You are on page 1of 7

Immune signaling pathways regulating bacterial and malaria parasite infection of the

mosquito Anopheles gambiae

Stephan Meister, Stefan M. Kanzok, Xue-li Zheng, Coralia Luna, Tong-Ruei Li, Ngo T. Hoa,
John Randall Clayton, Kevin P. White, Fotis C. Kafatos, George K. Christophides, and Liangbiao
Zheng
PNAS 2005;102;11420-11425; originally published online Aug 2, 2005;
doi:10.1073/pnas.0504950102
This information is current as of May 2007.

Online Information High-resolution figures, a citation map, links to PubMed and Google Scholar,
& Services etc., can be found at:
www.pnas.org/cgi/content/full/102/32/11420
Supplementary Material Supplementary material can be found at:
www.pnas.org/cgi/content/full/0504950102/DC1
References This article cites 36 articles, 16 of which you can access for free at:
www.pnas.org/cgi/content/full/102/32/11420#BIBL
This article has been cited by other articles:
www.pnas.org/cgi/content/full/102/32/11420#otherarticles
E-mail Alerts Receive free email alerts when new articles cite this article - sign up in the box
at the top right corner of the article or click here.
Rights & Permissions To reproduce this article in part (figures, tables) or in entirety, see:
www.pnas.org/misc/rightperm.shtml
Reprints To order reprints, see:
www.pnas.org/misc/reprints.shtml

Notes:
Immune signaling pathways regulating bacterial
and malaria parasite infection of the mosquito
Anopheles gambiae
Stephan Meister*†‡, Stefan M. Kanzok‡§¶, Xue-li Zheng§储, Coralia Luna§, Tong-Ruei Li**, Ngo T. Hoa§,
John Randall Clayton*¶, Kevin P. White**, Fotis C. Kafatos*†, George K. Christophides*†,††‡‡, and Liangbiao Zheng§††‡‡
*European Molecular Biology Laboratory, Meyerhofstrasse 1, 69117 Heidelberg, Germany; and Departments of §Epidemiology and Public Health and
**Genetics, Yale University School of Medicine, 60 College Street, New Haven, CT 06520

Contributed by Fotis C. Kafatos, June 13, 2005

We show that, in the malaria vector Anopheles gambiae, expres- the Anopheles gambiae REL2 and show that it regulates expression
sion of Cecropin 1 is regulated by REL2, an NF-␬B-like transcription of the AMP genes CEC1, CEC3, and GAM1 and a key parasite
factor orthologous to Drosophila Relish. Through alternative splic- antagonist, LRIM1 (13). Two REL2 gene products are produced via
ing, REL2 produces a full-length (REL2-F) and a shorter (REL2-S) alternative splicing: A full-length form (REL2-F) includes the
protein isoform lacking the inhibitory ankyrin repeats and death carboxyl-terminal ANK and death domains, and a shorter form
domain. RNA interference experiments show that, in contrast to (REL2-S) lacks these domains. RNAi-dependent gene silencing in
Drosophila Relish, which responds solely to Gram-negative bacte- adult Anopheles mosquitoes revealed that REL2-F and IMD are
ria, the Anopheles REL2-F and REL2-S isoforms are involved in implicated in defense against the Gram-positive bacterium Staph-
defense against the Gram-positive Staphylococcus aureus and the ylococcus aureus, whereas REL2-S is involved in defense against the
Gram-negative Escherichia coli bacteria, respectively. REL2-F also Gram-negative bacterium Escherichia coli. Thus, through alterna-
regulates the intensity of mosquito infection with the malaria tive splicing, Anopheles uses a single NF-␬B gene (REL2) to mediate
parasite, Plasmodium berghei. The adaptor IMD shares the same immune reactions, for which Drosophila employs two distinct genes,
activities as REL2-F. Microarray analysis identified 10 additional Relish and Dif. We show that the Anopheles Imd pathway leading to
genes regulated by REL2, including CEC3, GAM1, and LRIM1. REL2-F activation is also involved in limiting Plasmodium parasite
infections. Thus, a classical immune signaling pathway is implicated
innate immunity 兩 NF-␬B 兩 Relish 兩 Cecropin 兩 Imd
in parasite survival in the Anopheles mosquito vector. Finally,
REL2-F is involved in regulation of the melanization reaction;
rom insects to mammals, NF-␬B-like transcription factors play
F central roles in induction and regulation of innate immune
responses. In Drosophila, three NF-␬B-like proteins have been
silencing of REL2-F causes limited but consistent melanization of
Plasmodium ookinetes in the Anopheles midgut.

identified. Dorsal is largely engaged in early developmental pro- Materials and Methods
cesses, whereas Relish and Dif are involved in immune signaling cDNA Production. Total RNA was extracted from 25 larvae, pupae,
leading to transcriptional activation of genes, including those en- or adult Anopheles gambiae and ⬇5,000 embryos with TRIzol
coding antimicrobial peptides (AMPs) (1). The Toll pathway, which reagent (Invitrogen) and treated with DNaseI, and 0.2 ␮g were used
is used, principally, to counteract infections by fungi (2) and for reverse transcription using the Omniscript kit (Qiagen, Santa
Gram-positive bacteria (3), is activated through nuclear transloca- Clarita, CA).
tion of Dif; one of the transcriptionally induced genes encodes the
AMP Drosomycin (4, 5). In contrast, the immunodeficiency (Imd) dsRNA Construction. PCR amplicons tailed with T7 promoter se-
pathway deals primarily with Gram-negative bacterial infections (6) quences (see Table 2, which is published as supporting information
through nuclear translocation of Relish, leading to transcriptional on the PNAS web site) were used to synthesize dsRNAs with the
induction of an array of AMPs, including Cecropins A and B, MEGAshortscript kit (Ambion) that were DNaseI-treated, cleaned
Attacin, Drosocin, and Diptericin (7). up by phenol:chloroform extraction, precipitated with isopropanol,
Relish is a modular protein consisting of two main domains. The and resuspended in nuclease-free water.
Rel-homology domain (RHD) binds to DNA and activates gene
expression, whereas the inhibitory I␬B domain, harboring several
Luciferase Assays. Cells were plated in 24-well plates and, upon
ankyrin (ANK) repeats, blocks the nuclear localization signal and,
reaching ⬇70% confluence, were incubated for 30 min at room
thus, keeps Relish in the cytosol. After bacterial infection, the RHD
temperature with 7 ␮g of dsRNA in FBS-free medium. Transfec-
domain is separated from ANK by targeted proteolysis and is
tions with CEC1 or KIN1 reporter (300 ng per well) and reference
translocated into the nucleus. Recent studies have shown that the
caspase Dredd is involved, directly or indirectly, in the proteolytic
activation of Relish (8). In the yellow fever mosquito Aedes aegypti, Abbreviations: AMP, antimicrobial peptide; ANK, ankyrin; Imd, immunodeficiency; KD,
three products of the Relish gene are produced via alternative knockdown; RHD, Rel-homology domain; RNAi, interfering RNA.
splicing and encode the full-length protein, the RHD, and the ANK †Presentaddress: Department of Biological Sciences, Imperial College London, SW7 2AZ
domain (9). Activation of the Aedes full-length protein presumably London, United Kingdom.
requires proteolytic processing, as in Drosophila. ‡S.M. and S.M.K. contributed equally to this work.
The Anopheles gambiae genome sequence (10) has allowed us to ¶Present address: Department of Molecular Microbiology and Immunology, Johns Hopkins

compare putative immunity genes between Anopheles and Dro- Bloomberg School of Public Health, Baltimore, MD 21205.
sophila (11). The majority of intracellular components of the Toll 储Present
address: Department of Parasitology, Southern Medical University, Guangzhou,
and Imd pathways are conserved, but several differences exist. Most Guangdong Province, China.
striking is the absence of Dif in Anopheles, suggesting that a ††G.K.C. and L.Z. contributed equally to this work.
functional Toll pathway would require REL1 (the orthologue of ‡‡To whom correspondence may be addressed. E-mail: g.christophides@imperial.ac.uk or
Dorsal, previously designated as Gambif1) (12) or REL2 (the liangbiao.zheng@yale.edu.
orthologue of Relish) to substitute for Dif. Here, we characterize © 2005 by The National Academy of Sciences of the USA

11420 –11425 兩 PNAS 兩 August 9, 2005 兩 vol. 102 兩 no. 32 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0504950102


Fig. 1. Genomic organization and alternative splicing of the Anopheles REL2 gene. Introns (lines) and exons (open boxes) are shown, based on Ensembl gene
prediction (A) and actual data (B). The length (bp) of introns and exons is shown below and above the lines and boxes, respectively. E, putative ␬B elements.
(C) REL2 has two alternative 5⬘ exons. The 5⬘ end of REL2-S has not been confirmed and is indicated by dashed intron lines. Alternative splicing occurs at the 3⬘
end with a REL2-S-specific exon. Protein domains are depicted by shades of gray inside the exons. Q兾H, glutamine and histidine-rich region; RHD, Relish homology

IMMUNOLOGY
domain; IPT兾TIG, DNA-binding domain; NLS, nuclear localization signal; ANK, ANK-repeat domain; DD, death domain. Putative start- and stop-codons and target
regions of dsRNA constructs (RHD dsRNA and ANK dsRNA) are indicated.

pAct5C-Renilla (3 ng per well) DNA constructs were performed (Promega) and labeled as described in ref. 19. Equal volumes of
with the Qiagen Effectene kit (14). Cells were harvested 16–18 h Cy-3- and Cy-5-labeled samples were combined in hybridization
later and subjected to Dual Luciferase (Promega) assays according buffer supplemented with poly(A)-DNA and hybridized to mi-
to the manufacturer’s instructions. Relative light units per second croarray slides at 42°C overnight. Hybridized arrays were washed,
were measured in a luminometer (Berthold, Bundoora, Australia). air-dried, and scanned by using the ScanArray 3000 (GSI Lumonics,
Billerica, MA). Two independent experiments with dye swaps were
Gene Silencing in Adult Anopheles gambiae. One- to two-day-old performed. Expression ratios were calculated and analyzed by using
adult female mosquitoes were injected with 207 ng of dsRNA, as the TM4 microarray software package (www.tigr.org兾software兾
described in ref. 15, and, after 4 days, was subjected to infection tm4兾). After spot quality filtering, expression data were normalized
assays. by using the Lowess (Locfit) algorithm, and clustered with k-mean
and hierarchical clustering with the program TMEV (The Institute
Immune Challenge in Mosquito Cell Lines. Lipopolysaccharides from for Genomic Research) (19). Elements with at least 500 signal
E. coli O26:B6 (Sigma Aldrich) were dissolved in water and added intensity values were considered.
to cells at 10 ␮g兾ml. Micrococcus luteus and E. coli (DH5␣) bacteria
and Beauveria bassiana spores (American Type Culture Collection) Results
were heat-inactivated by boiling for 5 min in PBS and added to cells Genomic Organization and Expression of REL2. RT-PCR and RACE
at 100 microbes per cell. using RNA from adult female Anopheles gambiae were used to
determine the structural organization of REL2. This gene consists
Bacterial and Malaria Infection of Adult Anopheles gambiae. Bacteria of 11 exons, encompassing 10.9 kb (Fig. 1). It encodes putative
were grown to OD600 ⫽ 0.7, precipitated, washed, and resuspended protein(s) with a glutamine- and histidine-rich region, potentially
in PBS. Bacterial suspensions of 69 nl (E. coli at OD600 ⫽ 0.01 and implicated in protein–protein interactions, followed by RHD, IPT兾
S. aureus at OD600 ⫽ 0.4) were injected into the mosquito thorax TIG (DNA binding), ANK and death domains, and a nuclear
with a nanoinjector (Nanoject, Drummond Scientific, Broomall, localization signal.
PA). For malaria infections, mosquitoes were fed for 30 min at 21°C We detected two REL2 transcripts: full-length REL2-F, and a
on CD1 mice infected with transgenic GFP-fluorescing Plasmo- shorter form, REL2-S, encompassing a unique exon at its 3⬘ end.
dium berghei (16, 17). For mRNA splicing, this exon uses a splice acceptor that is 4 bp 3⬘
upstream to that used in REL2-F and results in a downstream early
Quantitative RT-PCR. cDNA was produced as described above. termination codon. Thus, REL2-F encodes a protein with all of the
Transcript abundance was measured with an Applied Biosystems above features, whereas REL2-S is missing the ANK and death
7500 Real-Time PCR system according to the manufacturer’s domains. Additional alternative splicing at the 5⬘ end of REL2 may
instructions. PCRs of 25 ␮l consisted of 1⫻ SYBR green mix, 900 also produce two transcript variants for each form (Fig. 1C).
nM forward and reverse primers, and cDNA, corresponding to 0.02 REL2-F shows significant similarity to the Aedes (55%) and
␮g of total RNA. Primer sequences are shown in Table 2. Drosophila (29%) Relish (see Fig. 7, which is published as support-
ing information on the PNAS web site).
Microarray Analysis. Slides were printed as described in ref. 18; REL2-F and -S transcripts are expressed constitutively through-
prehybridized in 5⫻ standard saline citrate (SSC) (1⫻ SSC ⫽ 0.15 out Anopheles gambiae development, albeit at different levels:
M sodium chloride兾0.015 M sodium citrate, pH 7), 0.1% SDS, and REL2-F is expressed more strongly (Fig. 2A). Both transcripts are
1% BSA at 42°C for 45 min; washed at room temperature with also expressed in cultured cell lines, Sua1B (Fig. 2B) and 4a3A (data
deionized water; dipped in 100% isopropanol; and dried. mRNA not shown). Cells challenged with heat-killed E. coli (Gram-
samples were prepared by using the PolyTract isolation system negative) or purified E. coli lipopolysaccharides (LPS) up-regulate

Meister et al. PNAS 兩 August 9, 2005 兩 vol. 102 兩 no. 32 兩 11421


Fig. 2. Expression profile of REL2 in mosquito cultured cells
and developmental stages. (A) Relative abundance of REL2-F
and REL2-S transcripts at different stages of mosquito devel-
opment. (B) Transcriptional regulation of REL2 transcripts in
immune-challenged cultured Sua1B cells. REL2-F was detected
by using primer pairs targeting the ANK domain. REL2-S was
detected with one primer specific to the REL2-S transcript (see
Materials and Methods). PCR conditions: 30 cycles, 60°C an-
nealing temperature, 1.5 min extension time, and 2 ␮g of total
RNA used per reaction. *, a band which includes the intron
sequence of REL2-S. It is unclear whether this transcript is part
of an additional splice variant of REL2. The identity of REL2-S
bands was confirmed by sequencing. Control RT-PCR with rpS7
showed no genomic DNA contamination. CTRL, control; B.b.,
B. bassiana; M.l., M. luteus; E.c., E. coli; 么, males; 乆, females.

REL2-F significantly and REL2-S to a lower level. We cannot (⬇60% and 68%, respectively), although not as efficiently as REL2
exclude the possibility that induction after LPS challenge was due KD. This is consistent with evidence from Drosophila imd mutants,
to the common peptidoglycan contaminants of LPS preparations in which Cecropin A expression is reduced but not abolished (23).
(20). No significant up-regulation was detected after cell challenge Surprisingly but consistently, silencing of CASPL1 and IKK1
with the fungus Beauveria bassiana or the Gram-positive bacterium or IKK2 had opposite effects in the two cell lines. In Sua1B cells,
M. luteus. KD of CASPL1, but not of IKK1 or IKK2, substantially reduced
CEC1 promoter expression, whereas, in 4a3A cells, KD of IKK1
Expression of CEC1 Is Regulated by REL2-F. A putative AMP target or IKK2, but not of CASPL1, reduced CEC1 promoter expres-
gene of Anopheles immune signaling pathways is CEC1 (previously sion. We used semiquantitative RT-PCR to estimate the level of
designated as Cecropin A) (14, 21). Compared with Drosophila gene silencing (Fig. 4A). Importantly, KD of IKK1 and IKK2 in
Cecropins, which are mostly active against Gram-negative bacteria, Sua1B was highly effective, suggesting that these genes are not
CEC1 has a wide spectrum of antibacterial activities and is induced implicated in the regulation of CEC1 in Sua1B cells. KD of
by Plasmodium and both Gram types of bacteria (21). Sua1B and CASPL1 in the same cells was less effective (⬇40%) and, yet,
4a3A cells have high levels of constitutive CEC1 expression that are
associated with substantial reduction in CEC1 activity (Fig. 3B).
further enhanced by treatment with heat-killed E. coli or M. luteus
A plausible explanation that requires further investigation is that
(data not shown). We examined the effect of REL proteins on
IKK1兾IKK2 and CASPL1 are implicated in alternative branches
CEC1 gene expression in Sua1B cells after transfection with a
CEC1-promoter–luciferase-reporter construct (14), silencing of
REL1 and REL2 genes by RNAi, and subsequent monitoring of
luciferase activity. Targeting either the RHD or ANK domains
of REL2 resulted in substantial reduction of luciferase activity,
typically 5- to 10-fold. In contrast, REL1 knockdown (KD) had no
significant effect (Fig. 3A).
Because Anopheles gambiae shows no transitive RNAi effects,
i.e., dsRNAs are effective only against mRNAs bearing the corre-
sponding sequences (22), targeting the RHD should silence both
REL2 forms, whereas ANK-domain KD should affect only
REL2-F. We observed that KDs using ANK dsRNA result in a
major (⬇91%) reduction in CEC1-promoter expression, essentially
indistinguishable from that observed with RHD dsRNA. Thus, it
appears that REL2-F largely controls constitutive CEC1-promoter
expression in Sua1B cells, whereas REL2-S has no major effect
(Fig. 3A). Similar results were obtained for the 4a3A cells (data not
shown). Proteolytic cleavage is necessary to remove the inhibitory
ANK domain of Drosophila Relish, and this processing would also
be expected for REL2-F activation. By extension, these results
suggest that REL2-F is constitutively activated in these mosquito
cells.
Other Anopheles orthologues of Drosophila Imd-pathway com-
ponents, the Imd adaptor (IMD) and the caspase Dredd
(CASPL1), were identified in ref. 11; orthologues of IKK␤ (IKK1)
and IKK␥ (IKK2) are reported here (see Table 3, which is published
Fig. 3. CEC1 promoter activity in cultured Anopheles cells. (A) REL2, but not
as supporting information on the PNAS web site). To examine the
REL1, is required for expression of CEC1 promoter in Sua1B cells, which is
degree of functional Imd-pathway conservation between Anopheles equally affected by dsRNAs encompassing the RHD and the ANK-repeats
and Drosophila, we silenced these genes in Sua1B and 4a3A cells by domain. (B) REL2, IMD, and CASPL1, but not IKK1 or IKK2, are required for
using RNAi and observed significant down-regulation (Fig. 3 B and CEC1 expression in Sua1B cells. (C) REL2, IMD, IKK1, and IKK2, but not CASPL1,
C). In both cell lines, IMD KD reduced CEC1-promoter expression are required for CEC1 expression in 4a3A cells. CTRL, control.

11422 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0504950102 Meister et al.


genesis, a protein disulfide isomerase, KIN1 (see below), and
LRIM1, a leucine-rich repeat immune protein with strong anti-
parasitic activity (13). The CEC3 gene is tightly clustered with
CEC1 and CEC2, and its promoter region includes two NF-␬B-
binding sites (14), suggesting regulatory similarities to CEC1. At
least 1, and as many as 10, ␬B elements were detected in adjacent
genomic regions of all regulated genes (see Table 4, which is
published as supporting information on the PNAS web site).
KIN1 encodes a short mature cationic polypeptide with a kinino-
gen domain, which is enriched in histidine (26 residues, 36%), plus
an amino-terminal signal peptide. KIN1 is strongly up-regulated by
immune challenges (18) and Plasmodium infection (25). Examina-
tion of a KIN1-promoter luciferase construct in Sua1B cells showed
that REL2 is, indeed, required for KIN1 expression (see Fig. 8,
which is published as supporting information on the PNAS web
site).

REL2 Is Required for Antibacterial Defense of Adult Anopheles gam-


biae. We investigated, by RNAi analysis, the possible role of REL2
and other components of the Imd pathway in defense against
bacterial infection in adult mosquitoes. DsRNA was injected into
newly emerged adult females that were infected, 4 days later, with
bacteria representative of the Gram-positive (S. aureus) or the
Fig. 4. Efficacy of RNAi-mediated gene silencing. (A) KD of IKK1, IKK2, and Gram-negative (E. coli) types, as described in ref. 15. Real-time

IMMUNOLOGY
CASPL1 genes in Sua1B cultured cells. RNAi effectiveness was determined by quantitative RT-PCR confirmed that RHD dsRNA injection si-
semiquantitative RT-PCR (Left). Quantification of PCR products (Right) was lences both REL2-S and REL2-F transcripts equally (⬇50%),
performed by using rpS7 as an internal reference and untreated cells as a
whereas ANK dsRNA decreases the levels of only REL2-F (Fig.
calibrator. (B) REL2 gene silencing in adult mosquitoes after injections with
either RHD or ANK dsRNA. Relative REL2-S and REL2-F transcript levels in KD 4B). Interestingly, REL2-S transcript levels were up-regulated by
compared with control GFP dsRNA-treated mosquitoes were determined in 1.5-fold in dsANK-injected mosquitoes.
the carcass, midgut, and whole mosquitoes 4 days after injections by using Mosquito survival was assessed daily for 7 days after bacterial
quantitative RT-PCR. infection (Fig. 5). Simultaneous KD of REL2-F and REL2-S
transcripts (RHD dsRNA) significantly compromised mosquito
defense against both S. aureus and E. coli. In contrast, KD of
of the CEC1-activation pathway that are differentially active in REL2-F alone (ANK dsRNA) reduced the survival of S. aureus- but
these two cell types. not of E. coli-infected mosquitoes. Similar results were obtained
with IMD KD, suggesting that IMD acts in concert with REL2-F in
REL2 Regulates the Expression of AMPs and Other Immunity Genes. defense against S. aureus. Survival of CASPL1-KD mosquitoes did
We performed a DNA microarray analysis (18) to identify genes not detectably decrease after infection (data not shown). Prelimi-
regulated by REL2 in Anopheles 4a3A cells by comparing the nary results indicated that KD of REL1 had no measurable effect
expression profiles of cells treated with either REL2 (RHD domain) on mosquito survival, suggesting that REL2 is sufficient to confer
or control lacZ dsRNAs. Cells were challenged, 24 h after dsRNA resistance to bacterial infections.
addition, by a 12-h exposure to heat-killed E. coli, and expression
profiles were assessed at 0, 2, 6, and 12 h. From 3,840 EST clones The Anopheles Imd Pathway Regulates Malaria Parasite Infection. It
present on the microarray, nine were consistently down-regulated has been reported previously that CEC1 and other AMP genes are
in REL2 KD cells by at least 1.7-fold and one by 1.2-fold (Table 1). up-regulated when Plasmodium crosses the Anopheles midgut (11,
Six of these genes have been categorized previously as defense兾 21, 26). Furthermore, our microarray results indicated that REL2
immunity-related (24). They encode the AMPs CEC3 and GAM1 regulates expression of the major Plasmodium antagonist LRIM1.
(previously CecB and gambicin, respectively), the serine protease We tested the effect of REL2 and IMD KD on P. berghei survival
CLIPB14, another two CLIP-domain serine proteases, a fibrinogen- to the oocyst stage in the mosquito midgut. At 4 days after dsRNA
domain lectin, a Brix-domain protein implicated in ribosome bio- treatment, mosquitoes were infected with GFP-marked parasites

Table 1. Effect of REL2 pathway gene silencing on parasite development


Gene KD Midguts Parasites per midgut Prevalence, % Mel, % P (KS) P (t)

Control (5) 79 86 87.34 0.07


IMD (5) 79 153 96.20 0.13 0.044 0.017
Control (4) 49 51 87.76 0.63
REL2-RHD (4) 49 104 91.84 4.37 0.147 0.035
Control (3) 42 82 97.62 0.00
REL2-ANK (3) 42 166 95.24 4.87 0.093 0.008

Numbers of oocysts and melanized parasites are reported as three experimental datasets for knocked-down
genes or REL2 domains and control dsRNA-treated mosquitoes (GFP). The number of replicates is indicated in
parenthesis in the first column. Each replicate used different batches of mosquitoes, which were fed on the same
infected mouse. Identical numbers of midguts from control and KD mosquitoes were assessed for each replicate
(see also Table 5). P values indicate whether the distributions of parasites in KD and control midguts in each
dataset are similar and were determined by KS and Student’s t test. Mel, melanized parasites (ookinetes); KS,
Kolmogorov–Smirnov test. Parasites per midgut are oocysts and melanized ookinetes.

Meister et al. PNAS 兩 August 9, 2005 兩 vol. 102 兩 no. 32 兩 11423


Fig. 6. REL2- and IMD-mediated Plasmodium killing and melanization in the
Fig. 5. Mosquito survival to bacterial infection. Survival of adult mosquitoes mosquito vector. (A) Averaged P. berghei oocyst load per midgut after KD of
to Gram-positive S. aureus (A) and Gram-negative E. coli (B) infections after KD IMD, REL2-ANK, and REL2-RHD. Oocyst load after individual gene KD was
of IMD, REL2-RHD, and REL2-ANK. GFP dsRNA-treated mosquitoes were used normalized relative to oocyst load of parallel dsGFP treatment and presented
as a control. The survival assay was repeated at least 3 and up to 10 times per as a percentage. Every gene KD was repeated three times or more. Black
gene. Error bars represent the standard error. sections of columns represent melanized ookinete fraction. Error bars repre-
sent the standard error. (B) Representative sections of infected Anopheles
gambiae midguts infected with GFP-tagged P. berghei. Arrows point to
(16, 17), and oocysts were counted 6–8 days later. Silencing of melanized ookinetes.
REL2 (RHD dsRNA) or IMD resulted in a statistically significant
twofold increase of oocyst numbers, suggesting that the pathway is,
indeed, implicated in parasite killing before or during midgut the Imd pathway, which is activated mainly in response to Gram-
invasion (Fig. 6, Table 1; and see Table 5, which is published as negative bacteria, are highly conserved between the two insects.
supporting information on the PNAS web site). Quantitatively Diversification of proteins implicated in immune recognition
similar results were obtained when silencing REL2-F alone (ANK upstream of both pathways is very pronounced. This category
dsRNA), indicating that REL2-F, rather than REL2-S, is impli- includes peptidoglycan-recognition proteins (PGRPs) and Gram-
cated in this reaction. Both REL2-RHD and REL2-ANK, but not negative binding proteins. For example, the Drosophila PGRP-LC,
IMD KDs, led to limited occurrence of parasite melanization (⬇5% the main receptor of the Imd pathway (27, 28), and its Anopheles
of the ookinetes). This finding would suggest that REL2-F is orthologue have similar architectures, but their PGRP domains
involved in the control of the melanization reaction, in a manner have apparently evolved independently from a single domain-
independent of IMD. containing ancestral protein (11). Thus, recognition properties of
PGRP-LC could differ substantially between the two species.
Discussion Interestingly, the other Imd-pathway-associated receptor, PGRP-
Previously, our comparative genomic analysis of immunity-related LE, has no Anopheles orthologue.
genes in the fruit fly Drosophila melanogaster and the malaria Here, we provide evidence from detailed gene-structure deter-
mosquito Anopheles gambiae (11) revealed that, in general, intra- mination and reverse genetic analysis of function, illuminating the
cellular components of immune signaling pathways are well con- structural and functional evolution of immune signaling in Dro-
served between the two insects. However, differences were noted, sophila and Anopheles gambiae. The first major difference we
particularly in potential components of the Toll pathway, which may detected in Anopheles is alternative splicing of the REL2 gene. This
indicate important variations in immune signaling. Importantly, the gives rise to two isoforms, REL2-F and REL2-S, which are differ-
NF-␬B-like transcription factor Dif, which activates gene expression entially involved in defense against Gram-positive and Gram-
upon recognition of Gram-positive bacteria and fungi, has no negative bacteria, respectively. In this way, the single mosquito gene
orthologue in Anopheles. This may imply that in a putative Toll REL2 mediates alternative immune responses, which, in the fly,
pathway, one of the other two Anopheles NF-␬B-like proteins, require two genes: REL2’s orthologue Relish and Dif. Thus, through
REL1 (the Dorsal orthologue) or REL2 (the Relish orthologue), posttranscriptional processing of REL2, Anopheles appears to com-
may substitute for the role of Dif. Furthermore, Toll genes have pensate for the absence of Dif.
evolved through independent gene duplication and sequence di- In addition to this gene substitution, we discovered a switch in the
vergence, and, thus, the Drosophila Toll receptor has no clear roles of the NF-␬B transcriptional factors during the ⬇250 million
orthologue in Anopheles. In contrast, intracellular components of years of evolution (29, 30) since the last common ancestor of

11424 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0504950102 Meister et al.


Drosophila and Anopheles. The Anopheles REL2-F is used for vated, inhibited by another I␬B-like protein, or inhibited by
defense against the Gram-positive bacterium S. aureus, whereas the REL2-F through heterodimer formation. The detected low con-
orthologous and structurally similar Relish in the fruit fly is stitutive expression of the CEC1 promoter in Anopheles cells is
required for defense against Gram-negative bacteria. Likewise, caused by REL2-F and not REL2-S, suggesting that REL2-S is
REL2-S deals with Anopheles defense against the Gram-negative normally associated with an inhibitory protein. Further studies are
bacterium E. coli, whereas Dif, which, like REL2-S, does not have required to determine the inhibitory partner of REL2-S. Our
an inhibitory domain, deals with fruit fly responses against Gram- preliminary data suggest that the putative partner is not CACT, the
positive bacteria. However, both REL2-F and Relish mediate their orthologue of Cactus (data not shown).
functions in concert with the death-domain adaptor IMD, indicat- Anopheles gambiae is the main vector of human malaria in
ing that the pathway structure has remained conserved. subSaharan Africa, causing death to more than one million people
REL2-F, unlike its fly counterpart Relish but similar to ortholo- every year. An important finding of our work is that the IMD兾
gous genes in other higher eukaryotes, encodes a death domain REL2-F signaling cascade is involved in limiting the number of
located at the carboxyl terminus of the deduced protein. This P. berghei oocysts that develop in the mosquito midgut. This
domain might be used for oligomerization with the respective reaction could be due to the detected low constitutive activation of
domain of IMD or other death-domain-encoding proteins. Perti- REL2-F, to pathway activation by Plasmodium, or both. A recent
nent information is available from studies on the Relish gene of study has shown that ectopic overexpression of CEC1 drastically
Aedes aegypti, the vector of yellow and dengue fever, which belongs reduces susceptibility of mosquitoes to Plasmodium (33). Other
to a different mosquito subfamily (Culicinae rather than Anopheli- REL2-regulated proteins that might explain the observed pheno-
nae). The Relish gene of Aedes is structurally similar to Anopheles type are Gambicin (GAM1), another AMP with antiparasitic
REL2, in that it encodes a death domain. Aedes REL2 produces activity in vitro (34), and LRIM1, a key antagonist of ookinete-to-
three alternatively spliced transcripts: two are similar to Anopheles oocyst development (13).
REL2-F and REL2-S, and the third encodes a protein encompass- The silencing of REL2-F also leads to melanization of some
ing only the ANK domain (9).
ookinetes during their passage through the Anopheles midgut
In Aedes, transgenic overexpression of a truncated version of the

IMMUNOLOGY
epithelium. This reaction is independent of IMD, pointing to either
REL2-S homologue (C8), which lacks the putative glutamine- and
an alternative signaling cascade or to the detected REL2-F con-
histidine-rich transactivator domain, results in susceptibility to
stitutive activation. In Drosophila, the Imd pathway seems to
Gram-negative but not to Gram-positive bacteria (31). Additional
regulate melanization upstream of Serpin27A (35), which is an
studies are required to determine whether C8 acts as a dominant-
inhibitor of the final steps of prophenoloxidase activation and is
negative allele of REL2-S. In this case, our postulated switch in
pathway function would probably have occurred in the common controlled by the Toll pathway (36).
ancestor of anopheline and culicine mosquitoes. In conclusion, our data suggest significant divergence of immune
Whether Aedes, like Anophele, also lacks Dif is not yet resolved. signaling between the Anopheles mosquito and the fruit fly D. mela-
However, recent studies have shown that the Aedes orthologue of nogaster. The documented differences most likely reflect different
REL1 is not involved in defense against bacteria but only against lifestyles and, consequently, different infectious agents that the two
fungal infections (32). Likewise, our preliminary results in Anoph- insects encounter during their lifetimes. In mosquitoes, one of these
eles indicate that REL2 is the only NF-␬B factor implicated in the agents is the malaria parasite Plasmodium.
antibacterial defense against E. coli and S. aureus. Thus, it is likely
that the two mosquitoes use similar strategies to deal with bacterial We thank G. Dimopoulos and members of our and the Aksoy labora-
tories for helpful discussions and the use of equipment and Dr. Gareth
infections.
Lycett for critical reading of the manuscript. This work was supported,
Control of NF-␬B activation is achieved through inhibitory I␬B in part, by National Institutes of Health Grants R01 A43035 (to L.Z.),
domains by the use of two alternative strategies. In the fly, Relish P01 AI044220-06A1 (to F.C.K.), and K22 HG00045 (to K.P.W.); by
contains its own inhibitory I␬B domain (ANK) located at the European Union Grant HPRN-CT-2000-00080 (to F.C.K.); and by the
carboxyl terminus of the protein, whereas the inhibitory domain for European Molecular Biology Laboratory. S.M.K. was supported, in part,
Dif is provided by another protein, Cactus. Accordingly, Anopheles by the Deutsche Forschungsgemeinschaft, and N.T.H. was supported by
REL2-S, in noninduced condition, might be constitutively acti- Wellcome Trust Fellowship 064874.

1. Hoffmann, J. A. & Reichhart, J. M. (2002) Nat. Immunol. 3, 121–126. 19. Hegde, P., Qi, R., Abernathy, K., Gay, C., Dharap, S., Gaspard, R., Hughes, J. E., Snesrud,
2. Lemaitre, B., Nicolas, E., Michaut, L., Reichhart, J. M. & Hoffmann, J. A. (1996) Cell 86, 973–983. E., Lee, N. & Quackenbush, J. (2000) BioTechniques 29, 548–550, 552–554, and 556.
3. Michel, T., Reichhart, J. M., Hoffmann, J. A. & Royet, J. (2001) Nature 414, 756–759. 20. Leulier, F., Parquet, C., Pili-Floury, S., Ryu, J. H., Caroff, M., Lee, W. J., Mengin-Lecreulx,
4. Meng, X., Khanuja, B. S. & Ip, Y. T. (1999) Genes Dev. 13, 792–797. D. & Lemaitre, B. (2003) Nat. Immunol. 4, 478–484.
5. Rutschmann, S., Jung, A. C., Hetru, C., Reichhart, J. M., Hoffmann, J. A. & Ferrandon, D. 21. Vizioli, J., Bulet, P., Charlet, M., Lowenberger, C., Blass, C., Muller, H. M., Dimopoulos,
(2000) Immunity 12, 569–580. G., Hoffmann, J., Kafatos, F. C. & Richman, A. (2000) Insect Mol. Biol. 9, 75–84.
6. Lemaitre, B., Kromer-Metzger, E., Michaut, L., Nicolas, E., Meister, M., Georgel, P., 22. Hoa, N. T., Keene, K. M., Olson, K. E. & Zheng, L. (2003) Insect Biochem. Mol. Biol. 33,
Reichhart, J. M. & Hoffmann, J. A. (1995) Proc. Natl. Acad. Sci. USA 92, 9465–9469. 949–957.
7. Silverman, N., Zhou, R., Stoven, S., Pandey, N., Hultmark, D. & Maniatis, T. (2000) Genes 23. Onfelt Tingvall, T., Roos, E. & Engstrom, Y. (2001) EMBO Rep. 2, 239–243.
Dev. 14, 2461–2471. 24. Dimopoulos, G., Casavant, T. L., Chang, S., Scheetz, T., Roberts, C., Donohue, M., Schultz,
8. Stoven, S., Silverman, N., Junell, A., Hedengren-Olcott, M., Erturk, D., Engstrom, Y., J., Benes, V., Bork, P., Ansorge, W., et al. (2000) Proc. Natl. Acad. Sci. USA 97, 6619–6624.
Maniatis, T. & Hultmark, D. (2003) Proc. Natl. Acad. Sci. USA 100, 5991–5996. 25. Nachou, D., Schlegelmilch, T., Christophides, G. K. & Kafatos, F. C. (2005) Curr. Biol. 15,
9. Shin, S. W., Kokoza, V., Ahmed, A. & Raikhel, A. S. (2002) Proc. Natl. Acad. Sci. USA 99, 9978–9983. 1185–1195.
26. Dimopoulos, G., Richman, A., Muller, H. M. & Kafatos, F. C. (1997) Proc. Natl. Acad. Sci.
10. Holt, R. A., Subramanian, G. M., Halpern, A., Sutton, G. G., Charlab, R., Nusskern, D. R.,
USA 94, 11508–11513.
Wincker, P., Clark, A. G., Ribeiro, J. M., Wides, R., et al. (2002) Science 298, 129–149.
27. Choe, K. M., Werner, T., Stoven, S., Hultmark, D. & Anderson, K. V. (2002) Science 296, 359–362.
11. Christophides, G. K., Zdobnov, E., Barillas-Mury, C., Birney, E., Blandin, S., Blass, C., Brey,
28. Gottar, M., Gobert, V., Michel, T., Belvin, M., Duyk, G., Hoffmann, J. A., Ferrandon, D.
P. T., Collins, F. H., Danielli, A., Dimopoulos, G., et al. (2002) Science 298, 159–165.
& Royet, J. (2002) Nature 416, 640–644.
12. Barillas-Mury, C., Charlesworth, A., Gross, I., Richman, A., Hoffmann, J. A. & Kafatos, 29. Yeates, D. K. & Wiegmann, B. M. (1999) Annu. Rev. Entomol. 44, 397–428.
F. C. (1996) EMBO J. 15, 4691–4701. 30. Gaunt, M. W. & Miles, M. A. (2002) Mol. Biol. Evol. 19, 748–761.
13. Osta, M. A., Christophides, G. K. & Kafatos, F. C. (2004) Science 303, 2030–2032. 31. Shin, S. W., Kokoza, V., Lobkov, I. & Raikhel, A. S. (2003) Proc. Natl. Acad. Sci. USA 100, 2616–2621.
14. Zheng, X. L. & Zheng, A. L. (2002) Insect Mol. Biol. 11, 517–525. 32. Shin, S. W., Kokoza, V., Bian, G., Cheon, H. M., Kim, Y. J. & Raikhel, A. S. (2005) J. Biol.
15. Blandin, S., Moita, L. F., Kocher, T., Wilm, M., Kafatos, F. C. & Levashina, E. A. (2002) Chem. 280, 16499–16507.
EMBO Rep. 3, 852–856. 33. Kim, W., Koo, H., Richman, A. M., Seeley, D., Vizioli, J., Klocko, A. D. & O’Brochta, D. A.
16. Vlachou, D., Zimmermann, T., Cantera, R., Janse, C. J., Waters, A. P. & Kafatos, F. C. (2004) J. Med. Entomol. 41, 447–455.
(2004) Cell. Microbiol. 6, 671–685. 34. Vizioli, J., Bulet, P., Hoffmann, J. A., Kafatos, F. C., Muller, H. M. & Dimopoulos, G. (2001)
17. Franke-Fayard, B., Trueman, H., Ramesar, J., Mendoza, J., van der Keur, M., van der Linden, Proc. Natl. Acad. Sci. USA 98, 12630–12635.
R., Sinden, R. E., Waters, A. P. & Janse, C. J. (2004) Mol. Biochem. Parasitol. 137, 23–33. 35. Takehana, A., Yano, T., Mita, S., Kotani, A., Oshima, Y. & Kurata, S. (2004) EMBO J. 23, 4690–4700.
18. Dimopoulos, G., Christophides, G. K., Meister, S., Schultz, J., White, K. P., Barillas-Mury, 36. Ligoxygakis, P., Pelte, N., Ji, C., Leclerc, V., Duvic, B., Belvin, M., Jiang, H., Hoffmann, J. A.
C. & Kafatos, F. C. (2002) Proc. Natl. Acad. Sci. USA 99, 8814–8819. & Reichhart, J. M. (2002) EMBO J. 21, 6330–6337.

Meister et al. PNAS 兩 August 9, 2005 兩 vol. 102 兩 no. 32 兩 11425

You might also like