Developmental Biology:
Molecular Basis of Cell and
doi: 10.1074/jbc.M606542200 originally published online October 18, 2006
2006, 281:38854-38866.J. Biol. Chem.
10.1074/jbc.M606542200Access 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
http://www.jbc.org/content/281/50/38854.full.html#ref-list-1
This article cites 47 references, 29 of which can be accessed free at
by guest on December 26, 2015http://www.jbc.org/Downloaded from
... Dectin-2. Dectin-2 (CLEC6A) defends against fungi and mycobacteria [115,116]. Its activation stimulates the secretion of proinflammatory chemokines and cytokines, including TNFα and IL-12 [116,117], and the receptor enables liver-resident Kupffer cells to recognise and phagocytose cancer cells [118]. ...
... Dectin-2 (CLEC6A) defends against fungi and mycobacteria [115,116]. Its activation stimulates the secretion of proinflammatory chemokines and cytokines, including TNFα and IL-12 [116,117], and the receptor enables liver-resident Kupffer cells to recognise and phagocytose cancer cells [118]. The agonistic antibody BDC-3042 preferentially binds to TAMs in patient tumour samples [117], and it is being investigated in a phase 1/2 trial as a monotherapy and with pembrolizumab (NCT06052852). ...
Article
Full-text available
Tumour-associated macrophages (TAMs) sustain a tumour-supporting and immunosuppressive milieu and therefore aggravate cancer prognosis. To modify TAM behaviour and unlock their anti-tumoural potential, novel TAM-reprogramming immunotherapies are being developed at an accelerating rate. At the same time, scientific discoveries have highlighted more sophisticated TAM phenotypes with complex biological functions and contradictory prognostic associations. To understand the evolving clinical landscape, we reviewed current and past clinically evaluated TAM-reprogramming cancer therapeutics and summarised almost 200 TAM-reprogramming agents investigated in more than 700 clinical trials. Observable overall trends include a high frequency of overlapping strategies against the same therapeutic targets, development of more complex strategies to improve previously ineffective approaches and reliance on combinatory strategies for efficacy. However, strong anti-tumour efficacy is uncommon, which encourages re-directing efforts on identifying biomarkers for eligible patient populations and comparing similar treatments earlier. Future endeavours will benefit from considering the shortcomings of past treatment strategies and accommodating the emerging complexity of TAM biology.
... With the help of TLR2, Dectin-1 can also stimulate the MAPK and NF-κB pathwa ys , triggering pr oinflammatory r esponses (Gantner et al. 2003 ). Dectin-2 can interact with Fc receptor γ (FcR γ ), recognizing high-mannose structures and mediating innate immunity (McGreal et al. 2006, Sato et al. 2006. ...
Article
Full-text available
Candida albicans (C. albicans) is a prevalent opportunistic pathogen that causes mucocutaneous and systemic infections, particularly in immunocompromised individuals. Macrophages play a crucial role in eliminating C. albicans in local and bloodstream contexts, while also regulating antifungal immune responses. However, C. albicans can induce macrophage lysis through pyroptosis, a type of regulated cell death. This process can enable C. albicans to escape from immune cells and trigger the release of IL-1β and IL-18, which can impact both the host and the pathogen. Nevertheless, the mechanisms by which C. albicans triggers pyroptosis in macrophages and the key factors involved in this process remain unclear. In this review, we will explore various factors that may influence or trigger pyroptosis in macrophages induced by C. albicans, such as hypha, ergosterol, cell wall remodeling, and other virulence factors. We will also examine the possible immune response following macrophage pyroptosis.
... Dectin-2 binds mannose carbohydrates, such as the mannose-coated lipoarabinomannan of mycobacteria (36) and a-mannan found in Candida albicans (37). Upon binding, Dectin-2 associates with the Fc receptor gamma chain (38) and signals via SYK and CARD9/BCL-10/MALT1 (CBM complex), triggering NF-kB activation and subsequent proinflammatory cytokine production (39). Despite the importance of furfurman in inducing immune responses, safety issues related to its origin in pigs remains a concern since the fungus could affect animals. ...
Article
Full-text available
Introduction Conventional foot-and-mouth disease (FMD) vaccines have been developed to enhance their effectiveness; however, several drawbacks remain, such as slow induction of antibody titers, short-lived immune response, and local side effects at the vaccination site. Therefore, we created a novel FMD vaccine that simultaneously induces cellular and humoral immune responses using the Dectin-2 agonist, D-galacto-D-mannan, as an adjuvant. Methods We evaluated the innate and adaptive (cellular and humoral) immune responses elicited by the novel FMD vaccine and elucidated the signaling pathway involved both in vitro and in vivo using mice and pigs, as well as immune cells derived from these animals. Results D-galacto-D-mannan elicited early, mid-, and long-term immunity via simultaneous induction of cellular and humoral immune responses by promoting the expression of immunoregulatory molecules. D-galacto-D-mannan also enhanced the immune response and coordinated vaccine-mediated immune response by suppressing genes associated with excessive inflammatory responses, such as nuclear factor kappa B, via Sirtuin 1 expression. Conclusion Our findings elucidated the immunological mechanisms induced by D-galacto-D-mannan, suggesting a background for the robust cellular and humoral immune responses induced by FMD vaccines containing D-galacto-D-mannan. Our study will help to facilitate the improvement of conventional FMD vaccines and the design of next-generation FMD vaccines.
... Dectin-2, another type of C-type lectin expressed by myeloid cells, selectively binds to hyphal mannose of C. albicans instead of the yeast form (Sato et al., 2006). Although the cytoplasmic region of Dectin-2 lacks the ITAM motif which is included in Dectin-1, Fc receptor γ (FcRγ) chain fills in and transduces the signal down to NF-κb for proinflammatory elements production. ...
Article
Full-text available
Candida species are significant causes of mucosal and systemic infections in immune compromised populations, including HIV-infected individuals and cancer patients. Drug resistance and toxicity have limited the use of anti-fungal drugs. A good comprehension of the nature of the immune responses to the pathogenic fungi will aid in the developing of new approaches to the treatment of fungal diseases. In recent years, extensive research has been done to understand the host defending systems to fungal infections. In this review, we described how pattern recognition receptors senses the cognate fungal ligands and the cellular and molecular mechanisms of anti-fungal innate immune responses. Furthermore, particular focus is placed on how anti-fungal signal transduction cascades are being activated for host defense and being modulated to better treat the infections in terms of immunotherapy. Understanding the role that these pathways have in mediating host anti-fungal immunity will be crucial for future therapeutic development.
Article
Full-text available
Patients with decreased levels of CD18 (β2 integrins) suffer from life-threatening bacterial and fungal infections. CD11b, the α subunit of integrin CR3 (CD11b/CD18, αMβ2), is essential for mice to fight against systemic Candida albicans infections. Live elongating C. albicans activates CR3 in immune cells. However, the hyphal ligands that activate CR3 are not well defined. Here, we discovered that the C. albicans Als family proteins are recognized by the I domain of CD11b in macrophages. This recognition synergizes with the β-glucan-bound lectin-like domain to activate CR3, thereby promoting Syk signaling and inflammasome activation. Dectin-2 activation serves as the “outside-in signaling” for CR3 activation at the entry site of incompletely sealed phagosomes, where a thick cuff of F-actin forms to strengthen the local interaction. In vitro, CD18 partially contributes to IL-1β release from dendritic cells induced by purified hyphal Als3. In vivo, Als3 is vital for C. albicans clearance in mouse kidneys. These findings uncover a novel family of ligands for the CR3 I domain that promotes fungal clearance.
Article
Full-text available
Kupffer cells, the liver tissue resident macrophages, are critical in the detection and clearance of cancer cells. However, the molecular mechanisms underlying their detection and phagocytosis of cancer cells are still unclear. Using in vivo genome-wide CRISPR-Cas9 knockout screening, we found that the cell-surface transmembrane protein ERMAP expressed on various cancer cells signaled to activate phagocytosis in Kupffer cells and to control of liver metastasis. ERMAP interacted with β-galactoside binding lectin galectin-9 expressed on the surface of Kupffer cells in a manner dependent on glycosylation. Galectin-9 formed a bridging complex with ERMAP and the transmembrane receptor dectin-2, expressed on Kupffer cells, to induce the detection and phagocytosis of cancer cells by Kupffer cells. Patients with low expression of ERMAP on tumors had more liver metastases. Thus, our study identified the ERMAP–galectin–9-dectin-2 axis as an ‘eat me’ signal for Kupffer cells.
Article
Pattern recognition receptors (PRRs) are crucial immune modulators that orchestrate innate and adaptive immune systems for the regulation of inflammatory responses. Several PRR families and their ligands associated with immune modulation have been identified, which promoted the development of natural and synthetic PRR agonistic ligands as adjuvants in immunotherapy applications. However, conventional adjuvants are mainly based on small molecules, peptides, lipids, and oligonucleotides, which suffer from unfavorable drug-like properties for in vivo applications, including vulnerability to biodegradation, undesirable pharmacokinetic profiles, and poor cellular uptake. Nanoparticle formulation is a promising approach for addressing the issues with conventional adjuvants. This review aims to provide a broad understanding of nano-adjuvants utilized in cancer immunotherapy. For this purpose, we introduce the background, summarize the current knowledge on PRRs and their ligands for some representative classes, and highlight various nanoparticle platforms that are utilized to construct nano-adjuvants in cancer immunotherapy. We also discuss some design considerations for optimal nano-adjuvant formulations with potent adjuvant activity and desired in vivo performance. Nanoparticles provide a robust and versatile platform to shape conventional adjuvants into more drug-like formulations, and the preclinical studies and clinical trials have demonstrated their potential in cancer immunotherapy. The emergence of cancer immunotherapy in clinics will fuel continuous efforts toward highly efficient nano-adjuvant systems that can strongly boost the antitumor immune responses in cancer immunotherapy. The accumulated knowledge gained through the progress will provide important insights into optimal nano-adjuvant formulations and potentially guide their clinical translation.
Article
Full-text available
Interstitial Cystitis/Bladder Pain Syndrome (IC/BPS) is a chronic disorder characterized by pelvic and/or bladder pain, along with lower urinary tract symptoms that have a significant impact on an individual’s quality of life. The diverse range of symptoms and underlying causes in IC/BPS patients pose a significant challenge for effective disease management and the development of new and effective treatments. To facilitate the development of innovative therapies for IC/BPS, numerous preclinical animal models have been developed, each focusing on distinct pathophysiological components such as localized urothelial permeability or inflammation, psychological stress, autoimmunity, and central sensitization. However, since the precise etiopathophysiology of IC/BPS remains undefined, these animal models have primarily aimed to replicate the key clinical symptoms of bladder hypersensitivity and pain to enhance the translatability of potential therapeutics. Several animal models have now been characterized to mimic the major symptoms of IC/BPS, and significant progress has been made in refining these models to induce chronic symptomatology that more closely resembles the IC/BPS phenotype. Nevertheless, it's important to note that no single model can fully replicate all aspects of the human disease. When selecting an appropriate model for preclinical therapeutic evaluation, consideration must be given to the specific pathology believed to underlie the development of IC/BPS symptoms in a particular patient group, as well as the type and severity of the model, its duration, and the proposed intervention’s mechanism of action. Therefore, it is likely that different models will continue to be necessary for preclinical drug development, depending on the unique etiology of IC/BPS being investigated.
Article
Full-text available
Contact between dendritic cells (DC) and resting T cells is essential to initiate a primary immune response. Here, we demonstrate that ICAM-3 expressed by resting T cells is important in this first contact with DC. We discovered that instead of the common ICAM-3 receptors LFA-1 and alphaDbeta2, a novel DC-specific C-type lectin, DC-SIGN, binds ICAM-3 with high affinity. DC-SIGN, which is abundantly expressed by DC both in vitro and in vivo, mediates transient adhesion with T cells. Since antibodies against DC-SIGN inhibit DC-induced proliferation of resting T cells, our findings predict that DC-SIGN enables T cell receptor engagement by stabilization of the DC-T cell contact zone.
Article
Full-text available
The cell wall is essential to nearly every aspect of the biology and pathogenicity of Candida albicans. Although it was initially considered an almost inert cellular structure that protected the protoplast against osmotic offense, more recent studies have demonstrated that it is a dynamic organelle. The major components of the cell wall are glucan and chitin, which are associated with structural rigidity, and mannoproteins. The protein component, including both mannoprotein and nonmannoproteins, comprises some 40 or more moieties. Wall proteins may differ in their expression, secretion, or topological location within the wall structure. Proteins may be modified by glycosylation (primarily addition of mannose residues), phosphorylation, and ubiquitination. Among the secreted enzymes are those that are postulated to have substrates within the cell wall and those that find substrates in the extracellular environment. Cell wall proteins have been implicated in adhesion to host tissues and ligands. Fibrinogen, complement fragments, and several extracellular matrix components are among the host proteins bound by cell wall proteins. Proteins related to the hsp70 and hsp90 families of conserved stress proteins and some glycolytic enzyme proteins are also found in the cell wall, apparently as bona fide components. In addition, the expression of some proteins is associated with the morphological growth form of the fungus and may play a role in morphogenesis. Finally, surface mannoproteins are strong immunogens that trigger and modulate the host immune response during candidiasis.
Article
Full-text available
The transmembrane domain of the alpha chain of the T-cell receptor is responsible both for its assembly with the CD3 delta chain and for rapid degradation of the unassembled chain within the endoplasmic reticulum. The determinant for both assembly and degradation is located in a segment of eight residues containing two basic amino acids. We show here that placement of a single basic residue in the transmembrane domain of the Tac antigen can induce interaction with the CD3 chain, through its transmembrane acidic residue. This interaction is most favoured when the interacting residues are located at the same level in the membrane. The ability to induce protein-protein interaction by placing charge pairs within transmembrane domains suggests an approach to producing artificial dimers.
Article
Full-text available
Walls of the two cellular forms (blastoconidia and mycelia) of Candida albicans ATCC 26555 were obtained from cells metabolically labeled (6-h pulse) with 14C-protein hydrolysate and [3H]threonine. Walls were purified by thorough washings with buffered and sodium dodecyl sulfate solutions and digested with Zymolyase 20T. The enzymatic treatment released four major high-molecular-weight mannoproteins (HMWM), with apparent molecular masses of 650, 500, 340, and 200 kilodaltons (HMWM-650, HMWM-500, HMWM-340, and HMWM-200, respectively), from yeast cells, whereas two high-molecular-mass mannoproteins (HMWM-260 and HMWM-180) were solubilized from mycelial cells. Some additional minor low-molecular-weight species were also detected in the enzymatic digests of walls from both types of cell. Single and dual pulse-chase experiments indicated that the HMWM-260 and HMWM-180 species reflect de novo synthesis of new proteins specific for the mycelia and do not represent a topological rearrangement of blastoconidium wall components. Monoclonal antibodies were raised against the HMWM-260 species (quantitatively the predominant component in the mycelial walls), and polyclonal rabbit antibodies were obtained against yeast or mycelial cell walls. Anti-mycelial cell wall polyclonal antibodies were adsorbed to whole killed blastoconidia to remove antibodies against common blastoconidium and mycelial wall antigens. Titration by enzyme-linked immunosorbent assay revealed that the monoclonal antibodies could recognize an epitope of the protein moiety of the HMWM-260 mannoprotein. Immunoblotting and immunofluorescence techniques using these monoclonal and polyclonal antibodies confirmed that the HMWM-260 and HMWM-180 species are specific components of the envelope of the mycelial cell walls.
Article
Full-text available
An antigen (Ag), as defined by reactivity with a monoclonal antibody (mAb 10G), was found on the cell wall surface and on the plasma membrane of Candida albicans yeast cells (Li, R. K., and Cutler, J. E. (1991) J. Gen. Microbiol. 137, 455-464). The 10G Ag from the cell surface location was solubilized by treatment with beta-mercaptoethanol and Zymolyase. Antigenic activity of the extract was lost following periodate oxidation, but was stable to proteolytic enzyme and heat treatments. Mannoproteins in the extract were fractionated by hexadecyltrimethylammonium bromide. A fraction containing the 10G Ag was degraded by mild acid hydrolysis and a tetrahexose eluted from a P-2 size exclusion column was found to be the epitope (10G epitope) part of the 10G Ag. By use of gas-liquid chromatography, mass spectroscopy, and H-1 proton NMR analysis, the 10G epitope was determined to be a linear beta-1,2-linked tetramannose. The 10G Ag, which was immunopurified from the Zymolyase extract, and the 10G epitope were involved in the attachment of C. albicans to mouse spleen marginal zone macrophages. Both the 10G Ag and the purified 10G epitope specifically inhibited C. albicans adherence in an ex vivo binding assay. Latex beads adsorbed with the 10G Ag or the 10G epitope adhered specifically to the splenic marginal zone macrophages. These data show that a beta-1,2-linked mannotetraose part of a C. albicans cell wall mannoprotein is an adhesin site that may play a role in pathogenesis of disseminated candidiasis.
Article
Contact between dendritic cells (DC) and resting T cells is essential to initiate a primary immune response. Here, we demonstrate that ICAM-3 expressed by resting T cells is important in this first contact with DC. We discovered that instead of the common ICAM-3 receptors LFA-1 and alphaDbeta2, a novel DC-specific C-type lectin, DC-SIGN, binds ICAM-3 with high affinity. DC-SIGN, which is abundantly expressed by DC both in vitro and in vivo, mediates transient adhesion with T cells. Since antibodies against DC-SIGN inhibit DC-induced proliferation of resting T cells, our findings predict that DC-SIGN enables T cell receptor engagement by stabilization of the DC-T cell contact zone.
Article
Antigen presenting cells (macrophages and dendritic cells) express pattern recognition molecules that are thought to recognize foreign ligands during early phases of the immune response. The best known of these are probably the Toll-like receptors, but a number of other receptors are also involved. Several of these recognize endogenous as well as exogenous ligands, suggesting that they play a dual role in normal tissue function and host defense.
Article
The polysaccharide β(1,3)-D-glucan is a component of the cell wall of many fungi. Synthesis of the linear polymer is catalysed by UDP-glucose β(1,3)-D-glucan β(3)-D-glucosyltransferase. Because this enzyme has a key role in fungal cell-wall synthesis, and because many organisms that are responsible for human mycoses, including Candida albicans, Aspergillus fumigatus and Cryptococcus neoformans, produce walls that are rich in β(1,3)-glucan, it has been and remains the focus of intensive study. From early characterization of the enzymatic activity in Saccharomyces cerevisiae, advances have been made in purification of the enzyme, identification of essential subunits and description of regulatory circuitry that controls expression and localization of different components of the multisubunit enzyme complex. Progress in each of these areas has been enhanced dramatically by the availability of specific inhibitors of the enzymatic reaction that produces β(1,3)-glucan. These natural product inhibitors have utility both as tools to dissect the biology of β(1,3)-glucan synthase and as sources for development of semisynthetic derivatives with clinical utility in treatment of human fungal disease. This review will focus on the biochemistry, genetics and regulation of the enzyme.
Article
The innate immune system is essential for host defense and is responsible for early detection of potentially pathogenic microorganisms. Upon recognition of microbes by innate immune cells suchas macrophages and dendritic cells, diverse signaling pathways are activated that combine to define inflammatory responses that direct sterilization of the threat and/or orchestrate development of the adaptive immune response. Innate immune signaling must be carefully controlled, and regulation comes in part from interactions between activating and inhibiting signaling receptors. Toll-like receptors (TLR) have recently emerged as key receptors responsible for recognizing specific conserved components of microbes including lipopolysaccharides from Gram-negative bacteria, CpG DNA, and flagellin. Full activation of inflammatory responses by TLR may require the assembly of receptor signaling complexes including other transmembrane proteins that may influence signal transduction. In addition to TLR, many additional receptors participate in innate recognition of microbes, and recent studies demonstrate strong interactions between signaling through these receptors and signaling through TLR. Useful models for these interacting signaling pathways are now emerging and should pave the way for understanding the molecular mechanisms that drive the rich diversity of inflammatory responses.
Article
The mannose receptor recognizes the patterns of carbohydrates that decorate the surfaces and cell walls of infectious agents. This macrophage and dendritic cell pattern-recognition receptor mediates endocytosis and phagocytosis. The mannose receptor is the prototype of a new family of multilectin receptor proteins (membrane-spanning receptors containing eight-ten lectin-like domains, which appear to play a key role in host defense) and provides a link between innate and adaptive immunity. Recent advances include the identification of three new members of the mannose receptor family, additional work on defining the molecular requirements for sugar binding, a role for the mannose receptor in antigen presentation of lipoglycan antigens and evidence that the mannose receptor is associated with a signal transduction pathway leading to cytokine production.