Download Functional study of hemolymph coagulation in Zhi Wang Drosophila

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Plant disease resistance wikipedia , lookup

Herd immunity wikipedia , lookup

Phagocyte wikipedia , lookup

Adoptive cell transfer wikipedia , lookup

DNA vaccination wikipedia , lookup

Infection wikipedia , lookup

Infection control wikipedia , lookup

Hospital-acquired infection wikipedia , lookup

Neonatal infection wikipedia , lookup

Complement system wikipedia , lookup

Sociality and disease transmission wikipedia , lookup

Molecular mimicry wikipedia , lookup

Cancer immunotherapy wikipedia , lookup

Adaptive immune system wikipedia , lookup

Social immunity wikipedia , lookup

Immune system wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Immunosuppressive drug wikipedia , lookup

Hygiene hypothesis wikipedia , lookup

Drosophila melanogaster wikipedia , lookup

Innate immune system wikipedia , lookup

Immunomics wikipedia , lookup

Psychoneuroimmunology wikipedia , lookup

Transcript
Functional study of hemolymph coagulation in
Drosophila larvae
Zhi Wang
Doctoral thesis
Stockholm 2012
Department of Molecular Biology and Functional Genomics
Stockholm University
i
©Zhi Wang, Stockholm 2012
ISBN 978-91-7447-501-2
Printed in Sweden by Universitetsservice US-AB,
Stockholm 2012
Distributor: Stockholm University Library
ii
Abstract
Many pathogen infections in nature are accompanied by injury and subsequent coagulation.
Despite the contribution of hemolymph coagulation to wound sealing, little is known about its
immune function. Based on the molecular knowledge of Drosophila innate immunity, this
thesis investigated the immune function of clot both in vitro and in vivo, the immune relevant
genes involved in a natural infection model, involving entomopathogenic nematodes (EPN)
and the factors leading to crystal cell activation.
Transglutaminase (TG) and its substrate Fondue (Fon) have been identified as bona fide clot
components in Drosophila larvae. By knocking down TG or Fon via RNAi, we observed an
increased susceptibility to EPN in larvae. In addition, this increased susceptibility was
associated with an impaired ability of hemolymph clots to entrap bacteria. Immunostaining
revealed that both clot components (Fon and TG) were able to target microbial surfaces. All
these data suggest an immune function for the Drosophila hemolymph clot. Strikingly,
similar results were obtained when we ran parallel experiments with human FXIIIa, an
ortholog of Drosophila TG, indicating a functional conservation. We also found evidence for
the regulation on both clot and immunity by eicosanoids in Drosophila larvae.
The combination of EPN infection with the Drosophila model system allowed us to discover
an immune function for TEP3 and Glutactin. However the molecular mechanism underlying
the involvement of these two proteins in this particular host-pathogen interaction remains to
be elucidated.
Prophenoloxidase, the pro-form of enzyme involved in hardening the clot matrix, has been
shown to be released by rupture of crystal cells. This cell rupture is dependent on activation of
the JNK pathway, Rho GTPases and Eiger. Our work further identified the cytoskeletal
component, Moesin, and the cytoskeletal regulator Rac2 as mediators of cell rupture. Despite
the possible role of caspases in crystal cell activation, such cell rupture was turned out to be
different from apoptosis. The implication of Rab5 in this process indicated that proper
endocytosis is required for cell activation and subsequent melanization. Our findings
furthered not only our understanding of the release of proPO via cell rupture but also our
knowledge on different paths of immune cell activation.
III
List of papers
This thesis is based on the following publications and manuscript, which are referred to in the
text by Roman numerals. Some additional results are presented.
I.
Pavel Dobes, Zhi Wang, Robert Markus, Ulrich Theopold, Pavel Hyrsl. An improved
method for nematode infection assays in Drosophila larvae. (in press in Journal Fly)
II.
Wang Z*, Wilhelmsson C*, Hyrsl P*, Loof TG*, Dobes P, Klupp M, Loseva O,
Mörgelin M, Iklé J, Cripps RM, Herwald H, Theopold U. Pathogen entrapment by
transglutaminase--a conserved early innate immune mechanism. PLoS Pathog. 2010
Feb 12;6(2):e1000763.
III. Hyrsl P, Dobes P, Wang Z, Hauling T, Wilhelmsson C, Theopold U. Clotting
factors and eicosanoids protect against nematode infections. J Innate Immun.
2011;3(1):65-70. Epub 2010 Oct 16.
IV. Zhi Wang, Gawa Bidla, Robert Markus and Ulrich Theopold . Crystal cell activation
through rupture. (Manuscript)
*These author contributed equally to the work
Paper not included in the thesis
Waldholm J, Wang Z, Brodin D, Tyagi A, Yu S, Theopold U, Farrants AK, Visa N. SWI/SNF
regulates the alternative processing of a specific subset of pre-mRNAs in Drosophila
melanogaster. BMC Mol Biol. 2011 Nov 2;12:46.
Reprints were made with permission from the publishers.
IV
Abbreviations
AA
Arachidonic acid
AMP
Antimicrobial peptide
Bsk
Basket
Crq
Croquemort
COX
Cyclooxygenase
Cyt
Cytochrome
DAMP
Damage-associated molecular pattern
DAP-type
meso-diaminopimelic acid-type
DCV
Drosophila C virus
DSCAM
Down syndrome cell adhesion molecule
dSR-CI
Drosophila scavenger receptor- CI
EGF
Epidermal growth factor
EPN
Entomopathogenic nematode
Fon
Fondue
GFP
Green fluorescent protein
Glu
Glutamine
GNBP
Gram-negative binding protein
GRP
Glucanases recognition protein
GTPase
Guanosine triphosphate hydrolase enzyme
Hml
Hemolectin
Ig
Immunoglobulin
IJs
Infective juveniles
Imd
Immunodeficiency
JAK
Janus kinases
JNK
Jun N-terminal kinase
LOX
Lipoxygenase
LPS
Lipopolysaccharide
LTA
Lipoteichoic acid
Lys
Lysine
MP1/MP2
Melanization protein 1 and 2, respectively
NF-kB
Nuclear factor-kappa B
PAF
Platelet activating factor
PAMP
Pathogen-associated molecular pattern
PGN
Peptidoglycan
V
PGRP
Peptidoglycan recognition protein
PLA2
Phospholipase A2
PM
Peritrophic matrix
PO
Phenoloxidase
PPAS
Prophenoloxidase activating system
proPO
Prophenoloxidase
PRR
Pattern recognition receptor
Rel
Relish
Rho
Ras homology
RNAi
Ribonucleic acid (RNA) interference
RNS
Reactive nitrogen species
ROS
Reactive oxygen species
Serpin
Serine protease inhibitor
TEP
Thioester containing protein
TF
Tissue factor
TG
Transglutaminase
TLR
Toll-like receptor
TNF
Tumor necrosis factor
Tcs
Toxin complex
Wt
Wild type
VI
Contents
Introduction ............................................................................................................................... 9
Insect models to study innate immunity .............................................................................. 10
Insect pathogens and entomopathogenic nematode (EPN) infection .............................. 10
Insect pathogens .......................................................................................................... 10
Entomopathogenic nematode infection ....................................................................... 11
The immune system of insects ........................................................................................ 13
Sensing and recognition – the first step of immune activation.................................... 14
Recognition receptors/molecules............................................................................. 14
Danger signals ......................................................................................................... 17
Defense mechanisms-elimination of invaders ............................................................. 17
Phagocytosis and encapsulation .............................................................................. 17
Local and systemic induction of AMPs................................................................... 18
Melanization and Coagulation ................................................................................. 19
Eicosanoids and immunity .......................................................................................... 23
The present study..................................................................................................................... 25
Aim of this study ................................................................................................................. 25
Results and discussion ......................................................................................................... 26
Improvement of a natural entomopathogenic nematode infection in Drosophila larvae
(Paper I) ........................................................................................................................... 26
Immune function of Drosophila clot in vivo (Paper II and paper III) ............................. 27
Recognition Molecules and immune related genes involved in EPN infection in
Drosophila larvae ............................................................................................................ 29
Crystal cell activation via rupture (Paper IV) .................................................................. 33
Concluding remarks ................................................................................................................ 36
Acknowledgements ................................................................................................................. 39
References ............................................................................................................................... 41
7
8
Introduction
Multicellular organisms, including vertebrates, invertebrates and plants, are
surrounded by a great variety of infectious agents including viruses, bacteria, fungi,
protozoa and multicellular parasites. These infectious agents are usually prevented
from entering the host body by a combination of physical, chemical and biochemical
barriers. Once the barriers are breached by penetrating or through injuries, the
infectious agents have the potential to multiply unchecked in their host bodies and
cause disease, eventually killing their hosts. However, most of the infections are
eliminated and leave little permanent damage. This is due to the immune system,
which surveys the whole host body and combats infections.
Infectious microbes differ in size, lifestyle and the way they induce infection, so a
wide variety of immune responses are required for the host to cope with all kinds of
infections. Any immune response consists of two phases: recognition and elimination.
Broadly, the immune responses fall into two categories: innate (or non-adaptive)
immune responses and adaptive immune responses. The innate immune system is an
ancient form of host defense found in all multicellular organisms (reviewed in
(Medzhitov & Janeway 1998; Janeway & Medzhitov 2002)). It relies on germ-line
encoded receptors that recognize conserved molecular patterns associated with
pathogens (PAMPs for pathogen associated molecular patterns) as non-self (Janeway
1992). This interaction between the host receptors and PAMPs directly induces host
defense responses, such as phagocytosis and induction of antimicrobial peptides and
reactive oxygen species (ROS) (Janeway 1992), which constitute primary defense
mechanisms against infections.
In addition to innate immune responses, vertebrates also rely on their adaptive
immune responses, which are mainly mediated by lymphocytes. There are
considerable interactions between these two systems. The activated innate immune
system releases proinflammatory molecules like cytokines to guide and turn on a
series of adaptive immune responses, including lymphocyte division, maturation and
antibody secretion (reviewed in (Hoebe et al. 2004)). Unlike the innate immune
system, whose recognition receptors are constitutively expressed, the adaptive
immune system depends on specific antigen receptors created by a somatic
mechanism of gene rearrangement and clonal selection to recognize pathogens. The
clonal selection offers vertebrates exquisite specificity and memory, the ability to
rapid response when re-exposed to the same pathogen. However, this takes time,
which makes adaptive immune responses less effective during the initial defense
against infections than innate immune responses (reviewed in (Dushay & Eldon
1998)).
The Immune system is complex. Although many the microbes are recognized as nonself by hosts, not all of them are eliminated by host immune system. For instance, the
bacteria inhabiting the animal gut are not entirely eliminated and rather maintained
9
and regulated by the host despite the low induction of antimicrobial peptides in the
gut epithelia (Ryu et al. 2008). In some cases, host cells are removed by activating the
host immune system when they undergo apoptosis or cancer and are recognized as
altered self (reviewed in (Miyake & Yamasaki 2012)). Thus, the immune system is
not only a means to combat infectious agents but also the mechanism to maintain
homeostasis (Lazzaro & Rolff 2011).
Insect models to study innate immunity
Insects, like other invertebrates, rely entirely on innate immune responses due to the
lack of adaptive immune system. Therefore the study of the interactions of insects
with microbes provides a powerful tool for discovering early immune reactions
without interference with adaptive responses that are present in vertebrate systems.
The study of insect immunity dates back to the late 19 century when Louis Pasteur
initiated the study of silkworm disease. Since then the study of infection and
immunity in insects has grown and contributed greatly to an increasing knowledge of
innate immunity for both invertebrates and vertebrates. The first concept of an
inducible antibacterial defense was introduced to the whole immunity field in the
1970s by the pioneering studies in Drosophila melanogaster (Boman et al. 1972).
Subsequently, the primary structure of the first antimicrobial peptide (AMP),
designated cecropin, was identified in the giant silk moth Hyalophora cecropia
(Steiner et al. 1981). This has led to the blooming of the biochemical characterization
of antimicrobial peptides and other immune proteins.
Due to the conserved nature of many fundamental aspects of the immune system
across all animals, insects have served as valuable models for the study of immune
system function and the interaction between pathogen and host. The best example is
the discovery of the Toll signaling pathway in Drosophila in 1990s (Lemaitre et al.
1996) , which paved the way for the subsequent identification of the Toll-like receptor
pathway in mammals (Poltorak et al. 1998). With this work earning a share of the
2011 Nobel Prize in Physiology or Medicine, the importance of Drosophila as a
model system was further emphasized.
Insect pathogens and entomopathogenic nematode (EPN) infection
Insect pathogens
Insects are infected by an incredibly large number and diversity of pathogen species.
The different types of pathogen include viruses, bacteria, fungi, protozoans,
nematodes and parasitoids. They differ in size, lifestyle, and the way of causing
infection; even the induced host defense responses are different depending on
pathogen species.
10
Insect viruses may be double- or single-stranded DNA (dsDNA and ssDNA,
respectively) or double- or single-stranded RNA. In Drosophila, viruses are important
natural pathogens and can be transmitted horizontally such as Drosophila C Virus
(DCV) or vertically like Sigma virus (reviewed in (Lemaitre & Hoffmann 2007)).
They are eradicated by the host via the RNA interference machinery (van Mierlo et al.
2011). Most bacteria gain the entry into their insect hosts by injuries or ingestion.
Some bacterial species get into the insect host via special vector-nematodes. Such
bacteria are usually associated with their specific nematode species forming a
symbiosis. Bacterial endosymbionts like Wolbachia infect their hosts by vertical
transmission from a female host to her offspring. In the host body, bacteria usually
induce immune responses like the production of antimicrobial peptides or
phagocytosis, resulting in clearance of the bacteria. Note that some bacteria such as
Mycobacterium marinum benefit from being phagocytosed because the immune cell
provides a temporal shelter for them preventing contact with other effector molecules
(Dionne et al. 2003).
Unlike viruses and bacteria, entomopathogenic fungi almost always invade their
insect hosts by penetrating directly through the cuticle, and eventually the infecting
hyphae reach the insect hemocoel. Beauveria bassiana is a widely used fungus in the
study of insect immune responses to fungi. Recently an opportunistic human pathogen,
Canida albicans has been applied to infect Drosophila (Davis et al. 2011; Glittenberg
et al. 2011).
Protozoans are also among animal enemies. Some of them are responsible for serious
human diseases, such as malaria, vectored by mosquitoes; and some of them are
specific to and causing diseases in insects. The microsporidia, one group of insect
protozoans, are mostly investigated as insecticides. They are ingested by insect host
and infect host gut epithelial cells, resulting in the debilitation and eventual death of
the host (Texier et al. 2010).
In nature, parasitoid wasps parasitize mainly insects. They oviposit into other insects’
host bodies. The suppression of the host defense aids the wasp offspring to complete
their development, and the mature adult wasps emerge from the host pupal case, mate
and disperse for another infection (Carton & Nappi 2001).
Because some stages of insect life cycle are in a humid environment, there is a
considerable risk for contact with nematodes. The parasitic nematodes are also
natural pathogens of insects, known as entomopathogenic nematodes (EPNs).
Entomopathogenic nematode infection
Entomopathogenic nematodes (EPNs) are soil-dwelling parasitic worms that infect
and kill insects (Bednarek & Gaugler 1997). In the life cycle (See figure 1), the nonfeeding and developmentally arrested infective juvenile (IJ) is the only stage of EPN
found outside of insect host in nature. Infective juveniles (IJs) are tightly associated
11
with their specific symbiotic bacteria inhabiting their intestinal lumen (Poinar 1977).
During an infection, IJs enter the insect host through their natural openings such as
spiracles, mouth or by penetrating the cuticle, and release the symbiotic bacteria that
kill and digest host tissues, thereby providing nutrients for the development of both
nematodes and bacteria. There is evidence to show that the symbiotic bacteria are
essential to nematodes, since axenic nematodes do not cause insect mortality and do
not proliferate inside the insect cadavers (Han & Ehlers 2000). The cues released by
insect hosts, such as CO2, are sensed by IJs and guide them to find the host, indicating
that IJs have evolved chemosensory mechanisms not only to detect insect hosts, but
also to locate the hosts (Ciche 2007; Hallem et al. 2011).
Figure 1. Live cycle of entomopathogenic nematode. Infective juvenile nematodes enter the
insect body through natural opening or by injury (right part of figure), then they release their
symbiotic bacteria (green dots in lower part) to kill the host. Nematodes reproduce and
produce offspring for 2 or 3 generations (left part). Infective juveniles leave the dead insect
and seek a new insect host (upper part).
Although EPNs have long been widely applied as biologic control agents for insect
pests, particularly the species in the genera Steinernema and Heterorhabditis
(Steinernematidae and Heterorhabditae, respectively) (Ehlers 2001), the interactions
between EPNs and their hosts, particularly at the level of the host immune response
are not fully understood. Early studies on the interaction between Hyalophora
cecropia and Neoaplectana carpocapsae/Xenorhabdus nematophilus have found that
the nematodes helped the bacteria by excreting an immune inhibitor that selectively
destroys immune proteins P5 and P9 (P. Gotz et al. 1981). In Manduca sexta, the
nematode Heterorhabditis bacteriophora has been observed to be encapsulated
without being (Li 2007). Recently, the EPN infection of Drosophila melanogaster has
been performed by Hallem et al., who found that the infection of Drosophila larvae
with Heterorhabditis/Photorhabdus resulted in the temporary dynamic expression of
AMP genes, a response that was induced specifically by the symbiotic bacterium
Photorhabdus. However, Drosophila larvae with defects in the induction of AMPs
12
show the same sensitivity towards EPNs as controls, indicating a dispensable role of
AMPs and in fact of both the Imd and Toll pathway in this infection model (Hallem et
al. 2007). Accumulating empirical evidences show how the two partners of EPN
symbiosis evade or suppress host immune system to achieve a successful infection of
insects.
The immune system of insects
Insects are constantly exposed to various pathogens, but only a few invaders lead to
infections. This is due to a robust and efficient immune system (Figure 2).
Figure 2. Schematic overview of insect immune system. The recognition of pathogens
elicits a large array of immune responses, including cellular immune responses and humoral
effector mechanisms (see text for more details).
Insect immune tissues comprise immunocytes, lymph gland (the “factory” of
immunocytes), epithelia, and fat body which is equivalent to the mammalian liver.
Drosophila produces three types of blood cells (hemocytes): plasmatocytes, crystal
cells and lamellocytes (reviewed in (Crozatier & Meister 2007)). Plasmatocytes are
the majority in the blood or hemolymph and account for 95% of total hemocytes.
They are macrophage-like phagocytes and found in all life stages. Crystal cells are
only found at embryonic and larval stages comprising 5% of the blood cell population.
With crystallized Prophenoloxidase (PPO), the pro-enzyme required in melanin
production, crystal cells are the main player in melanization. Lamellocytes
differentiate in the lymph gland or from sessile hemocytes at the larval stage upon
immune challenge by parasites or melanotic tumors (Sorrentino et al. 2002; Markus et
al. 2009). As in vertebrates, Drosophila blood cells also have two developmental
13
origins: one early in the embryonic head mesoderm and another during larval
development in the lymph gland (Holz et al. 2003; Crozatier & Meister 2007).
The first line of defense comprises the integument, the peritrophic matrix (PM), a
layer composed of chitin and glycoproteins that lines the insect intestinal lumen, and
the epithelia covering the outer and inner surfaces. For entry into the host body cavity
(hemocoel), pathogens have to breach these barriers. Once in the host body, they are
immediately recognized, leading to the activation of host immune cells and tissues
and effective immune responses. The binding of plasmatocytes to the invading
pathogens through the cell surface recognition receptors leads to the phagocytosis of
microbes. Pathogens too big to be engulfed by plasmatocytes like wasp eggs, activate
both plasmatocytes and the differentiation of lamellocytes and are eliminated by
formation of multicellular layered capsules. In addition to the recognition of
pathogens, even signals from damaged host cell can turn on humoral immunity,
including melanization, coagulation and production of AMPs (reviewed in (Lemaitre
& Hoffmann 2007)). Rather than being independent, immune responses are often
interconnected. Some evidences have shown that the depletion of phagocytic
plasmatocytes in Drosophila larvae largely increased the susceptibility of host to
bacteria (Basset et al. 2000; Brennan et al. 2007). Thus, cellular effectors and humoral
effectors work in synergy to mount a successful defense against pathogens.
Sensing and recognition – the first step of immune activation
Each immune reaction is initiated by either the recognition of a pathogen or sensing
the ‘danger’ due to the presence of pathogens or both. Thus the identification of these
cues is essential to understand immunity and has been actively investigated.
Recognition receptors/molecules
The recognition of pathogen is achieved by the interaction of host recognition
molecules with microbial conserved patterns, so called PAMPs which are only found
on evolutionarily distant organisms like bacteria and absent on eukaryotes (Janeway
1989). PAMPs are essential components of microbial cell wall, including
peptidoglycan (PGN) found in both Gram-positive and Gram-negative bacteria,
lipopolysaccharide (LPS) from the outer membrane of Gram-negative bacteria,
lipoteichoic acid (LTA) of Gram-positive bacteria and fungal β 1, 3-glucans
(reviewed in (Royet 2004; Charroux et al. 2009)). The host receptors capable of
recognizing or interacting with them are collectively called pattern recognition
receptors (PRRs).
Peptidoglycan consists of long glycan chains crosslinked to each other by short
peptide bridges which can be separated from glycan chain by cleavage of amidases
(reviewed in (Chaput & Boneca 2007)). Peptidoglycan from Gram-negative bacteria
differs from most Gram-positive PGN by the replacement of lysine (Lys) with mesodiaminopimelic acid (DAP) at the third position in the peptide bridge (Schleifer &
14
Kandler 1972), which allows host to differentiate different type of bacteria (Paredes et
al. 2011). Peptidoglycan recognition proteins (PGRPs) have been shown to interact
with PGNs in a wide range of species from insects to mammals, revealing the
evolutionary conservation in this family proteins (Yoshida et al. 1996; Sang et al.
2005).
Drosophila contains 13 PGRPs. Some of them lack amidase activity, such as PGRPSA,-SD, -LC, -LE, -LF. However their conserved PGRP domain is responsible for
sensing of bacterial elicitors upstream of the Toll and Imd pathways, which regulate
the production of AMPs (reviewed in (Ferrandon et al. 2007; Royet et al. 2011).
Study of loss-of-function mutants in PGRP-SA known as Semmelweis reveals its role
in activation of the Toll pathway in response to Gram-positive bacteria, but not to
fungal infection (Michel et al. 2001). PGRP-LC and –LE function synergistically in
the activation of the Imd pathway, since PGRP-LC and –LE (loss of function) double
mutants show a more dramatic susceptibility to Gram-negative bacterial infection than
either mutation alone(Takehana et al. 2004). In addition to its critical role in the
induction of AMP, PGRP-LC is also involved in the microbial recognition in
phagocytosis, as its depletion by RNA interference (RNAi) decreases the
phagocytosis of E.coli but not S.aureus in S2 cells (Ramet et al. 2002). Other
essential function have been attributed to PGRP-LE namely recognition of
intracellular Listeria monocytogenes and activation of proPO cascade (Takehana et al.
2004; Yano et al. 2008). Recently, the discovery of the negative regulation of the
Imd and JNK pathway by PGRP-LF in flies suggests that it has a function similar to
decoy receptors for mammalian cytokine receptors (Maillet et al. 2008; Charroux et al.
2009).
The Drosophila genome encodes six catalytic PGRPs (PGRP-SC1A, -SC1B, -SC2, LB, -SB1, and –SB2). PGRP–SB1 has been reported to show the amidase activity
towards DAP-type PGN (Mellroth & Steiner 2006). Using deletion of six catalytic
PGRPs separately or in combination, Lemaitre’s group has uncovered the role of
PGRP-LB as a negative regulator of the Imd pathway and a synergy of PGRP-SCs
with –LB in the systemic response (Paredes et al. 2011).
Gram-negative binding proteins (GNBPs) and β 1, 3-glucanases recognition protein
(βGRP) form another family of PRR. They share sequence similarities with bacterial
β 1, 3-glucanases but loss of the enzymatic activity due to the amino-acid
substitutions in residues important for catalytic cleavage. Like non-catalytic PGRPs,
they retain only the ability to bind to β 1, 3-glucans. (Yahata et al. 1990; Royet 2004).
A mutation in the Gram-negative binding protein 1 (GNBP1) gene leads to
compromised survival of mutant flies after Gram-positive infections, but not after
fungal or Gram-negative bacteria challenge. This demonstrates that GNBP1 and
PGRP-SA can jointly activate the Toll pathway (Gobert et al. 2003). The detection of
fungal infections relies on both the binding of GNBP3 to fungal cell wall and on
cleavage of Persephone by the fungal virulence factor PR1 (Gottar et al. 2006).
15
Like PGRPs, scavenger receptors are conserved across the animal kingdom. It is now
becoming clearer that they recognize not only PAMPs but also endogenous ligands
from damaged self cells (reviewed in (Pal & Wu 2009)) . In Drosophila, DScr-C1, a
homologue of mammalian class C macrophage specific scavenger receptor (SR-C),
shows a restricted expression on macrophages or hemocytes during embryonic
development and is important for phagocytosis of bacteria (Pearson et al. 1995).
Similar to DScr-C1, Croquemort (Crq) is required for phagocytosis by embryonic
hemocytes, not of bacteria instead of dying cells (Franc et al. 1999). In mammals,
macrophage receptor CD36 has been implicated in recognizing and internalizing
Gram-positive bacteria, but its Drosophila homologue, Croquemort (Crq) has so far
only been shown to identify and remove apoptotic cells in vivo (Franc et al. 1999;
Stuart et al. 2005). In some cases, receptors show exclusive specificity for one type of
bacteria, for example Peste, which promotes exclusively the phagocytosis of
Mycobaterium fortuitum (Philips et al. 2005). Eater and NimrodC1, both possessing
an epidermal growth factor-like motif (EGF-like) are new pattern recognition
receptors. They are found on the plasmatocytes surface and play an active role in the
removal of apoptotic cells and microorganisms (Kocks et al. 2005; Kurucz et al.
2007).
Thioester containing proteins (TEPs) are highly conserved. Due to the innate immune
function of vertebrate complement proteins, belonging to TEPs, insect TEPs are
thought to be involved in innate immunity. This has been corroborated by studies in
the mosquito Anopheles gambiae where TEP1 (aTEP1) was shown to act as a bona
fide opsonin to promote phagocytosis of bacteria (Levashina et al. 2001). The
Drosophila genome encodes 6 TEPs. Except TEP5, which does not appear to be
expressed, other TEPs have been actively studied in the context of immunity. So far,
evidence only showed an up-regulation of TEPs expression upon immune challenge
in Drosophila larvae and adult; however mutants in TEPs survived different bacterial
infection as well as wild type flies. One possibility is a redundancy in the activity of
TEPs (Bou Aoun et al. 2011) .
Lectins are carbohydrate binding proteins and proposed to participate in innate
immunity. Some Drosophila C-type lectins (DL2 and DL3) have been identified to
agglutinate Gram-negative E. coli to hemocytes in a calcium-dependent manner; and
agarose beads showed enhanced encapsulation and melanization by Drosophila
hemocytes in vitro after coating with the lectins (Ao et al. 2007). An interesting
finding made early highlights the ability to induce cecropin A1 with a hemagglutinin
from edible snail in the Drosophila blood cell-line mbn-2 (Theopold et al. 1996).
Down Syndrome Cell Adhesion Molecule (Dscam) is composed of several variable
exons flanked by constant exons, which can potentially generate more than 18,000
splice variants of immunoglobulin (Ig)-superfamily receptors (Schmucker et al. 2000).
A possible immune function for Dscam was discovered by Watson et al. in 2005. The
authors proposed a role of Dscam in phagocytosis based on the observation that
16
hemocyte-specific loss of Dscam impairs the efficiency of phagocytic uptake of
bacteria, possibly due to reduced bacterial binding (Watson et al. 2005).
Danger signals
In the absence of pathogens, the host immune system is also turned on such as in the
case of an aseptic injury. This indicates that not only exogenous pathogens but also
endogenous molecules released from damaged self, so called damage-associated
molecular patterns (DAMPs) are capable of activating the immune system (reviewed
in (Bianchi 2007; Miyake & Yamasaki 2012)). Examples of danger signals are heat
shock proteins, reactive oxygen intermediates and extracellular-matrix breakdown
products. For example, encapsulation of salivary glands is induced when
tumorigenesis occurs by overexpressing of the Ras oncogen in glands in Drosophila
larvae (unpublished data from Thomas Hauling). In flies, the turandot (Tot) peptide A
are upregulated in fat body and secreted into the hemolymph under severe stress such
as high temperature and exposure to oxidative agents (Ekengren et al. 2001). TotA
has also been shown to be induced by bacterial challenge (Agaisse et al. 2003).
Studies in Galleria mellonella where released collagen fragments and nucleic acid
from damaged cell synergize with PAMPs to stimulate an immune response adds
additional support to the idea that danger signals play a part in host recognition.
It is becoming more accepted that both exogenous signals (PAMPs) and endogenous
signals (DAMPs) work together to alarm the host for appropriate immune responses
and to maintain host homeostasis.
Defense mechanisms-elimination of invaders
Phagocytosis and encapsulation
Phagocytosis, thought to have evolved originally from engulfment of nutrients, is a
primary defense mechanism highly conserved across the animal kingdom (reviewed
in (Stuart & Ezekowitz 2005)). It is mediated by phagocytic cells, plasmatocytes in
Drosophila. Phagocytic receptors on the cell surface (see above) recognize the
invading microorganisms or apoptotic cells, followed by cell cytoskeleton remodeling.
The ingested particles (phagosomes) fuse with lysosomes to form phagolysosome and
finally are degraded. It is noteworthy that despite the increasing list of identified
phagocytic receptors, Eater and NimC1 are the only ones for which convincing
evidence for their activity has been provided (Kocks et al. 2005; Kurucz et al. 2007).
Bacterial pathogens have devised numerous and diverse strategies to overcome
phagocytosis. Most are aimed at blocking one or more of the steps in phagocytosis,
thereby halting the process. In EPN infection, the symbiotic bacteria Photorhabdus
luminescens suppress phagocytosis by secreting toxin complex (Tcs) to alternate the
actin cytoskeleton (reviewed in (Eleftherianos et al. 2010)). For intracellular bacteria,
they exploit the engulfment of phagocytes to reach their destination within the cell.
17
Encapsulation is another type of cellular response directed against foreign objects too
large to be phagocytosed by individual phagocytes, such as parasitic nematodes, wasp
eggs, and host tissues undergoing apoptosis. The recognition of foreign intruders
initiates the attachment of plasmatocytes to their surface leading to the formation of
several layers; subsequently, the differentiated lamellocytes encompass the
multicellular sheath and the parasite is eventually destroyed by the local production of
ROSs (Nappi et al. 1995), reactive nitrogen species (RNSs) (Nappi et al. 2000) and
melanization(Nappi & Vass 1993). However, some invaders like parasitic nematodes
invading Manduca sexta escape from this reaction although parts of the body may be
encapsulated (Park et al. 2007). The differentiation and release of lamellocytes is one
of the hallmarks in this immune response, therefore an increased number of
circulating lamellocytes is observed in tumor inducing Drosophila strains.
Local and systemic induction of AMPs
Upon bacterial or fungal infections, antimicrobial peptides (AMPs) are transiently
synthesized and serve as self-produced antibiotics. The production of AMPs is mainly
regulated by two signaling pathways: the Imd (immune deficiency) and the Toll
pathways. This immune response can be activated locally, such as in the epidermis at
the site of infection, or systemically, meaning that the synthesis occurs in fat body.
The Toll pathway is activated in response to most of Gram-positive bacteria and fungi
(reviewed in (Lemaitre & Hoffmann 2007)). Unlike the Toll like receptors (TLRs) in
mammals, Toll receptor in Drosophila does not directly bind to PAMPs. Instead it
binds to the cleaved extracellular Spätzle, leading to the activation of two Relish
factors, the nuclear transcriptional factors (NF-ƘB) Dif and Dorsor (Weber et al.
2003). Normally, Dif and Dorsal form an inactive complex with Cactus, an inhibitor
of NF-ƘB. Upon the activation of Toll, the intracellular signal transduction leads to
the degradation of Cactus, in turn resulting in the translocation of Dif and Dorsal to
the nucleus where genes encoding AMPs like drosomycin and cecropin are expressed
(reviewed in (Engstrom 1999)).
The immune deficiency (Imd) pathway is turned on preferably by recognition of
gram-negative bacteria or in some cases of certain positive bacteria (Leulier et al.
2003) . This recognition is mediated through the receptors PGRP-LC and PGRP-LE
(Kaneko et al. 2006). As consequence, the activated receptors signal to Imd protein
and leads to the cleavage of another member of NF-ƘB family, Relish (Stoven et al.
2003). The cleaved Relish translocates to the nucleus and activates antibacterial
peptide gene expression (reviewed in (Ganesan et al. 2011)). The AMPs like
diptericin and attacin induced by this pathway are different from the ones induced by
the Toll pathway. The observation that some types of bacteria like S.aureus,
L.monocytogens activate both signaling pathways (Hashimoto et al. 2009; Igboin et al.
2012) suggests that there is some functional overlap and synergy in gene regulation
between the two them (reviewed in (Igboin et al. 2012)).
18
The Drosophila Toll and Imd pathways share some similarities both at the level of the
recognition receptors (Toll) and in their signal transduction (Rel/NF-ƘB) with
vertebrates despite differences at some level (reviewed in (Ganesan et al. 2011)).
Once again this suggests that immune mechanisms have been conserved through
evolution.
The AMPs secreted by the fat body are released into the hemolymph 30min to 1h
after the onset of infection and persist for several days during which time they target
and kill the bacteria and fungi. Most of the AMPs have a broad specificity for both
bacteria and even fungi like metchnikowin and cecropins, whereas some exhibit
narrow specificity such as drosomycin which is a potent antifungal molecule
(reviewed in(Meister et al. 2000)).
Using GFP reporter transgenes, Tzou et.al have shown that the antimicrobial peptides
can be induced in surface epithelia in a tissue-specific manner, meaning local
activation of AMPs (Tzou et al. 2000). More importantly the Imd pathway, rather than
the Toll pathway plays a critical role in the activation of this local response to
infection. The evidence that drosomycin expression, which is regulated by the Toll
pathway during the systemic response, is regulated by the Imd in the respiratory tract
speaks for the existence of distinct regulatory mechanisms for local and systemic
induction of antimicrobial peptide genes in Drosophila.
Melanization and Coagulation
Melanization
Melanization is a specific defense mechanism for insects and many other
invertebrates but not found in vertebrates. It is believed to contribute to wound
healing, sclerotization and encapsulation (reviewed in (Cerenius et al. 2008)). It is
also suggested that PO in Drosophila contributes to the strengthen of teh clot (Bidla et
al. 2005).The reaction involves the rapid synthesis and deposition of melanin (brownblack pigments) at the site of infection or injury (reviewed in (Nappi 1973;
Eleftherianos & Revenis 2011)). During the synthesis of melanin, the key enzyme
phenoloxidase (PO) is activated from its pro-form (proPO) via a serine protease
cascade(Kanost et al. 2004). The triggers responsible for proPO activation include the
recognition of PAMPs by PGRP-SA or GNBP (Park et al. 2007), tissue damage
inflicted by mechanical wounding (Galko & Krasnow 2004) and endogenous elicitors
from aberrant tissue like breakdown of basement membrane (Brennan & Anderson
2004; Bidla et al. 2009). Following the stimulation, proPO stored mainly in crystal
cells in case of Drosophila, is released into the extracellular milieu via the rupture of
crystal cells (see below)(Bidla et al. 2007) and is activated via a serine protease
cascade, proPO activation system (PPAS) (Kanost et al. 2004; Wang & Jiang 2004;
Tang et al. 2006). Given the toxicity of the products and by-products of the
19
melanization reaction, a tight control of PO-activation is necessary. In Drosophila,
one serine protease inhibitor (Spn27) has been identified to inhibit the proPO
activating proteases M1 and M2 (De Gregorio et al. 2002; Tang et al. 2006). Mutants
of Spn27A prevent uncontrolled melanization (systemic melanization) upon infection
or injury, and resulting lethality (De Gregorio et al. 2002).
Although the effects of melanization on parasitoid killing have been well documented
in both Drosophila and A. gambiae (Christensen et al. 2005), its effects on the
clearance of bacteria and fungi have recently been debated (reviewed in (Cerenius et
al. 2008)). Leclerc et.al have shown that Drosophila mutants that fail to activate PO in
the hemolymph were as resistant to microbial infections as wild-type flies (Leclerc et
al. 2006), whereas the importance of melanization in resistance to infections is
supported by the increased susceptibility to fungal and some bacterial infection of the
flies with defects in PO activation due to the silencing of serine protease MP1
(CG3066) (Tang et al. 2006; Ayres & Schneider 2008). The increased mortality of
the Bc Drosophila mutants, which lack of active PO after aseptic injury, implies PO in
wound healing.
PO released via the activation of crystal cells
The Drosophila genome encodes three POs, namely, DoxA1 (Bc/CG5779), DoxA2
(CG2952) and CG8193. Except DoxA2, which is expressed in lamellocytes, the other
two POs are expressed in crystal cells (reviewed in (Meister 2004)). Due to the lack of
signal peptides it has been unclear for decades how proPO was released into the
hemolymph to function extracellularly. Recently crystal cell rupture has been put
forward to explain this conundrum. Accompanying cell rupture the crystals, which are
believed to contain proPO dissolve. This process is regulated by JNK pathway
because the knock down of Bsk (Jun kinase) by RNAi blocks the rupture of crystal
cell, resulting in loss of melanization in larval bleeds (Bidla et al. 2007). Moreover, it
was found that the constitute activation of a RhoGTPase in blood cells also led to the
loss of clot melanization and mutant crystal cells appeared incapable of rupture and
dissolving crystals. This implies both the cytoskeleton and small GTPases in the
activation of crystal cell.
Coagulation
An ancient defense mechanism
Coagulation has historically been regarded as a process separate from the immune
responses due to the role it plays in hemostasis upon injury. Only recently has it been
proposed that coagulation and innate immunity evolved from a common ancestral
immune system (Opal & Esmon 2003). Most animal species, share the involvement of
both humoral and cellular mechanism during coagulation, a proteolytic cascade based
on zymogen activation, dependence of calcium ions (Ca2+) and a requirement for
transglutaminase (TG) dependent crosslinking of clot proteins (reviewed in (Cerenius
& Soderhall 2011)). However, except TG and to a limited degree hemolectin from the
insect Drosophila, factors participating in invertebrate coagulation lack a vertebrate
20
counterpart, suggesting that coagulation has evolved independently in different
species (reviewed in (Jiang & Doolittle 2003; Theopold et al. 2004; Cerenius &
Soderhall 2011)).
Since insects have an open circulatory system, coagulation must be activated quickly
to prevent excessive loss of body fluid. Upon wounding or triggered by microbes,
insect blood cells, mainly granulocytes (Galleria) or plasmatocytes (Drosophila) are
attracted (Galko & Krasnow 2004) to the wound sites and undergo degranulation of
procoagulants such as hemolectin and GP150 from Drosophila plasmatocytes (Lesch
et al. 2007; Lindgren et al. 2008). At the presence of phenoloxidase and TG, cellular
clot components are crosslinked to humoral/hemolymph clot proteins which are
primarily secreted by fat body, forming a protein meshwork (Fig. 3). In Drosophila,
using recombinant Fondue-GFP, the clot, could be labeled with indicating the
incorporation of Fondue into clots (Lindgren et al. 2008). Moreover, knockdown of
Fondue or TG causes larval bleeding defects (Scherfer et al. 2006; Lindgren et al.
2008), which functionally proves that both Fondue and TG are the true Drosophila lot
components. However, the role of PO in coagulation differs between different insect
species. The genetic mutation of PO in Drosophila does not affect the soft/primary
clot formation, whereas these mutants display wounding defects, suggesting that PO
may be not essential for primary clot formation but play a role at later stage in
hardening the wound plug (Bidla et al. 2005). In contrast to Drosophila, PO in the
mosquito Anopheles appears to be more critical in clot formation since blocking of
PO by addition of PTU (PO inhibitor) impedes the procoagulant like lipophorin
particles to coalesce into sheet structures (Agianian et al. 2007).
Figure 3. Schematic overview of coagulation in Drosophila and humans. Trauma elicits
clotting in both species. In humans, clotting activates two proteolytic cascades (the contact
and tissue factor systems), which culminate in the activation of thrombin and FXIII and the
formation of a stable clot. In Drosophila larvae, hemolymph clotting also involves a
proteolytic cascade (activation of PO) and secretion of clot components like TG and Fondue.
21
Mammalian coagulation is initiated by trauma. First, circulating platelets adhere to the
injured site by binding of integrin to collagen exposed on damaged tissue (Clemetson
et al. 1999). As a result, activated platelets degranulate and release factors that recruit
more platelets aggregating at the wound site and form a cellular plug. This is also
known as primary hemostasis (Clemetson 1999). However the complete arrest of
bleeding requires the proteolytic blood coagulation cascade leading to the formation
of a fibrin clot, a process that is called secondary hemostasis.
During secondary hemostasis, proteolytic cascades are activated by two pathways (fig.
3): contact activation or the intrinsic pathway initiated by damaged host cells or
foreign objects and the tissue factor or extrinsic activation pathway initiated by tissue
factor from wound site (Spronk et al. 2003). Both of them converge onto a common
pathway leading to the activation of thrombin, which in turn converts plasma
fibrinogen to fibrin. The loose fibrin mesh is finally thickened and strengthened by
crosslinking via FactorXIIIa (FXIIIa), a homologue to Drosophila transglutaminase
(TG) (Lorand & Graham 2003; Theopold et al. 2004).
As many infections in nature are concurrent with blood or hemolymph coagulation, it
is generally accepted that clotting serves as an immune effector, in some cases, as the
mediator of other immune reactions, such as during the reciprocal interaction between
complement and coagulation in mammals (reviewed in (Markiewski et al. 2007)).
The in vitro evidence from Drosophila strongly supports that coagulation facilitates a
localized immune response by immobilizing the pathogens and to some extent killing
them (Bidla et al. 2005). Therefore coagulation is integrated into host immunity.
Transglutaminase, the best conserved protein in coagulation
As a crosslinking enzyme, TG catalyzes the isopeptide bond formation between lysine
(Lys) and glutamine (Glu) residues from appropriated substrates in a Ca2+ dependent
manner (Shibata et al. 2010). Among the different clotting factors in both invertebrate
and vertebrate coagulation systems, transglutaminase is the only one which is highly
conserved (reviewed in (Theopold et al. 2004; Loof et al. 2011)). In mammals, FXIII
is converted from pro-transglutaminase by thrombin into activated FXIII in the
presence of Ca2+ and fibrin. As a consequence, fibrin fibrils are assembled tightly in a
spotwelding-like fashion. The FXIIIa required stabilization of clot structure is very
important, because FXIII deficiency and other disorders of stable fibrin meshes can be
life threatening (Lorand & Graham 2003; Hsieh & Nugent 2008). Similar to FXIII,
TG in horseshoe crabs is not required to form coagulin gel, but the gel is stabilized by
its crosslinking of proxin and stablin, two hemocyte derived proteins (Matsuda et al.
2007). Different from the former, in insects like Drosophila and in crustaceans, TG is
apparently necessary for producing the clot matrix, which is based on the fact that
knockdown of TG in Drosophila or blocking TG activity prevents clot formation
(Wang et al. 2001; Lindgren et al. 2008).
22
Eicosanoids and immunity
Eicosanoids are a group of biologically active metabolites of certain C20
polyunsaturated fatty acids. The biosynthesis of eicosanoids starts with the hydrolysis
of arachidonic acids (AAs) from membrane phospholipid pool via phospholipase A2
(PLA2). Depending on the cell-specific complement of enzymes, the free AA is a
substrate for cyclooxygenases (COXs), lipoxygenases (LOXs) and cytochrome
P450(Cyt P450) (reviewed in (Stanley et al. 2009)). The biosynthesis of eicosanoids
can be blocked by inhibition of PLA2 or of the three pathways.
The regulatory role of eicosanoids in immunity was first observed when inhibitor of
eicosanoid biosynthesis were shown to impair the host ability to form
microaggregates and nodules following the bacterial injection (Miller et al. 1994).
Later, the same group found that eicosanoids mediate the immune reaction to LPS
(Bedick et al. 2000) and to fungi (Lord et al. 2002). Besides chemical inhibitors,
several bacteria such as Xenorhabdus nematophilus suppress host immunity by
blocking eicosanoids biosynthesis (Park et al. 2003). Similarly, prevention of
eicosanoids production reduces encapsulation of parasitoids in Drosophila (Carton et
al. 2002; Stanley 2011). It is clear that eicosanoids interact with cellular immune
responses.
Beyond their contribution to cellular immunity, eicosanoids have also been reported
to be functionally coupled to the Imd pathway in Drosophila (Yajima et al. 2003). In
mammals, eicosanoids have been demonstrated to regulate blood coagulation
(reviewed in (Simmons et al. 2004)). It is tempting to speculate that eicosanoid play a
similar role during invertebrate coagulation.
23
24
The present study
The presented work in my thesis is based on the collaboration with my colleagues at
Stockholm University and external collaborators during my PhD journey.
Aim of this study
Despite our increased understanding of the molecular bases for Drosophila clot
formation, there is no solid evidence from in vivo studies demonstrating the immune
function of the Drosophila clot. This study aimed to fill this gap using the genetically
tractable model organism, Drosophila melanogaster. The advantages offered by
natural infection with entomopathogenic nematode prompted us to apply and improve
this infection model in Drosophila larvae (Paper I). Since TG and other proteins like
Fondue and GP150 have been identified and confirmed as true components of the
Drosophila clot, we have studied their functional role in host immunity in vivo and
provided evidence for an immune function of the Drosophila clot (Paper II and III).
Additional work has focused on the study of the molecular mechanism by which
crystal cells release proPO to contribute to clot formation (Paper IV).
25
Results and discussion
Improvement of a natural entomopathogenic nematode infection in Drosophila larvae
(Paper I)
When studying immunity, a common procedure to induce an immune response is to
inject microbes and immune elicitors into the host. This can be time consuming and
technically tricky, especially in larvae. In the wild, infective juveniles (IJs) of
entomopathogenic nematode (EPN) infect insect hosts naturally and release their
symbiotic bacteria to kill insects and complete their life cycle (reviewed in
(Eleftherianos et al. 2010)). Therefore, EPNs have been applied to study interactions
between the insect immune system and insect-pathogenic nematodes and their
symbiotic bacteria since the early 1980’s (P. Gotz et al. 1981). However, the lack of
genetic knowledge in most of insect model organisms, like Hyalophora cecropia and
Maduca sexta, hampers our understanding of the host-pathogen interactions, in
particular the host immune responses. The completion of the Drosophila genome
sequencing project, microarray analysis as well as the use of genetic screen, have
broadened our knowledge of host immune responses (reviewed in (Tzou et al. 2002;
Lemaitre & Ausubel 2008)). The aim of this paper was to improve the EPN infection
in Drosophila larvae based on the technique established by Hallem et al (Hallem et al.
2007) and to develop a fast and efficient strategy for nematode infection, which is
suitable for large-scale screens to identify genes that modulate host-pathogen
interactions.
To study the specific host preferences displayed by different nematode species, we
compared the infectivity of Steinernema and Heterorhabditis towards Drosophila
larvae. Both nematode species displayed the ability to infect and kill Drosophila
larvae with the latter being less pathogenic (Fig. 1). Given the different foraging
strategies employed by these two species, we next assessed the effect of media
(agarose and tissue paper) on nematode infection. Tissue paper appeared more
efficient for both nematodes to kill the host larvae (Fig. 2), but only with Steinernema,
these effects were significant. The more passive waiting for hosts by Steinernema
may require host exposed more often than possible on agarose. Then we tested a key
parameter for Heterorhabditis infection, i.e. the dosage of the nematode. Host
mortality displayed dose dependence (Fig. 3). To identify negative effects on the
immune response against EPNs, 100 IJs/larva is the optimal dose for a single host
challenged at room temperature, while a ten times higher concentration may be
suitable to identify dominant activators of the immune responses.
Since the improved nematode infection was performed with 96-well plates, which
requires individual inspection of the larvae in each well, we tried to improve scoring
the infection by mixing 50 host larvae with H.bacteriophora/GFP-expressing
P.luminescens in a transparent plastic bag. Due to the GFP labeled symbiotic bacteria,
both infected and killed host larvae are scored as GFP positive (Fig. 4 B). In addition
26
to that, GFP-expressing P.luminescens allowed us to monitor the infection in real time.
Using the novel assay, we could fully confirm the increase in susceptibility in
previously described lines (Fig. 4 A).
So, this newly established infection model in Drosophila larvae will be applicable in
large scale screens aimed at identifying novel genes/pathways involved in innate
immune responses.
Immune function of Drosophila clot in vivo (Paper II and paper III)
Since many infections in nature are accompanied by breaching body surfaces leading
to the hemolymph or blood coagulation, coagulation has been proposed to prevent not
only blood loss but also dissemination of infectious agents from entry to the host body
(reviewed in (Sun 2006)). In insects, this is supported by the observation that both
Galleria and Drosophila clot entrap bacteria in vitro (Bidla et al. 2005). However,
evidence from in vivo studies demonstrating an immune function for the Drosophila
clot is missing, despite the increased susceptibility to certain bacteria observed in the
larvae where the expression of clot component was reduced or lost (Scherfer et al.
2006; Lesch et al. 2007).
To study the immune function of the Drosophila clot in vivo, we focused on
previously identified clot components. Since transglutaminase (TG) has been shown
to play pivotal role in Drosophila clot formation based on the clotting defects caused
by reduced TG level in Drosophila larvae (Lindgren et al. 2008), we wondered
whether TG knockdown larvae displayed any defects in wound healing or combating
microbial infection. To our surprise, no wounding defects were observed in larvae
with reduced TG level (Fig. 5 A; paper II). In contrast, they were more sensitive to
pathogenic bacterial infection by S.aureus or P.luminescens but not to E.coli (Fig. 5 B;
paper II), which implicates TG in fighting pathogenic bacterial infection. More
interestingly, in an EPN natural infection model, TG knockdown larvae succumbed
more quickly to H.bacteriophora/P.lum complex while mutants for the Toll or the
Imd pathway survived as well as wild type (wt) (Fig. 5 C; paper II), despite the fact
that both the Toll and the Imd pathway-dependent AMPs are induced after infection
(Hallem et al. 2007). In addition, we tested the contribution of other host immune
responses including melanization and phagocytosis to the defense against this natural
infection. It appeared that none of these immune reactions were effective. Therefore,
it is very likely that this natural infection model revealed a previously uncharacterized
immune mechanism, namely the immune function of the Drosophila clot. To further
confirm this, we infected knockdown (RNAi) lines for previously identified clotting
factors, including Hemomucin, GP150, Fbp1, Tiggrin and Fondue (Schmidt &
Theopold 1997; Karlsson et al. 2004; Scherfer et al. 2004; Lindgren et al. 2008).
Similar to TG, reduced level of Fondue, or GP150 rendered the host more susceptible
to nematode infection (Fig. 1 A and B; paper III). This lends further support for a
27
function of the Drosophila clot in the response against nematodes and their bacteria.
Note that the influence of genes that did not show an effect on mortality cannot be
excluded, for example due to a partial knockdown.
Given the ability of the clot to entrap microbes, we speculated that the increase in
mortality after knocking down clotting factors may be caused by an impaired
sequestration of bacteria. Clots from TG-RNAi larvae did in fact entrap fewer bacteria
compared to those from wt larvae (Fig. 6 C; paper II), in line with quantification data
in Fig. 6 B; paper II. Moreover, a more brittle appearance of clots from TG-RNAi
larvae was observed, similar to what was found after inhibition of TG activity by
addition of MDC (Lindgren et al. 2008).
TG crosslinks selected glutamines and lysines in proteins leading to ε-(γ-glutamyl)
lysine bridges (reviewed in (Lorand & Graham 2003)). Therefore TG may target the
proteins from both pathogens and Drosophila itself. Using an antibody recognizing ε(γ-glutamyl) lysine bridges, we detected TG activity present as a punctuate pattern or
aggregates on the surface of zymosan beads that had been mixed with hemolymph
(Fig. 1 A; paper II). Such a punctuate pattern was also observed when we replaced the
antibody by biotin-cadaverine (B-cad), a small primary amine capable of replacing
lysine during TG-mediated crosslinking and which can serve to mark host proteins
involved in crosslinking. Other than on the surface of zymosan beads, TG activity was
also detected on the surface of both bacteria and nematodes in contact with
hemolymph. In all cases, the pattern after B-cad incorporation appeared to localize to
small deposits on the microbial surfaces (Fig. 1 B and C; paper II). Using GFP-tagged
Fondue, we traced Fondue not only to the clot but also to the surface of P.luminescens
as local aggregates (Fig. 2 C-F; paper III). The similarities in the patterns on pathogen
surfaces for both clot factors, TG and Fondue, suggest that the same mechanism is
active during sequestration of all pathogens.
To learn more about their composition, we performed affinity purification of bacterial
lysates after incubation with the biotinylated cadaverine. We identified hexamerin
subunits as the major constituent of the aggregates, while less abundant protein
components included phenoloxidase and lipophorin (Fig. 2 C; paper II). Taken
together, we propose a hypothetic mode for TG mediated immune function of
hemolymph clot: upon contact with hemolymph, microbes are almost instantaneously
targeted by TG activity and crosslinked to humoral clot factors, leading to the
formation of small aggregates and ultimately to sequestration by the clot matrix (Fig.
7; paper II).
Due to the evolutionary similarity between Drosophila TG and human Factor XIII,
we speculated that human Factor XIII might play the same role as Drosophila TG. To
test our hypothesis, we performed parallel experiments to the ones in Drosophila
using human blood. First, we asked whether human Factor XIII was able to target
microbes. This was shown to be the case using artificial substrate. Both E.coli and
S.aureus were targeted by human Factor XIII in contact with normal human plasma
28
although the reaction did not show the same pattern as on microbial surface as in
Drosophila (Fig. 3 A; paper II). Next we found that a fibrin meshwork formed from
Factor XIII deficient plasma, but it appeared less crosslinked compared to normal
plasma. And the ability to sequester bacteria differed substantially between Factor
XIII deficient and normal plasma, with a significant reduction of bacterial entrapment
in Factor XIII deficient clot (Fig. 3 B; paper II). Thus, human Factor XIII function
similarly as Drosophila TG, by targeting microbial surfaces and crosslinking them to
the clot fiber leading to their sequestration. This also indicates a functional
conservation between Drosophila TG and human Factor XIII.
Finally, the fact that eicosanoids are involved in regulation of vertebrate blood
coagulation (reviewed in (Simmons et al. 2004)) and insect immune responses
(reviewed in (Stanley 2011)) made them potential mediators in insect clotting. In the
nematode infection model, we detected increased mortality in Drosophila larvae
where eicosanoid biosynthesis had been blocked by injecting chemical inhibitors (Fig.
3 A; paper III), strongly suggesting the role of eicosanoids in the defense against
nematode infection. Furthermore the injection of a chemical inhibitor was also found
to interfere with the targeting of microbial surface by Fondue-GFP (Fig. 3A inset;
paper III), which provides evidence for eicosanoids as mediators in insect hemolymph
coagulation. Additionally, RNAi knockdown of the phospholipase A2
(CG11124/sPLA2), a potential upstream enzyme in the eicosanoids biosynthesis
pathway, also caused high mortality in the nematode infection model (Fig. 3 B; paper
III). Thus both genetic and chemical inhibition of eicosanoids biosynthesis impeded
the host immune responses against nematodes in Drosophila, which is in line with
the observation that some nematode symbiotic bacteria like P.lum evade host
immune reactions by suppressing PLA2 in other insect models (Kim et al. 2005).
Taken together we could show for the first time that the TG-dependent clotting
system in insects has an important role in the immune response to bacterial infection
or during a natural nematode infection. This immune function seems conserved during
evolution. Eicosanoids, appear to be involved in host immune responses potentially
by regulating the clotting system in Drosophila, although, further investigations into
their role in clotting are required.
Recognition Molecules and immune related genes involved in EPN infection in
Drosophila larvae
Entomopathogenic nematode infection as a natural infection model probably reflects
the natural order of microbe detection better than artificial injection of microbes,
which may bypass early stages of microbial encounter. Although the studies from
Hallem et.al and ours have shown that both phagocytosis and induction of AMPs
appeared not effective in the host defense against EPN, this does not exclude a
potential role of recognition molecules in combating EPN infection. To look for the
29
recognition molecules involved in this particular interaction between host and
pathogen, we performed EPN infection in mutants for several known pattern
recognition receptors (Fig. 4).
100
Mortality (%)
Wt
80
**
60
**
40
PGRP-SA
PGRP-LC
PGRP-LE
PGRP-LF
GNBP3
20
BcImd
0
Figure 4. Identification of PRRs that contribute to the defense against Heterorhabditis/Photorhabdus. Mutants of PRRs were infected by Heterorhabditis/Photorhabdus with a
dose of 25IJs per larvae at 29 °C and mortality was scored after 48. The double mutant
Bc/Imd served as a positive control. Data presented are means ±SD. ** P<0.01.
Mutants of pattern recognition receptors like PGRP-SA, -LC, -LE and GNBP3, which
sense directly the bacterial elicitors upstream of the Toll or the Imd pathway seemed
not sensitive to H.bacteriophora/P.luminescens complex. This result is in line with
the observation that mutants for the Toll and the Imd pathway survived EPN infection
as well as Wt animals (Hallem et al. 2007). Surprisingly, a lack of PGRP-LF caused
increased mortality of Drosophila larvae. PGRP-LF is a membrane associated PGRP
and has been demonstrated to act as a negative regulator of the Imd pathway by
interacting with and sequestering PGRP-LCx (Maillet et al. 2008; Basbous et al.
2011). In addition, PGRP-LF negatively regulates the activation of the JNK pathway.
Therefore, mutants of PGRP-LF show not only the activation of both the Imd and
JNK pathway, but also displayed developmental defects, which made the increased
mortality of these mutants in EPN infection more difficult to interpret. More
investigations have to been performed to confirm the role of PGRP-LF in the host
defense against the H.bacteriophora/P.luminescens complex. Thus far, there is no
solid prove that PRRs are required for the host to combat EPN natural infection.
TEPs are complement like proteins with structural similarity to vertebrate
complement C3 protein. These genes have been shown to be expressed in Drosophila
larvae and adult at a low level, but significantly upregulated after an immune
challenge (Lagueux et al. 2000; Bou Aoun et al. 2011). Recent microarray study on
the induction of immune related gene upon H.bacteriophora/P.luminescens infection
revealed the up-regulation of TEP2 (unpublished work from Michal Zurovec). This
prompted us to investigate the involvement of TEPs in activation of immunity
mechanisms with nematode infection model.
30
Mortality (%)
100
W1118
80
TEP2
60
**
**
40
TEP3
TEP4
TEP23
20
TEP234
0
Figure 5. TEP3 contributes to host defense against Heterorhabditis/Photorhabdus. Tep
mutants were subjected to the infection by Heterorhabditis/Photorhabdus with a dose of
25IJs per larvae at 29 °C and mortality was scored after 48. Data presented are means ±SD.
** P<0.01.
As shown in Fig. 5, the mutation of TEP2 appeared not effective for host in surviving
nematode infection despite its up-regulation shown in the microarray data. This is not
surprising since the upregulation of TEPs upon immune stimulation may not directly
account for their immune function. Contrarily, the TEP3 single mutants and the
double mutants for TEP2 and TEP3 (TEP23) displayed a clearly increased
susceptibility to nematode infections. To our knowledge this is the first time it could
be shown that a Drosophila TEP has an immune function in vivo. Although the exact
underlying mechanism remains to be elucidated, based on our previous study of
immune function of the clot, we postulate that Drosophila TEP3 may act as part of the
clotting system rather than as an opsonin. To test our hypothesis, we will next look
into the contribution of TEPs to hemolymph coagulation and to clot componentsmediated targeting of microbial surface.
Mortality (%)
100
Wt
Hemoless
80
60
**
40
20
0
Figure 6. Hemoless larvae are more susceptible to nematode infection. Apoptosis-induced
depletion of hemocytes in Drosophila larvae (Hemoless) led to significantly higher mortality
48h after infection with Heterorhabditis/Photorhabdus at a dose of 25IJs per larvae at 29 °C.
Data presented are means ±SD. ** P<0.01.
The Immune defects observed in hemocytes depleted Drosophila larvae (Hemoless)
reveal the role of hemocytes in multiple aspects of host immune defenses (Charroux
& Royet 2009). To evaluate the contribution of hemocytes to the host defense against
31
nematode, we infected the hemoless larvae where the majority of hemocytes are
depleted by induction of apoptosis with the H.bacteriophora/P.luminescens complex.
Remarkably, such larvae showed strong susceptibility to nematode infection (Fig. 6).
This result raised the question what was missing from the hemocytes to account for
the increased mortality. As a reservoir of immune genes, hemocytes express over
2,500 gene transcripts with significantly high hemocyte enrichment (Irving et al.
2005). To look for an answer to this question, we focused on a list of genes with
enriched expression in hemocytes including recognition molecules like Eater and
Lectins as well as extracellular matrix components such as Glutactin. We performed
targeted screen for genes involved using nematode infections. Strikingly, a reduction
of several genes’ expression compromised host immunity, leading to higher mortality,
although some of the results are at the border of significance and have to be
confirmed. The interesting genes comprise those encoding Glutactin, a protein
containing leucine rich repeats (CG4950) and an immunoglobulin-like protein
(CG7607) (Fig. 7).
100
Mortality (%)
80
Wt
PGRP-SA
PGRP-LC
Galectin
Lectin24Db
Lectin28C
SPARC
Eater
SR-CI
Glutactin
CG4950(LRR)
CG7607(Ig-like)
60
** ** **
40
20
0
Figure 7. Targeted screen for hemocyte enriched genes involved in defense against
nematode infection. RNAi lines for hemocyte-enriched genes were crossed with a hemocyte
specific driver (HeGAL4) and subjected to the infection by Heterorhabditis/Photorhabdus at
a dose of 25IJs per larvae at 29 °C, and mortality was scored after 48. Data presented are
means ±SD. ** P<0.01.
These in vivo data may therefore indicate an immune function for the three genes.
Glutactin, one of the basement membrane components, has been found associated
with the insect body wall, predominantly along the digestive tract, including the
proventriculus (a specialization of the anterior alimentary canal), midgut, and hindgut.
Since P.luminescens has been shown to occupy the specific niche between
32
extracellular matrix (ECM) and basal membrane of the folded midgut epithelium in
Manduca sexta (Silva et al. 2002), we speculate that reduced production of Glutactin
in hemocytes in Drosphila larvae might weaken the protection by the ECM or impede
entry more generally , resulting in the immune compromised phenotype. It is worth
looking into the role of other ECM components like Viking in host immune defense.
The function of the other two hits from screen, namely CG4950 and CG7607 are
unknown, however the protein domain they contain might indicate their involvement
in immunity.
Crystal cell activation via rupture (Paper IV)
Upon wounding or bleeding, Drosophila hemolymph clot is melanized (Bidla et al.
2005). And this melanization is tightly associated with crystal cell rupture, which is
regulated by the JNK pathway, RhoGTPases and Eiger (Bidla et al. 2007). The fact
that many factors leading to crystal cell activation remain unknown prompted us to
look for such factors. By using the melanization of hemolymph clot as readout we
analyzed mutants or RNAi lines for defects in this process.
Loss of clot melanization in the larvae overexpressing dominant active Rho1,
implicates small RhoGTPase in the process but due to the known side effects of
dominant active GTPases, are hard to interpret (Williams et al. 2007). To identify
possible small RhoGTPase, we analyzed available classical mutants for Rac1, Rac2
and Mtl. We found that clot melanization was lost in preparation from Rac2 mutants
(Fig. 2A, C and C’). This result is in agreement with the loss of melanization on
parasitoid eggs in Rac2Δ mutants (Williams et al. 2005). Thus, both findings suggest
a role for Rac2 in regulating localized melanization. Interestingly, the loss of control
in localized melanization was also reflected in vivo by a less concentrated
melanization at wound sites in Rac2Δ mutants, and it was pronounced in a Spn27A
mutant background where the inhibition of proPO activating system (PAS) is lost (Fig.
S2).
On the other hand, diffuse melanization was observed in the remainder of the
hemolymph in Rac2Δ mutants and could be measured photometrically (Fig. 2D). This
diffuse melanization was strongly inhibited in a JNK (Bsk) knockdown background
(Fig. 2D), adding more supports for the previous finding that crystal cell activation is
under control of JNK pathway, moreover, putting Rac2 genetically upstream of JNK
in the regulation of crystal cell rupture.
Previously, ectopic expression of the viral inhibitor of apoptosis protein p35 driven by
a hemocyte-specific Gal4 led to the inhibition of crystal cell rupture (Fig. 3B), but
partial melanization in the clot was observed (Fig. 3A). In this study overexpresssion
of p35 specifically in crystal cells caused the complete loss of clot melanization (Fig.
3A). This excludes plasmatocytes as the primary source for the observed defects and
33
indicates that apoptosis-like process may contribute to the activation of crystal cell
(Bertin et al. 1996; Kester & Nambu 2011). Next, we tested the role of all
proapoptotic proteins and caspases in melanization. Knockdown of two caspases,
namely Strica and Decay caused reduced melanization in the clot (Table. S1). The
reduced rather than complete loss of clot melanization may be explained by the
redundancy of caspases in crystal cell activation or incomplete silencing of caspase
gene. In the knockdown of proapoptotic genes by RNAi, we found that both GrimRNAi and Reaper-RNAi larvae displayed abolished clot melanization, with difference
that a diffuse melanization in hemolymph was observed in Grim-RNAi, reminiscent
of Rac2 phenotype. This lends support to the idea that deregulation of crystal cell may
activate the cell but this activation is not directed towards the clot and leads to diffuse
rather than localized melanization.
Since Rho1 is proven involved in both regulation of cytoskeleton (Hall 1998) and
activation crystal cell (Bidla et al. 2007), this triggered us to look for potential
interactors of Rho1 and cytoskeletal effectors, including Moesin, Rock, Zipper, which
might be involved in the same context. It turned out that reduced expression of
Moesin (Moe), an anchoring protein between cell membrane and cortex actin
(Edwards et al. 1997), caused not only loss of clot melanization and inactivated
crystal cells but also rendered the animal more susceptible to aseptic injury (Fig. 4 A,
B, G). Thus a functional role of Moe in the regulation of crystal cell activation gets
additional support. Using GFP-tagged Moe, we performed live imaging on crystal cell
rupture and revealed that there is an apparent increase in cell diameter, right before
the cell lysis (Fig. 4C-E). It has been shown that Moe functions antagonistically to the
Rho pathway to maintain epithelial integrity (Speck et al. 2003), however this
interaction between Rho1 and Moe in regulation of crystal cell rupture remains to be
shown.
A recent study by Mukherjee et al revealed the importance of endocytosis to maintain
crystal cells integrity in lymphgland (Mukherjee et al. 2011). This finding is
consistent with what our observations when we constitutively deactivated Rab5 in
crystal cell. Conversely the expression of Rab5 dominant active form specifically in
crystal cells led to the accumulation of vesicles (Fig. S4), in line with Rab5’s function
in vesicular traffic (Nielsen et al. 2008). As a result, the rupture of crystal cell was
disturbed, indicating the proper endocytosis is required for their immune activation.
Melanization requires tight control to keep it local and avoid unwanted or systemic
activation (reviwed in (Tang 2009)). This control is achieved by setting up
checkpoints at different levels, such as by using serpins which controls the serine
proteases cascade dependent PAS (De Gregorio et al. 2002; Ligoxygakis et al. 2002;
Gubb et al. 2010). Our findings on the regulators on crystal cell rupture identify
another check point to control melanization, namely the regulation of crystal cell
rupture. Although there are indications to imply p35 and caspases in melanization,
we have not been able to test if crystal cell rupture is has further similarities to
apoptosis. It is clear though that crystal cell activation is not the same as mammalian
34
neutrophil activation forming neutrophil extracellular traps (NETs), which involve
extracellular DNA fibers (Papayannopoulos & Zychlinsky 2009), since we observe
intact nuclei after crystal cell rupture (Fig. 1G-I). It is noteworthy that rupture may
not be the exclusive way to activate crystal cell and release proPO, in fact the proPO
containing crystals are observed dissolved in intact crystal cells in normal clot
preparation and bleeds. They may release proPO towards the targets in case of
septicemia via other means like exocytosis.
35
Concluding remarks
In this thesis, I focused mainly on studies of the immune function of the hemolymph
clot in Drosophila larvae and crystal cell activation.
Our understanding of the Drosophila clot has gone through different stages, from
morphological studies to molecular level, and now this work brought functional insight
furthering our knowledge of blood coagulation. For the first time, clot formation as an
immune defense mechanism in Drosophila was evident from our in vivo infection
studies on clot factors knockdown larvae. Reduced level of some clot factors rendered
larvae more sensitive to both bacterial and natural, namely EPN infections. And this
immune defect was shown to correlate with impaired bacterial entrapment mediated by
clot factors. Studies on two phylogenetically conserved clot components, Drosophila
TG and human FXIII, led to the conclusion that the clot acts as a conserved early innate
immune mechanism. Using the natural infection model, we also proved the involvement
of eicosanoids in Drosophila immunity. Moreover, the fact that blocking of
eicosanoinds biosynthesis inhibited clot factor (Fondue) to target microbial surface
implicated eicosanoids in the clot’s immune function.
Even though we have made progress in understanding of hemolymph coagulation, some
questions are left unanswered. Previous work has shown that entrapped bacteria were
eventually killed and this was independent of PO activity. An open question is what in
the clot causes the bacterial killing. It would be of interest to find whether clot factors
may gain antimicrobial activity after being processed perhaps by proteolytic cleavage.
As both Drosophila TG and human FXIII are able to target microbial surfaces, to
identify the specific membrane targets on pathogens would contribute to the knowledge
of the interaction between host and pathogen. In the mammalian system, blood
coagulation interacts constantly with lipids and eicosanoids are proven regulators of
blood coagulation. Whether this hold true in insect such as Drosophila remains to be
elucidated. Thus far, the clot factors identified in Drosophila larvae are most likely
incomplete. This is because the isolation and characterization of clot constituents has
been hampered by the rapid nature of clotting. One future task is to complete the list of
clot components. This will contribute to a better understanding of the clot as an integral
part of innate immunity in both vertebrates and invertebrates.
With the EPN infection model, we have identified TEP3, Glutactin and other two genes
(CG4950 and CG7607) that possibly act in the specific defense. These genes are not
novel but their role in Drosophila immunity is novel except for TEP3, which had been
implied in immune reactions. Due to the sequence similarity with vertebrate
complement proteins, insect TEPs have attracted attention for over a decade. However,
immune studies primarily found evidence for their up-regulation after immune
challenge and the ability to bind to specific bacteria, therefore promoting their
phagocytosis by S2 cells (Stroschein-Stevenson et al. 2006; Stroschein-Stevenson et al.
2009). Here we observed immune defects in larvae with loss of TEP3 expression,
36
strongly suggesting a functional role for TEP3 in the response against EPNs. It remains
unknown at this stage whether TEP3 is involved in the recognition of the pathogen, in
this case P.luminescens, or whether it takes part in the barrier defense. The possible
immune function for Glutactin in this study may further our understanding of the
extracellular matrix (ECM), particularly whether it contributes to host immunity. The
fact that we blocked Glutactin expression in hemocytes, indicates that hemocyte derived
ECM components participate host defense. One possible explanation is that hemocytes
sourced ECM components as readymade ‘building blocks’ are required for sealing and
repairing of wound, even sequestrations of invading microbes. Looking at other ECM
components will provide more information on their immune function. It is quite clear
that the natural infection model we applied in our studies is able to discover new
immune relevant genes. We believe that the combination of EPN infections and the
genetically tractable Drosophila system will continue to help us gain a more
comprehensive picture of insect immune defense.
Our studies on crystal cell activation provided more information on the regulation of
crystal cell rupture, one way to release proPO. We identified cytoskeletal components,
Moe, and cytoskeletal regulator Rac2 to as mediators of cell rupture. Following up
findings on the inhibition of crystal cell rupture by overexpression of p35, we tested the
role of all caspases and proapoptotic proteins in the activation of crystal cell. There was
signs of both caspases and proapoptotic proteins taking part in the process, but it is an
open question how many features crystal cell rupture shares with apoptosis. Lastly Rab5
was also characterized as a regulator of cell rupture. One piece of important
information we learned from this study was that crystal cell activation is not the same as
the activation of neutrophil. Although both involve terminal differentiation of cell, the
nucleus of crystal cell remains intact while the nucleus of neutrophil burst to release
their content during the activation.
Crystal cell rupture was observed to initiate always at one point on cell membrane. It
would be interesting to look at the rearrangement of cytoskeleton or depletion of lipids
at the initial point upon sensing of the directional signal. It has been shown that Moe
regulates Rho1 activity upstream during JNK dependent apoptosis. We would like to
know if this Moe and Rho regulated JNK pathway is shared during activation of crystal
cell. The answer to this would add to a more comprehensive picture of crystal cell
activation.
In summary, this thesis work furthers our understanding of clot’s immune function in
both Drosophila and humans. Additionally, the new insight into host-pathogen
interaction provided by this study may broaden our understanding of the medical
implications of clot, and lead to the new strategies for enhancing the clearance of
pathogens.
37
38
Acknowledgements
Many people have contributed to my work with support of many kinds.
First, my sincere gratitude goes to my supervisor, Ulrich Theopold. Thank you so
much for your optimism and open mindedness, for being always supportive in trying
new ideas and techniques, for your constant encouragement of scientific
independence. I have been greatly enjoyed being a member of your research group,
which, among other things, has provided me the opportunity to meet many fascinating
people. I am also very grateful to you and your wife Gaby for the help with my life
here and for the summer BBQ and winter fondue at your place.
I also want to thank Christos Samakovlis for offering me the opportunity to come to
Stockholm to pursue my study and for introducing me into fly world.
Thanks Michael, for offering me the one-week training on wasp parasitization,
dissection and immunostaining in Aberdeen University.
I express my deep gratitude to past and present lab colleagues: Thomas, for all the
laughs we have shared during these years. Being your partner at work has really
entertained us and even people around us. I cannot imagine my PhD journey without
you. Robert M, for teaching me all the skills in photography and microscopy, and
sharing all your PhD experience and wisdom with us. Robert K, for being such an
active and young gentleman to remind me how ‘old lady’ I am from time to time.
Thank you for being so open to share all your interest and knowledge in science and
concerts. I wish you could have joined our group earlier, so we can have more our
two-person journal club. Hope you can keep it with other people. Badrul, thank you
for the help at my last stay of PhD study, wish you good luck with your PhD study.
Maja, your coming gave me the opportunity to talk with girl in the lab and office
again. Thank you for showing me how efficiently a woman can work when she
becomes mom and for all the discussion in science and life in general. Gawa, for the
guidance you gave to me at the beginning and showing me that science can be
pursued without protocol to some extent. Christine, it was wonderful experience to
share project with you. And thank you for inviting me for meal at your place and
meeting your nice family. Tine, for presenting me the importance of being organized
and for being so helpful whenever I needed your help. Malin, for showing me the
complicated Gateway system when I first joined the group and your expertise in
dancing. Also there were many project students who had contributed to the
accomplishment of the projects. They are Martina Klupp, Carina Krützmann, Kathrin
Handge, Raha Riazi, Ashti Amin Said, Yvonne de Graaf and Marietherese Wuelbern.
Thanks for all their work.
Thanks to all the collaborators involved in this thesis. Pavel Hyrsl and Pavel Dobes
from Masaryk University, Czech, thank both of you for introducing me to the world
of nematodes and for the great time we shared inside and outside lab. Special thanks
39
gave to Pavel H and your wife Allena for guesting me in Brno, that’s really kind of
you. Pavel D, for teaching me to pick up the delicious mushrooms. Heiko
Herwald,Torsten G. Loof and Matthias Mörgelin from Lund University, for the work
on FXIII. Jennifer Iklé and Richard M. Cripps for sending us antibody against
Transglutaminase. Olga Loseva, for the technique support for proteomics.
My special appreciations go to the ‘fly people’: Shiva, for sharing your smiles and
nice chats. Monica, for giving me so many good advices regarding science and life in
Sweden. Widad, for the nice trip to Washington D.C. Anne, thank you for taking off
some of the teaching tasks from me. Gunnel, for showing me how to keep our
important working space, insect room, in order and sharing the musics from FM107.1.
Britta, thank you so much for taking so much work from us and for making our study
go smoothly.
Thanks to all the staff from department of molecular biology and functional genomics
for the times we shared in our kitchen and for creating the conductive and very
friendly working environment. Johan and Simei, it was nice experience to share the
project with you.
A ‘big thanks’ to all my friends outside the lab. Your smiles, greetings and caring sent
by phone calls or emails, information and advices made my stay in Sweden one of the
best time period in my life.
My deepest gratitude to my family in China. 感谢妈妈, 爸爸,姐姐,姐夫还有甜
甜对我的坚持不懈的信任和支持。 尤其感谢姐姐姐夫这么多年来在家对妈妈的
照顾。谢谢我所有的亲戚对我一直以来的关心和爱护。 感谢我的公公婆婆对我
们在外生活的体谅和对我由衷的关爱。
Finally, I am so grateful to my husband, Liang, for always being such a considerate
man and having faith in me. Thank you for your endless love of me and of our little
family. Thanks for your coming, my little daughter to be, you will be the biggest
treasure in the world for me. Love both of you so much!
40
References
Agaisse H, Petersen UM, Boutros M et al. (2003) Signaling role of hemocytes in Drosophila
JAK/STAT-dependent response to septic injury. Developmental cell 5, 441-450.
Agianian B, Lesch C, Loseva O et al. (2007) Preliminary characterization of hemolymph
coagulation in Anopheles gambiae larvae. Developmental and comparative
immunology 31, 879-888.
Ao J, Ling E, Yu XQ (2007) Drosophila C-type lectins enhance cellular encapsulation.
Molecular immunology 44, 2541-2548.
Ayres JS, Schneider DS (2008) A signaling protease required for melanization in Drosophila
affects resistance and tolerance of infections. PLoS biology 6, 2764-2773.
Basbous N, Coste F, Leone P et al. (2011) The Drosophila peptidoglycan-recognition protein
LF interacts with peptidoglycan-recognition protein LC to downregulate the Imd
pathway. EMBO reports 12, 327-333.
Basset A, Khush RS, Braun A et al. (2000) The phytopathogenic bacteria Erwinia carotovora
infects Drosophila and activates an immune response. Proceedings of the National
Academy of Sciences of the United States of America 97, 3376-3381.
Bedick JC, Pardy RL, Howard RW et al. (2000) Insect cellular reactions to the
lipopolysaccharide component of the bacterium Serratia marcescens are mediated
by eicosanoids. Journal of insect physiology 46, 1481-1487.
Bednarek A, Gaugler R (1997) Compatibility of soil amendments with entomopathogenic
nematodes. Journal of nematology 29, 220-227.
Bertin J, Mendrysa SM, LaCount DJ et al. (1996) Apoptotic suppression by baculovirus P35
involves cleavage by and inhibition of a virus-induced CED-3/ICE-like protease.
Journal of virology 70, 6251-6259.
Bianchi ME (2007) DAMPs, PAMPs and alarmins: all we need to know about danger. Journal
of leukocyte biology 81, 1-5.
Bidla G, Dushay MS, Theopold U (2007) Crystal cell rupture after injury in Drosophila
requires the JNK pathway, small GTPases and the TNF homolog Eiger. Journal of cell
science 120, 1209-1215.
Bidla G, Hauling T, Dushay MS et al. (2009) Activation of insect phenoloxidase after injury:
endogenous versus foreign elicitors. Journal of innate immunity 1, 301-308.
Bidla G, Lindgren M, Theopold U et al. (2005) Hemolymph coagulation and phenoloxidase in
Drosophila larvae. Developmental and comparative immunology 29, 669-679.
Boman HG, Nilsson I, Rasmuson B (1972) Inducible antibacterial defence system in
Drosophila. Nature 237, 232-235.
Bou Aoun R, Hetru C, Troxler L et al. (2011) Analysis of thioester-containing proteins during
the innate immune response of Drosophila melanogaster. Journal of innate
immunity 3, 52-64.
Brennan CA, Anderson KV (2004) Drosophila: the genetics of innate immune recognition and
response. Annual review of immunology 22, 457-483.
Brennan CA, Delaney JR, Schneider DS et al. (2007) Psidin is required in Drosophila blood
cells for both phagocytic degradation and immune activation of the fat body.
Current biology : CB 17, 67-72.
Carton Y, Frey F, Stanley DW et al. (2002) Dexamethasone inhibition of the cellular immune
response of Drosophila melanogaster against a parasitoid. The Journal of
parasitology 88, 405-407.
Carton Y, Nappi AJ (2001) Immunogenetic aspects of the cellular immune response of
Drosophilia against parasitoids. Immunogenetics 52, 157-164.
41
Cerenius L, Lee BL, Soderhall K (2008) The proPO-system: pros and cons for its role in
invertebrate immunity. Trends in immunology 29, 263-271.
Cerenius L, Soderhall K (2011) Coagulation in invertebrates. Journal of innate immunity 3, 38.
Chaput C, Boneca IG (2007) Peptidoglycan detection by mammals and flies. Microbes and
infection / Institut Pasteur 9, 637-647.
Charroux B, Rival T, Narbonne-Reveau K et al. (2009) Bacterial detection by Drosophila
peptidoglycan recognition proteins. Microbes and infection / Institut Pasteur 11,
631-636.
Charroux B, Royet J (2009) Elimination of plasmatocytes by targeted apoptosis reveals their
role in multiple aspects of the Drosophila immune response. Proceedings of the
National Academy of Sciences of the United States of America 106, 9797-9802.
Christensen BM, Li J, Chen CC et al. (2005) Melanization immune responses in mosquito
vectors. Trends in parasitology 21, 192-199.
Ciche T (2007) The biology and genome of Heterorhabditis bacteriophora. WormBook : the
online review of C elegans biology, 1-9.
Clemetson JM, Polgar J, Magnenat E et al. (1999) The platelet collagen receptor glycoprotein
VI is a member of the immunoglobulin superfamily closely related to FcalphaR and
the natural killer receptors. The Journal of biological chemistry 274, 29019-29024.
Clemetson KJ (1999) Primary haemostasis: sticky fingers cement the relationship. Current
biology : CB 9, R110-112.
Crozatier M, Meister M (2007) Drosophila haematopoiesis. Cellular microbiology 9, 11171126.
Davis MM, Alvarez FJ, Ryman K et al. (2011) Wild-type Drosophila melanogaster as a model
host to analyze nitrogen source dependent virulence of Candida albicans. PloS one 6,
e27434.
De Gregorio E, Han SJ, Lee WJ et al. (2002) An immune-responsive Serpin regulates the
melanization cascade in Drosophila. Developmental cell 3, 581-592.
Dionne MS, Ghori N, Schneider DS (2003) Drosophila melanogaster is a genetically tractable
model host for Mycobacterium marinum. Infection and immunity 71, 3540-3550.
Dushay MS, Eldon ED (1998) Drosophila immune responses as models for human immunity.
American journal of human genetics 62, 10-14.
Edwards KA, Demsky M, Montague RA et al. (1997) GFP-moesin illuminates actin
cytoskeleton dynamics in living tissue and demonstrates cell shape changes during
morphogenesis in Drosophila. Developmental biology 191, 103-117.
Ehlers RU (2001) Mass production of entomopathogenic nematodes for plant protection.
Applied microbiology and biotechnology 56, 623-633.
Ekengren S, Tryselius Y, Dushay MS et al. (2001) A humoral stress response in Drosophila.
Current biology : CB 11, 714-718.
Eleftherianos I, ffrench-Constant RH, Clarke DJ et al. (2010) Dissecting the immune response
to the entomopathogen Photorhabdus. Trends in microbiology 18, 552-560.
Eleftherianos I, Revenis C (2011) Role and importance of phenoloxidase in insect hemostasis.
Journal of innate immunity 3, 28-33.
Engstrom Y (1999) Induction and regulation of antimicrobial peptides in Drosophila.
Developmental and comparative immunology 23, 345-358.
Ferrandon D, Imler JL, Hetru C et al. (2007) The Drosophila systemic immune response:
sensing and signalling during bacterial and fungal infections. Nature reviews
Immunology 7, 862-874.
Franc NC, Heitzler P, Ezekowitz RA et al. (1999) Requirement for croquemort in phagocytosis
of apoptotic cells in Drosophila. Science 284, 1991-1994.
42
Galko MJ, Krasnow MA (2004) Cellular and genetic analysis of wound healing in Drosophila
larvae. PLoS biology 2, E239.
Ganesan S, Aggarwal K, Paquette N et al. (2011) NF-kappaB/Rel proteins and the humoral
immune responses of Drosophila melanogaster. Current topics in microbiology and
immunology 349, 25-60.
Glittenberg MT, Silas S, MacCallum DM et al. (2011) Wild-type Drosophila melanogaster as
an alternative model system for investigating the pathogenicity of Candida albicans.
Disease models & mechanisms 4, 504-514.
Gobert V, Gottar M, Matskevich AA et al. (2003) Dual activation of the Drosophila toll
pathway by two pattern recognition receptors. Science 302, 2126-2130.
Gottar M, Gobert V, Matskevich AA et al. (2006) Dual detection of fungal infections in
Drosophila via recognition of glucans and sensing of virulence factors. Cell 127,
1425-1437.
Gubb D, Sanz-Parra A, Barcena L et al. (2010) Protease inhibitors and proteolytic signalling
cascades in insects. Biochimie 92, 1749-1759.
Hall A (1998) Rho GTPases and the actin cytoskeleton. Science 279, 509-514.
Hallem EA, Dillman AR, Hong AV et al. (2011) A sensory code for host seeking in parasitic
nematodes. Current biology : CB 21, 377-383.
Hallem EA, Rengarajan M, Ciche TA et al. (2007) Nematodes, bacteria, and flies: a tripartite
model for nematode parasitism. Current biology : CB 17, 898-904.
Han R, Ehlers RU (2000) Pathogenicity, development, and reproduction of Heterorhabditis
bacteriophora and Steinernema carpocapsae under axenic in vivo conditions.
Journal of invertebrate pathology 75, 55-58.
Hashimoto Y, Tabuchi Y, Sakurai K et al. (2009) Identification of lipoteichoic acid as a ligand
for draper in the phagocytosis of Staphylococcus aureus by Drosophila hemocytes. J
Immunol 183, 7451-7460.
Hoebe K, Janssen E, Beutler B (2004) The interface between innate and adaptive immunity.
Nature immunology 5, 971-974.
Holz A, Bossinger B, Strasser T et al. (2003) The two origins of hemocytes in Drosophila.
Development 130, 4955-4962.
Hsieh L, Nugent D (2008) Factor XIII deficiency. Haemophilia : the official journal of the
World Federation of Hemophilia 14, 1190-1200.
Igboin CO, Griffen AL, Leys EJ (2012) The Drosophila melanogaster host model. Journal of
oral microbiology 4.
Irving P, Ubeda JM, Doucet D et al. (2005) New insights into Drosophila larval haemocyte
functions through genome-wide analysis. Cellular microbiology 7, 335-350.
Janeway CA, Jr. (1989) Approaching the asymptote? Evolution and revolution in immunology.
Cold Spring Harbor symposia on quantitative biology 54 Pt 1, 1-13.
Janeway CA, Jr. (1992) The immune system evolved to discriminate infectious nonself from
noninfectious self. Immunology today 13, 11-16.
Janeway CA, Jr., Medzhitov R (2002) Innate immune recognition. Annual review of
immunology 20, 197-216.
Jiang Y, Doolittle RF (2003) The evolution of vertebrate blood coagulation as viewed from a
comparison of puffer fish and sea squirt genomes. Proceedings of the National
Academy of Sciences of the United States of America 100, 7527-7532.
Kaneko T, Yano T, Aggarwal K et al. (2006) PGRP-LC and PGRP-LE have essential yet distinct
functions in the drosophila immune response to monomeric DAP-type peptidoglycan.
Nature immunology 7, 715-723.
Kanost MR, Jiang H, Yu XQ (2004) Innate immune responses of a lepidopteran insect,
Manduca sexta. Immunological reviews 198, 97-105.
43
Karlsson C, Korayem AM, Scherfer C et al. (2004) Proteomic analysis of the Drosophila larval
hemolymph clot. The Journal of biological chemistry 279, 52033-52041.
Kester RS, Nambu JR (2011) Targeted expression of p35 reveals a role for caspases in
formation of the adult abdominal cuticle in Drosophila. The International journal of
developmental biology 55, 109-119.
Kim Y, Ji D, Cho S et al. (2005) Two groups of entomopathogenic bacteria, Photorhabdus and
Xenorhabdus, share an inhibitory action against phospholipase A2 to induce host
immunodepression. Journal of invertebrate pathology 89, 258-264.
Kocks C, Cho JH, Nehme N et al. (2005) Eater, a transmembrane protein mediating
phagocytosis of bacterial pathogens in Drosophila. Cell 123, 335-346.
Kurucz E, Markus R, Zsamboki J et al. (2007) Nimrod, a putative phagocytosis receptor with
EGF repeats in Drosophila plasmatocytes. Current biology : CB 17, 649-654.
Lagueux M, Perrodou E, Levashina EA et al. (2000) Constitutive expression of a complementlike protein in toll and JAK gain-of-function mutants of Drosophila. Proceedings of
the National Academy of Sciences of the United States of America 97, 11427-11432.
Lazzaro BP, Rolff J (2011) Immunology. Danger, microbes, and homeostasis. Science 332, 4344.
Leclerc V, Pelte N, El Chamy L et al. (2006) Prophenoloxidase activation is not required for
survival to microbial infections in Drosophila. EMBO reports 7, 231-235.
Lemaitre B, Ausubel FM (2008) Animal models for host-pathogen interactions. Current
opinion in microbiology 11, 249-250.
Lemaitre B, Hoffmann J (2007) The host defense of Drosophila melanogaster. Annual review
of immunology 25, 697-743.
Lemaitre B, Nicolas E, Michaut L et al. (1996) The dorsoventral regulatory gene cassette
spatzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell
86, 973-983.
Lesch C, Goto A, Lindgren M et al. (2007) A role for Hemolectin in coagulation and immunity
in Drosophila melanogaster. Developmental and comparative immunology 31, 12551263.
Leulier F, Parquet C, Pili-Floury S et al. (2003) The Drosophila immune system detects
bacteria through specific peptidoglycan recognition. Nature immunology 4, 478-484.
Levashina EA, Moita LF, Blandin S et al. (2001) Conserved role of a complement-like protein
in phagocytosis revealed by dsRNA knockout in cultured cells of the mosquito,
Anopheles gambiae. Cell 104, 709-718.
Li XC, RS.; Cowles, EA.; Gaugler, R.; Cox-Foster, DL.. (2007) Relationship between the
successful infection by entomopathogenic nematodes and the host immune
response. Int J Parasitol 37, 10.
Ligoxygakis P, Pelte N, Ji C et al. (2002) A serpin mutant links Toll activation to melanization
in the host defence of Drosophila. The EMBO journal 21, 6330-6337.
Lindgren M, Riazi R, Lesch C et al. (2008) Fondue and transglutaminase in the Drosophila
larval clot. Journal of insect physiology 54, 586-592.
Loof TG, Schmidt O, Herwald H et al. (2011) Coagulation systems of invertebrates and
vertebrates and their roles in innate immunity: the same side of two coins? Journal
of innate immunity 3, 34-40.
Lorand L, Graham RM (2003) Transglutaminases: crosslinking enzymes with pleiotropic
functions. Nature reviews Molecular cell biology 4, 140-156.
Lord JC, Anderson S, Stanley DW (2002) Eicosanoids mediate Manduca sexta cellular
response to the fungal pathogen Beauveria bassiana: a role for the lipoxygenase
pathway. Archives of insect biochemistry and physiology 51, 46-54.
44
Maillet F, Bischoff V, Vignal C et al. (2008) The Drosophila peptidoglycan recognition protein
PGRP-LF blocks PGRP-LC and IMD/JNK pathway activation. Cell host & microbe 3,
293-303.
Markiewski MM, Nilsson B, Ekdahl KN et al. (2007) Complement and coagulation: strangers
or partners in crime? Trends in immunology 28, 184-192.
Markus R, Laurinyecz B, Kurucz E et al. (2009) Sessile hemocytes as a hematopoietic
compartment in Drosophila melanogaster. Proceedings of the National Academy of
Sciences of the United States of America 106, 4805-4809.
Matsuda Y, Osaki T, Hashii T et al. (2007) A cysteine-rich protein from an arthropod stabilizes
clotting mesh and immobilizes bacteria at injury sites. The Journal of biological
chemistry 282, 33545-33552.
Medzhitov R, Janeway CA, Jr. (1998) An ancient system of host defense. Current opinion in
immunology 10, 12-15.
Meister M (2004) Blood cells of Drosophila: cell lineages and role in host defence. Current
opinion in immunology 16, 10-15.
Meister M, Hetru C, Hoffmann JA (2000) The antimicrobial host defense of Drosophila.
Current topics in microbiology and immunology 248, 17-36.
Mellroth P, Steiner H (2006) PGRP-SB1: an N-acetylmuramoyl L-alanine amidase with
antibacterial activity. Biochemical and biophysical research communications 350,
994-999.
Michel T, Reichhart JM, Hoffmann JA et al. (2001) Drosophila Toll is activated by Grampositive bacteria through a circulating peptidoglycan recognition protein. Nature 414,
756-759.
Miller JS, Nguyen T, Stanley-Samuelson DW (1994) Eicosanoids mediate insect nodulation
responses to bacterial infections. Proceedings of the National Academy of Sciences
of the United States of America 91, 12418-12422.
Miyake Y, Yamasaki S (2012) Sensing necrotic cells. Advances in experimental medicine and
biology 738, 144-152.
Mukherjee T, Kim WS, Mandal L et al. (2011) Interaction between Notch and Hif-alpha in
development and survival of Drosophila blood cells. Science 332, 1210-1213.
Nappi AJ (1973) Hemocytic changes associated with the encapsulation and melanization of
some insect parasites. Experimental parasitology 33, 285-302.
Nappi AJ, Vass E (1993) Melanogenesis and the generation of cytotoxic molecules during
insect cellular immune reactions. Pigment cell research / sponsored by the European
Society for Pigment Cell Research and the International Pigment Cell Society 6, 117126.
Nappi AJ, Vass E, Frey F et al. (1995) Superoxide anion generation in Drosophila during
melanotic encapsulation of parasites. European journal of cell biology 68, 450-456.
Nappi AJ, Vass E, Frey F et al. (2000) Nitric oxide involvement in Drosophila immunity. Nitric
oxide : biology and chemistry / official journal of the Nitric Oxide Society 4, 423-430.
Nielsen E, Cheung AY, Ueda T (2008) The regulatory RAB and ARF GTPases for vesicular
trafficking. Plant physiology 147, 1516-1526.
Opal SM, Esmon CT (2003) Bench-to-bedside review: functional relationships between
coagulation and the innate immune response and their respective roles in the
pathogenesis of sepsis. Crit Care 7, 23-38.
P. Gotz, and AB, Boman HG (1981) Interactions between Insect Immunity and an InsectPathogenic Nematode with Symbiotic Bacteria. Proc R Soc Lond B 212, 18.
Pal S, Wu LP (2009) Pattern recognition receptors in the fly: lessons we can learn from the
Drosophila melanogaster immune system. Fly 3, 121-129.
Papayannopoulos V, Zychlinsky A (2009) NETs: a new strategy for using old weapons. Trends
in immunology 30, 513-521.
45
Paredes JC, Welchman DP, Poidevin M et al. (2011) Negative regulation by amidase PGRPs
shapes the Drosophila antibacterial response and protects the fly from innocuous
infection. Immunity 35, 770-779.
Park JW, Kim CH, Kim JH et al. (2007) Clustering of peptidoglycan recognition protein-SA is
required for sensing lysine-type peptidoglycan in insects. Proceedings of the
National Academy of Sciences of the United States of America 104, 6602-6607.
Park Y, Kim Y, Putnam SM et al. (2003) The bacterium Xenorhabdus nematophilus depresses
nodulation reactions to infection by inhibiting eicosanoid biosynthesis in tobacco
hornworms, Manduca sexta. Archives of insect biochemistry and physiology 52, 7180.
Pearson A, Lux A, Krieger M (1995) Expression cloning of dSR-CI, a class C macrophagespecific scavenger receptor from Drosophila melanogaster. Proceedings of the
National Academy of Sciences of the United States of America 92, 4056-4060.
Philips JA, Rubin EJ, Perrimon N (2005) Drosophila RNAi screen reveals CD36 family member
required for mycobacterial infection. Science 309, 1251-1253.
Poinar GOJ, Thomas, G.M., and Hess, R. (1977) Characteristics of the specific bacterium
associated with Heterorhabditis baceriophora heterorhabditidae rhabditida.
Nematologica 21, 8.
Poltorak A, He X, Smirnova I et al. (1998) Defective LPS signaling in C3H/HeJ and
C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282, 2085-2088.
Ramet M, Manfruelli P, Pearson A et al. (2002) Functional genomic analysis of phagocytosis
and identification of a Drosophila receptor for E. coli. Nature 416, 644-648.
Royet J (2004) Infectious non-self recognition in invertebrates: lessons from Drosophila and
other insect models. Molecular immunology 41, 1063-1075.
Royet J, Gupta D, Dziarski R (2011) Peptidoglycan recognition proteins: modulators of the
microbiome and inflammation. Nature reviews Immunology 11, 837-851.
Ryu JH, Kim SH, Lee HY et al. (2008) Innate immune homeostasis by the homeobox gene
caudal and commensal-gut mutualism in Drosophila. Science 319, 777-782.
Sang Y, Ramanathan B, Ross CR et al. (2005) Gene silencing and overexpression of porcine
peptidoglycan recognition protein long isoforms: involvement in beta-defensin-1
expression. Infection and immunity 73, 7133-7141.
Scherfer C, Karlsson C, Loseva O et al. (2004) Isolation and characterization of hemolymph
clotting factors in Drosophila melanogaster by a pullout method. Current biology :
CB 14, 625-629.
Scherfer C, Qazi MR, Takahashi K et al. (2006) The Toll immune-regulated Drosophila protein
Fondue is involved in hemolymph clotting and puparium formation. Developmental
biology 295, 156-163.
Schleifer KH, Kandler O (1972) Peptidoglycan types of bacterial cell walls and their taxonomic
implications. Bacteriological reviews 36, 407-477.
Schmidt O, Theopold U (1997) Helix pomatia lectin and annexin V, two molecular probes for
insect microparticles: possible involvement in hemolymph coagulation. Journal of
insect physiology 43, 667-674.
Schmucker D, Clemens JC, Shu H et al. (2000) Drosophila Dscam is an axon guidance receptor
exhibiting extraordinary molecular diversity. Cell 101, 671-684.
Shibata T, Ariki S, Shinzawa N et al. (2010) Protein crosslinking by transglutaminase controls
cuticle morphogenesis in Drosophila. PloS one 5, e13477.
Silva CP, Waterfield NR, Daborn PJ et al. (2002) Bacterial infection of a model insect:
Photorhabdus luminescens and Manduca sexta. Cellular microbiology 4, 329-339.
Simmons DL, Botting RM, Hla T (2004) Cyclooxygenase isozymes: the biology of
prostaglandin synthesis and inhibition. Pharmacological reviews 56, 387-437.
46
Sorrentino RP, Carton Y, Govind S (2002) Cellular immune response to parasite infection in
the Drosophila lymph gland is developmentally regulated. Developmental biology
243, 65-80.
Speck O, Hughes SC, Noren NK et al. (2003) Moesin functions antagonistically to the Rho
pathway to maintain epithelial integrity. Nature 421, 83-87.
Spronk HM, Govers-Riemslag JW, ten Cate H (2003) The blood coagulation system as a
molecular machine. BioEssays : news and reviews in molecular, cellular and
developmental biology 25, 1220-1228.
Stanley D (2011) Eicosanoids: progress towards manipulating insect immunity. J Appl
Entomol 135, 12.
Stanley D, Miller J, Tunaz H (2009) Eicosanoid actions in insect immunity. Journal of innate
immunity 1, 282-290.
Steiner H, Hultmark D, Engstrom A et al. (1981) Sequence and specificity of two antibacterial
proteins involved in insect immunity. Nature 292, 246-248.
Stoven S, Silverman N, Junell A et al. (2003) Caspase-mediated processing of the Drosophila
NF-kappaB factor Relish. Proceedings of the National Academy of Sciences of the
United States of America 100, 5991-5996.
Stroschein-Stevenson SL, Foley E, O'Farrell PH et al. (2006) Identification of Drosophila gene
products required for phagocytosis of Candida albicans. PLoS biology 4, e4.
Stroschein-Stevenson SL, Foley E, O'Farrell PH et al. (2009) Phagocytosis of Candida albicans
by RNAi-treated Drosophila S2 cells. Methods Mol Biol 470, 347-358.
Stuart LM, Deng J, Silver JM et al. (2005) Response to Staphylococcus aureus requires CD36mediated phagocytosis triggered by the COOH-terminal cytoplasmic domain. The
Journal of cell biology 170, 477-485.
Stuart LM, Ezekowitz RA (2005) Phagocytosis: elegant complexity. Immunity 22, 539-550.
Sun H (2006) The interaction between pathogens and the host coagulation system.
Physiology (Bethesda) 21, 281-288.
Takehana A, Yano T, Mita S et al. (2004) Peptidoglycan recognition protein (PGRP)-LE and
PGRP-LC act synergistically in Drosophila immunity. The EMBO journal 23, 46904700.
Tang H (2009) Regulation and function of the melanization reaction in Drosophila. Fly 3, 105111.
Tang H, Kambris Z, Lemaitre B et al. (2006) Two proteases defining a melanization cascade in
the immune system of Drosophila. The Journal of biological chemistry 281, 2809728104.
Texier C, Vidau C, Vigues B et al. (2010) Microsporidia: a model for minimal parasite-host
interactions. Current opinion in microbiology 13, 443-449.
Theopold U, Samakovlis C, Erdjument-Bromage H et al. (1996) Helix pomatia lectin, an
inducer of Drosophila immune response, binds to hemomucin, a novel surface mucin.
The Journal of biological chemistry 271, 12708-12715.
Theopold U, Schmidt O, Soderhall K et al. (2004) Coagulation in arthropods: defence, wound
closure and healing. Trends in immunology 25, 289-294.
Tzou P, De Gregorio E, Lemaitre B (2002) How Drosophila combats microbial infection: a
model to study innate immunity and host-pathogen interactions. Current opinion in
microbiology 5, 102-110.
Tzou P, Ohresser S, Ferrandon D et al. (2000) Tissue-specific inducible expression of
antimicrobial peptide genes in Drosophila surface epithelia. Immunity 13, 737-748.
van Mierlo JT, van Cleef KW, van Rij RP (2011) Defense and counterdefense in the RNAibased antiviral immune system in insects. Methods Mol Biol 721, 3-22.
47
Wang R, Liang Z, Hal M et al. (2001) A transglutaminase involved in the coagulation system
of the freshwater crayfish, Pacifastacus leniusculus. Tissue localisation and cDNA
cloning. Fish & shellfish immunology 11, 623-637.
Wang Y, Jiang H (2004) Prophenoloxidase (proPO) activation in Manduca sexta: an analysis
of molecular interactions among proPO, proPO-activating proteinase-3, and a
cofactor. Insect biochemistry and molecular biology 34, 731-742.
Watson FL, Puttmann-Holgado R, Thomas F et al. (2005) Extensive diversity of Ig-superfamily
proteins in the immune system of insects. Science 309, 1874-1878.
Weber AN, Tauszig-Delamasure S, Hoffmann JA et al. (2003) Binding of the Drosophila
cytokine Spatzle to Toll is direct and establishes signaling. Nature immunology 4,
794-800.
Williams MJ, Habayeb MS, Hultmark D (2007) Reciprocal regulation of Rac1 and Rho1 in
Drosophila circulating immune surveillance cells. Journal of cell science 120, 502-511.
Williams MJ, Sutherland WH, McCormick MP et al. (2005) Aged garlic extract improves
endothelial function in men with coronary artery disease. Phytotherapy research :
PTR 19, 314-319.
Yahata N, Watanabe T, Nakamura Y et al. (1990) Structure of the gene encoding beta-1,3glucanase A1 of Bacillus circulans WL-12. Gene 86, 113-117.
Yajima M, Takada M, Takahashi N et al. (2003) A newly established in vitro culture using
transgenic Drosophila reveals functional coupling between the phospholipase A2generated fatty acid cascade and lipopolysaccharide-dependent activation of the
immune deficiency (imd) pathway in insect immunity. The Biochemical journal 371,
205-210.
Yano T, Mita S, Ohmori H et al. (2008) Autophagic control of listeria through intracellular
innate immune recognition in drosophila. Nature immunology 9, 908-916.
Yoshida H, Kinoshita K, Ashida M (1996) Purification of a peptidoglycan recognition protein
from hemolymph of the silkworm, Bombyx mori. The Journal of biological chemistry
271, 13854-13860.
48