For Insect Pheromone Biochemistry and Molecular Biology
Biochemistry and Evolution of OBP and CSP proteins
Jean-François Picimbon
University of Lund, Department of Ecology, Solvegatan 37, SE 62-223 Lund, Sweden
I. Introduction
One perceives from the world only what one has been prepared to perceive. In humans
and in most mammals, all different senses are used to make sense of life. In contrast, in insects,
chemical senses involving odorants and contact chemosensory molecules play the vital role. The
olfactory system is the primary sense insects use in analyzing the environment, in crucial tasks
such as finding food, nest, mates and conspecifics. Contact chemosensation is specialized to
analyze specific substrates to assist in the identification of suitable oviposition sites, the
recognition of host-plants, the selection of tastants and the search for further nutrient chemicals.
Dedicated to survival, both olfactory and contact chemosensory systems in insects have
developed to extremely high levels of sensitivity and selectivity.
The sensitivity and selectivity of olfaction and contact chemosensation are due 1) in the
brain, to the existence of neuronal network of neurons tuned to a specific chemical stimulus, and
2) in the periphery, to the existence of olfactory/chemosensory receptor neurons housed in
sensory micro-organs called sensilla. The sensilla can best be viewed as simple cuticular porous
extrusions that increase the surface that captures airborne odorants or chemicals dissolved in
water droplets. They contain the receptive olfactory or chemosensory structures (Schneider,
1969). The olfactory sensilla are most numerous on the antennae and mediate the reception of
1
sex pheromones and plant volatiles, as well as other odorants. Low volatility pheromones may
also be detected by contact chemoreceptors on the front legs (Xu et al. , 2002; Park et al. , 2002).
In contrast, general chemosensory receptor sensilla are distributed over the whole insect body,
but mainly occur on the legs, and contain neurons responding to hydrophobic tastants, CO2,
temperature, humidity or a combination of different modalities. The antennae, legs and their
sensillar complement represent a wide spectrum of structures, shapes and lengths, but the
cellular organization of sensilla follows a universal scheme based on conserved morphological
features. Most prominent is the presence of a lymphatic fluid, the sensillar lymph, that entirely
fills the sensillar lumen into which the dendrites of the sensory neurons extend. Thus, the pores
that penetrate the surface of the sensilla are not in direct contact with the receptor proteins which
reside on the sensory dendritic membrane, and chemical molecules have to cross the sensillar
lymph before to interact with the dendritic receptors. The problem of chemical reception is dual:
1) the chemical molecules (largely hydrophobic in nature) face a hydrophilic environment after
penetrating the sensillum, 2) the chemicals in the lymph are exposed to a high concentration of
chemical degrading enzymes. Specific binding mechanisms are therefore required, not only to
solubilize but also to protect and transport the odorant molecules in the sensillar lymph,
upstream to the sensory receptors (Vogt and Riddiford, 1981; 1986; Vogt et al. , 1985; Vogt,
1987).
The Odorant Binding Proteins (OBPs) and the ChemoSensory Proteins (CSPs) are
proteins from the lymph that are thought to accomplish these tasks, solubilizing and protecting
the odorant and contact chemosensory molecules. This chapter describes the biochemical and
evolutionary aspects of these two families of peripheral sensory proteins of insects. Particular
attention will be paid to the sub-classification of binding proteins, the diversity of gene structures
2
and the phyletic and molecular relatedness between binding proteins from different insect
species.
II- The family of odorant-binding proteins
A- The concept of pheromone-binding protein
About twenty years ago, Vogt proposed that small water-soluble proteins called OBPs
might aid in odor reception, by keeping the lipophilic odorants soluble and active in the lymph,
thus allowing their transport and integrity through the aqueous barrier. The first insect OBP
identified as such was the pheromone-binding protein (PBP) found in the male antennae of the
large silkmoth Antheraea polyphemus (Vogt and Riddiford, 1981; Table 1).
In pioneer binding studies, pheromone alone in a glass vial containing water quickly
absorbed to the glass wall (Kaissling et al. , 1985). However, when PBP was added a certain
amount of pheromone remained in solution. The degraded pheromone molecule was not held in
solution by PBP, indicating a degree of specificity with respect to which odorants are
solubilized. The antibody of an OBP-related protein from the blowfly Phormia regina blocks
the response of the taste receptor cell to a stimulant containing hydrophobic molecules (Ozaki et
al. , 1995). The mutation of one OBP gene from Drosophila, lush, results in abnormal
chemoattractive behavior to ethanol (Kim et al. , 1998). These results indicate that reception of
tastants and chemical molecules soluble in water requires transport of these molecules by OBPs.
It is generally assumed that at, the molecular level, the nature of the diverse chemosensory
modalities is similar to that of odorant and tastant reception and that multiple types of binding
proteins are involved in the diverse chemosensations (Vogt et al. , 1991a,b; Shanbhag et al. ,
2001; Koganezawa and Shimada, 2002).
3
The question thus is to what extent PBP and other binding proteins participate in the
recognition of odor messages? The PBPs have been so called by virtue of sex pheromone
binding, specific association with sex-pheromonal sensilla (Vogt and Riddiford, 1981; Vogt et
al., 1989; Du et al. , 1994; Steinbrecht, 1996; Steinbrecht et al. , 1992, 1995; Laue et al. , 1994).
A PBP from Bombyx mori has been crystallized with the pheromone Bombykol packed into a
hydrophobic binding pocket (Sandler et al. , 2000). So far, no binding studies have really
assessed the degree of specificity of PBPs, i.e. the ability of the carrier-protein to bind to one
specific compound. Rather, the binding spectra of PBPs seem broad. The PBPs, like other
binding proteins, will bind many chemicals but differently (Vogt et al. , 1989; Du et al. , 1994;
Feng and Prestwich, 1997; Wojtaseck et al. , 1999; Campanacci et al. , 2001a; Bette et al. ,
2002).
Differential binding representing a fine tuning of PBP-pheromone interaction has been
documented by different studies of various insect species. A PBP from male moths will bind the
chemical component of the female pheromone better than any other components. PBPs may act
as selective filters, since PBPs differentially bind specific pheromone components; pheromonePBP interactions may be based on the recognition of the chain hydrocarbons of the pheromone
molecules (Du and Prestwich, 1995; Feixas et al. , 1995; Maïbèche-Coisné et al. , 1997; Maida et
al. , 2000; Picimbon and Gadenne, 2002). The odor recognition by PBPs might be as sensitive as
that exhibited by sensory receptors. In the gypsy moth, Lymantria dispar, two PBPs have been
shown to discriminate between two enantiomeric forms of the pheromone (Vogt et al. , 1989;
Plettner et al. , 2000). Given such a potential for greater binding and pheromone specificity
allowed by a duplication of PBPs, the evolution of PBP genes may have followed the
diversification of the pheromone systems.
4
B. The repertoire of PBPs in moths
The most prominent examples of pheromone diversification are found in Noctuidae. The
sex pheromones of the Noctuidae species are mixtures of at least three compounds that differ
mainly in carbon chain length (Lofstedt et al. , 1982; Teal et al. , 1986; Attygale et al. , 1987;
Picimbon et al. , 1997). In many noctuid moth species, multiple PBPs has been reported (Vogt et
al. , 1989; Merritt et al. , 1998; Maïbèche-Coisné et al. , 1998; Picimbon and Gadenne, 2002;
Abraham et al. , 2002). Based on sequence homology and phylogenetic analysis, a subclassification of noctuid PBPs has been proposed (Picimbon and Gadenne, 2002; Abraham et al.
, 2002).
A neighbor joining tree of selected PBPs from moths shows that noctuid PBPs segregate
into two sub-classes (Fig. 1). Sub-class 1 (group 1 or Grp1) corresponds to Grp1-PBPs from
Agrotis ipsilon, A. segetum, Heliothis virescens, H. zea and Mamestra brassicae (Aips-1, Aseg1, Hvir-1, Hzea-1 and Mbra-2). The Grp1-PBPs show about 86% identity between each other.
The sub-class 2 (group 2 or Grp2) corresponds to Grp2-PBPs from Agrotis, Heliothis and
Mamestra species (Aips-2, Aseg-2, Hvir-2, Mbra-1; Fig. 1A) as well as proteins from nonnoctuid species such as Manduca sexta (Msex-2, Msex-3; Robertson et al., 1999; Picimbon and
Gadenne, 2002). Sub-class 2 also contains Ycag PBP, a PBP from Yponomeuta cagnagellus
(Robertson et al. , 1999; Willett, 2002; Picimbon and Gadenne, 2002). The noctuid Grp2-PBPs
show about 72% identity between each other, about 50% to non-noctuid Grp2-PBPs and only
32-47% to other PBPs. The specific grouping of orthologous PBPs strongly suggest that different
types of PBP are likely utilized by most species of moth and that within a species multiple
subtypes are expressed, perhaps for binding a large repertoire of pheromone molecules.
5
Yponomeutidae and Noctuidae are phylogenetically distant. It is very likely that Grp2PBPs are expressed by other Lepidoptera lineages that share common ancestry with these two
Families, and that these PBPs are tuned to a pheromone structures conserved across the different
member species. In contrast, the Grp1-PBPs have been reported only in Noctuidae species.
However, Antheraea pernyi, Ostrinia nubilalis, Manduca sexta and Lymantria dispar use
pheromone components structurally similar to the main pheromone components of Noctuidae
(Klun et al. , 1973; Bestmann et al. , 1987; Tumlinson et al. , 1989; Gries et al. , 1996).
Therefore, it cannot be excluded that pyralid, sphingid and lymantrid species also express Grp1
types of PBP and that these have simply not yet been identified. The PBPs so far identified in
these species correspond to very specific groups of protein (Merritt et al. , 1998; Willett and
Harrisson, 1999; Robertson et al. , 1999; Vogt et al. , 1999; Picimbon and Gadenne, 2002). In
particular, the two PBPs from L. dispar, Ldis-1 and Ldis-2, are very divergent from other moth
PBPs. Interestingly, they both preferentially associate to enantiomers of disparlure, an epoxyde
component used as primary pheromone by L. dispar (Plettner et al. , 2000). It could well be that
in L. dispar, primary PBPs bind to disparlure and secondary PBPs binds to minor pheromone
components.
An electrophoretic analysis of antennal proteins has failed to find secondary PBPs and
only Ldis-1 and Ldis-2 appear to be detectable by a biochemical approach. Similarly, in the
noctuid M. brassicae, only protein bands corresponding to Grp1 and Grp2 PBPs have been
identified (Nagnan-Le Meillour, 1996). In the bombycid B. mori, only one PBP protein could be
found using either a biochemical approach or homology screening of a cDNA library (Maida et
al. , 1993; Krieger et al. , 1996). The failure to find additional PBPs in these species may be due
to the fact that the degree of expression is markedly different for each PBP. A primary PBP
6
tuned to high concentrations of a primary pheromone component may be expressed more than
secondary binding proteins tuned to lower concentrations of secondary pheromone compounds.
Analyzing the presence of Grp-1 and Grp-2 PBPs in non-noctuid insects will be a first step to
investigate PBP diversity with respect to recognition of multicomponent pheromone blends.
The Grp1 and Grp2 PBPs have been identified in two closely related noctuid species,
Agrotis ipsilon and A. segetum, whose females emit pheromone blends that consist of three main
components and minor components that vary locally. The major pheromone component of A.
ipsilon is (Z)-11-hexadecenyl acetate (Z11-16:Ac), while the major pheromone component in A.
segetum is (Z)-5-decenyl acetate (Z5-10:Ac). The antennae from A. ipsilon and A. segetum both
have neurons responding to (Z)-7-dodecenyl acetate (Z7-12:Ac) and (Z)-9-tetradecenyl acetate
(Z9-14:Ac), two of their major pheromone compounds (Löfstedt et al. , 1982; Toth et al. , 1992;
Picimbon, 1995,1998; Picimbon et al. , 1997; Gadenne et al. , 1997; Gemeno and Haynes, 1998;
Wu et al. , 1999). The Grp-1 PBPs of these two species, Aips-1 and Aseg-1, are virtually
identical (Laforest et al. , 1999; Picimbon and Gadenne, 2002). The Grp-2 PBPs from these
species are more divergent: Aips-2 and Aseg-2 show only 76% identity. We could speculate that
the conserved Aips-1/Aseg-1 proteins bind either Z7-12:Ac or Z9-14:Ac and that the more
variable Aips-2 and Aseg-2 bind to Z11-16:Ac and Z5-10:Ac respectively.
C- Gene structures encoding OBPs
To explore the functions of Grp-1 and Grp-2 PBPs in depth, cutting edge protein
expression, ligand-binding, structural analysis and immunocytochemistry experiments are
required. However, determination of the gene structures might also be informative since the
7
diversification of species and of their pheromone systems may have led to specific gene
regulations of PBP expression.
The genes encoding Aips-1, Aips-2, Aseg-1 and Aseg-2 have been characterized from
genomic DNA (Abraham et al. , 2002; Fig. 2). All four share the same two introns - three exons
structure but differ in length. The three exons encode similar portions of the protein. The first
exon (exon 1) is the shortest and has the same size in all Agrotis PBP genes. This exon pattern
may be a general feature across the Grp-PBP genes. The first intron of these genes exhibit little
variability. However, intron 2 of both A. ipsilon PBPs are significantly longer than the introns of
the corresponding A. segetum PBPs. Phylogenetical analysis of the introns from the four
Agrotis genes suggests that the Aips-1/Aseg-1 and Aips-2/Aseg-2 are respectively closely
related, and that these Grp1 and Grp2 genes may have evolved from gene duplication that
occurred before the divergence of A. ipsilon and A. segetum. This duplication may have
occurred even earlier, before the split of Yponomeutoidea, Sphingiodea and Noctuoidea lineages
as suggested by the presence of Grp2-PBPs in Y. cagnagellus and M. sexta.
In the sphingid M. sexta, Msex-1 is similar to the PBP from B. mori and divergent from
the Grp2 proteins (Vogt et al. , 2002). The Msex-1 gene displays the same structure of three
exons - two introns as the noctuid PBP genes encoding Aseg-1/Aips-1 and Aseg-2/Aips-2 as
well as other non-noctuid PBP genes (Krieger et al. , 1991; Willett and Harrisson, 1999; Willett,
2000; Abraham et al. , 2002). Therefore, the lepidopteran PBP genes have a conserved pattern
of two introns and three exons with a variability being observed in the intron length.
8
In contrast, the PBP-related proteins (PBPRPs) from Drosophila melanogaster are
highly variable with respect to exon-intron structure. The gene encoding the protein PBPRP5
has a single coding exon, while the genes encoding PBPRP1 and PBPRP2 have four and five
coding exons, respectively. The gene encoding the OS-F protein has 4 exons of varying size and
a very long second intron (McKenna et al. , 1994; Pikielny et al. , 1994; Hekmat-Scaffe et al. ,
1997, 2000). The genes encoding the olfactory proteins OS-E and LUSH exhibit intron-exon
patterns similar to those of lepidopteran PBP genes, suggesting that OS-E and the ethanolbinding protein LUSH from D. melanogaster and the moth PBPs may have a common ancestor.
D- Relationships of Moth and Drosophila melanogaster OBPs
This hypothesis may imply the existence of OS-E and LUSH orthologs in some moth
species. A protein similar to LUSH has been identified from the antennae of A. ipsilon
(Picimbon, unpublished data). In addition, the PBPRPs from D. melanogaster display
significant similarities to a specific subclass of moth OBPs that includes binding proteins whose
function is unknown, the so-called Antennal Binding Proteins-X (ABPX).
The ABPX proteins have highly conserved amino acid sequences across different moth
species and the overall ABPX sequence displays significant similarity with DmelPBPRP1 (Fig.
3). In particular, the proteins AipsABPX-1 and PBPRP1 share common amino acid residues
including the six cysteines characteristic of OBPs and the motifs 23-TGA-25, 89-SCGTQ-93 and
99-CDTA-102. The ABPX-1s and PBPRP1 may then represent an OBP-1 type of protein
defined by key residues that may underlie specific functions. The ABPX/PBPRP-specific amino
acid residues Arginine at position 16, Lysine at position 47, Proline at position 76, Threonine at
position 92 are replaced respectively by Leucine, Glycine, Lysine and Lysine residues that are
9
conserved in the different types of binding protein from B. mori (Krieger et al. , 1993; Picimbon,
2001). These replacements may be relevant to support the function of ABPX. Based on the
crystal structure of the bombykol-PBP complex, the Tryptophane at position 101 and the Valine
105 have been shown to contact the molecule of Bombykol (Sandler et al. , 2000). These are
replaced by two Threonine residues characteristic of OBP-1. Therefore, the Threonine residues
characteristic of OBP1s may be of crucial importance for the binding specificities of these
proteins.
The notion of relatedness between ABPXs of moths and the PBPRPs of D. melanogaster
(DmelPBPRPs) is strongly supported by the identification of additional ABPXs in the moth
species A. ipsilon: AipsABPX-2 and AipsABPX-3. The protein ABPX2 has significant
similarity to DmelPBPRP2 and DmelPBPRP5 on the basis of specific amino acids that may
represent an OBP-2 group, while AipsABPX-3 appears more similar to DmelPBPRP4 and may
represent an OBP-3 group (Picimbon et al. , unpublished). These relationships suggest the
existence of a multiplicity of ABPX in moths similar to that of PBPRP in flies and may indicate
a unique importance in insect olfactory behaviors. Alternatively one could speculate that ABPXs
may represent "intermediary" molecules between dipteran PBPRPs and lepidopteran PBPs.
A phylogenetic analysis of insect OBPs focusing on moth ABPXs and DmelPBPRPs
shows the ABPXs falling outside the groups of DmelPBPRPs, reflecting the phylogenetic
distance between Lepidoptera and Diptera (Fig. 4). The ABPXs from Agrotis ipsilon cluster
with the ABPX proteins from other moth species. In particular, ABPX-1s from A. ipsilon and H.
virescens group with ABPXs of B. mori, A. pernyi and M. sexta and not with other identified
noctuid PBPs. There are well supported separations between the branches containing A. ipsilon
10
proteins. One could speculate that these A. ipsilon genes represent multiple rearrangements or
duplications of the ABPX ancestral gene and that the events which produced the ABPX diversity
are common in noctuid species. Similar duplications may have occurred to produce diverse
PBPRPs in D. melanogaster but it seems that ABPX and PBPRP genes have evolved
independently, despite established similarities in there sequences.
As the PBPRP genes are quite different in sequence from the moth PBP genes, and as the
Drosophila and moth olfactory genes diverged very early in the evolutionary course of these
insects, moth PBPs in moths may have evolved for binding pheromones, while ABPX may have
persisted and developed for binding more generalist odorants. Identification of the genes
encoding AipsABPX and studies of the structure/activity relationships of ABPX/PBPRP need to
expand clarify the evolutionary and functional relationships between PBPRP, ABPX and PBP
proteins.
III- The family of chemosensory proteins
A- The concept of Chemosensory Protein
In the context of evolution of pheromone olfaction and general chemosensation, we can
speculate that general chemosensory proteins may be more highly conserved than olfactory PBPs
when compared across species. Indeed, little sequence diversification would be expected to
occur among genes encoding chemosensory proteins tuned to bind chemicals of common
importance to all species.
In insects, a class of putative general chemosensory proteins has been described and has
gained increasing interest over the last few years, expanding our understanding of the complexity
11
of the repertoire of sensory binding proteins. The first member of this novel class of proteins was
found in Drosophila melanogaster and called OS-D (Olfactory Specific-protein type D) or A10
(McKenna et al. , 1994; Pikielny et al. , 1994); OS-D is abundant in sensory appendages and
contains 4 cysteines. Similar proteins (Table 2) have been since identified in several species and
variously referred to as OS-Ds, SAPs (Sensory Appendage Proteins) or CSPs (Chemosensory
Proteins) (Danty et al. , 1998; Angeli et al. , 1999; Picimbon and Leal, 1999; Robertson et al. ,
1999; Picimbon et al. , 2000a). The first strong evidence that these proteins have a role in
chemosensation came from immunocytochemistry experiments in the grasshopper Schistocerca
gregaria that demonstrated OS-D protein in the lymph of the contact chemosensory sensilla
(Angeli et al., 1999).
B- Comparison of CSPs to OBPs
On the basis of the immunocytochemical localization of CSP, Angeli et al. (1999) have
suggested that CSPs have an OBP-like function. However, the CSPs cannot be regarded as
OBPs considering the basic definition of an OBP: a lymphatic, acidic α-helical, mainly
hydrophobic carrier protein of 14-16 Kds characterized by six cysteines linked by three disulfide
bridges, a flexible structure and an expression pattern restricted to the olfactory sensilla of the
antennae (Vogt and Riddiford, 1981; Breer et al. , 1992; Prestwich et al. , 1995; Steinbrecht,
1996; Wojtasek and Leal, 1999; Scaloni et al. , 1999; Leal et al. , 1999; Campanacci et al. , 1999;
Sandler et al. , 2000; Kowcun et al. , 2001; Horst et al. , 2001). The CSPs are lymphatic acidic
α-helical proteins, but 1) they have a molecular weight in the range of 12-13 Kds, 2) they show a
set of only four conserved cysteines and two disulfide bridges, 3) they are highly hydrophilic, 4)
they show high structural stability, and 5) they are not specific to the antennae but also found in
12
other parts of the insect body, more particularly in the legs (Angeli et al. , 1999; Picimbon et al. ,
2000a,b, 2001; Picone et al. , 2001; Campanacci et al. , 2001b).
The developmental patterns and structures of CSPs and OBPs are also different. The
CSPs are produced synchronously to the shedding of the cuticle, very early during adult
development, in contrast to OBPs which are produced late during adult development. This
demonstrates that the chemosensory CSPs and olfactory OBPs are controlled by independent
mechanisms (Vogt et al., 1993; Picimbon et al., 2001; Gavillet and Picimbon, 2002). Moreover,
X-Ray structure analysis of CSPs has revealed a novel type of α-helical fold with six helices
connected by α-α-loops and a narrow channel expanding over the protein hydrophobic core.
This structural feature may confer specific binding properties to CSPs, in particular the ability to
interact with long linear acyl chains of hydrophobic components (Lartigue et al. , 2002).
C- The repertoire of CSPs
A phylogenetic analysis of currently known insect CSPs, based on amino acid sequence
comparisons, is shown in Fig. 4B. With few exceptions, proteins from different insect Orders
segregate to different branches, consistent with the phylogenetic distance between these Orders.
The CSPs from moths segregate into 3 groups, noted as CSP1, CSP2 and CSP3; each group
includes taxa from multiple lepidopteran Families. Three CSPs from H. virescens, which
segregate to these 3 groups, share about 50% amino acid identity between each other; similar
between-group identities are seen for proteins of M. sexta and M. brassicae (Robertson et al. ,
1999; Nagnan-Le Meillour et al. , 2000; Picimbon et al. , 2001; Jacquin-Joly et al. , 2001).
Within-group identities for the M. brassicae are unusually high. The CSP1 proteins
CSPMbraA4 and CSPMbraA1 differ by only two residues, and are virtually identical to
13
CSPMbraA3/A6, CSPMbraA2 and CSPMbraA5. Similarly, the CSP2 M. brassicae proteins
MbraCSPB1 and MbraCSPB3 differ by only one amino acid. These proteins may actually be
alleles representing the same locus (Picimbon et al., 2000a; Nagnan- Le Meillour et al. , 2000;
Jacquin-Joly et al. , 2001). The CSP1 group, and especially BmorCSP1, attracted a protein from
the phasmid E. calcarata ( 39% identity) even though moths and phasmids are quite distant
phylogenetically (Picimbon et al. , 2000b; Marchese et al. , 2000).
CSPs of the orthopteroids (C. cactorum, Locusta migratoria, S. gregaria) are highly
conserved within species and with few exceptions (e.g. CLP1) separate into species specific
groups. One CSP group per orthopteroid species is in sharp contrast with moths where CSPs
from a given species fall into three groups. However, the S. gregaria protein SgreOS-D1 is
attracted to the CSPs of L. migratoria, sharing 79% identity with the L. migratoria CSPs but
only 55% identity to the other CSPs from S. gregaria (SgreCSPs), so perhaps there are multiple
orthopteroid CSP classes that have simply not been much identified.
Analysis of the D. melanogaster and A. gambiae genomes reveals the full repertoire of
CSPs within single species. D. melanogaster has 6 highly divergent CSPs; amino acid sequence
identities of PEBmeIII, DmelOS-D and RH74005 range from 16-23% and from 14-23% when
compared with the moth proteins. DmelOS-D is somewhat more similar, sharing about 45%
sequence identity with the moth proteins. Four of the 7 A. gambiae CSPs cluster in a single
group and share about 72% identity, but the others segregate by themselves. Although A.
gambiae and D. melanogaster are both Diptera, only a single pair of CSPs share significant
similarity to be attracted together (agCG5020865 and CG30172, 65% identity); these may be
14
orthologs. Overall, the divergence seen in CSP sequences is intriguing; divergent CSPs may
have different and specific binding properties for distinct chemical ligands.
D- Structural and binding properties of CSPs
Differences in the amino acid sequences of CSPs presumably relate to differences in the
ligand binding properties of each protein (Fig. 5). The type-1 and 2 CSPs of moths (CSP1 and
CSP2) are proteins with about 107-112 amino acids and all exhibit the diagnostic elements of
CSP: Aspartatic acid 6 and 88, Lysine 43, Glutamine 61, Proline 89 and four Cysteines at
positions 29, 36, 54 and 57. Overall, the CSP1 proteins are highly conserved, and most have the
sequence (1) -D-YTDKYD-----EIL-N-RLL--Y--CV---GKC--EGKELK--L--A---GC-KC---QEG----I--LIKN----W--L----DP---WR-KYEDRA-A-GI-IP-- (110). Six isoforms (alleles?) of
CSPMbraA proteins, all CSP1s, differ at only three sites, 59, 69 and 92 (Ala/Thr, Ala/Val and
Val/Gly). The CSPMbraA proteins differ from other CSP1 proteins in being one amino acid
longer (Glu ) and have the C-terminal sequence DRAKAAGIVIPEE (Nagnan- Le Meillour et al.,
2000; Jacquin-Joly et al., 2001). The three dimensional structure has been determined for
CSPMbraA6. This CSP protein binds aliphatic molecules with 12-18 carbons, suggesting that
CSPs generally bind hydrophobic compounds (Lartigue et al. , 2002). Thus, multiple subtypes
of CSP may mediate transport of carbon chains of different lengths. If CSP1s bind specific
hydrophobic chains, the large repertoire of proteins may bind the large number of diverse alkyl
chain components. We could hypothesize that a small CSP1a protein would bind to a C12
molecule, while the longer CSP1b protein would bind to molecules longer than C12.
Other types of CSP may have totally different binding properties. The CSP2s are
characterized by diagnostic CSP residues and by other conserved amino acids that may underlie
15
specific functions. Overall, the CSP2 proteins are highly conserved, and most have the sequence
(1) ----YTD-YD-V-LDEIL-N-R--VPY-KCILD-GKCAPD-KELKEHI-EALE-ECGKCT--QKGGTRRVI-HLINHE---W-EL--K-DP--K---KYEKEL---K- (108). The amino acid pattern (1) ---YTD-D-----EIL-N-R----Y--C----GKC----KELK----------C-KC---Q--G----------N-----W--L----DP-----K------- (106) is conserved in CSP1 and CSP2. CSP3s are so far only represented by the
proteins HvirCSP2 and SAP4 which share 84% sequence identity but are only about 50%
identical to other CSP1 or CSP2. These CSP3s have the sequence
(1) ASTYTDKWDNINVDEILESXRLXKXYVDCLLDXGRCTPDGKALKETLPDALEXXCS
KCTEKQKAGS-KVIR-LVNKR--LWKELSAKYDPNN-YQ--YKDKIXXXKQG- (106/107).
The α-helical regions of the moth CSP1, CSP2 and CSP3 proteins differ in amino acid
sequence. The α-helix (α1) in the N-terminal region anchors a narrow hydrophobic channel and
includes residues 5 (or 12)-18; α1 is conserved in all three CSP groups with the sequence YTD-D-----EIL (Fig. 5). In CSP1s, α1 interacts with residues Tyr 8 and Glu 39 to stabilize the
hydrophilic channel; however, CSP2s lack Glu 39. And in CSP3s, a non-polar residue Trp
replaces the polar, uncharged Tyr 8. Residue Glu 39 of CSP1s is replaced by Asp in CSP2s and
CSP3s. In contrast, the α-helix (α6) in the C-terminal region is highly variable and may support
ligand specificity. Furthermore, of thirteen residues of CSPMbraA6 that have been shown to
interact with water molecules, only six are conserved in CSP1s (Asp 9, Tyr 26, Leu 43, Leu 47,
Glu 63 and Gly 65) and only the residues Asp 9, Tyr 26 and Leu 43 are found in all three types
of moth CSP, while the other residues are highly variable (Fig. 5). Lartigue et al. (2002)
suggested that Tyr 26 may rotate towards the protein surface upon ligand binding, and that the
previous position of its hydroxyl group may then be occupied by the side chain of Leu 43. Since
16
both residues Tyr 26 and Leu 43 are highly conserved in all insect CSPs, this mechanism may be
common for all CSPs.
Two tryptophan residues from MbraA6 were proposed to interact with ligands; these are
conserved in all CSP1s, but only one (Trp 82) is conserved throughout the CSPs, suggesting that
different CSPs may bind different ligands. The overall sequence similarities of CSP1s, CSP2s
and CSP3s suggest that they all bind to lipid compounds. However, the diversity of CSPs not
only in moths but also in other insect species, suggests that CSPs transport diverse types of
chemicals. The CSPs of D. melanogaster, A. gambiae, B. mori and E. calcarata are highly
variable, sharing only 40% sequence identity (Picimbon et al. , 2000b; Marchese et al. , 2000;
this chapter). The CSPs from M. sexta (SAP1 and SAP5) are also very divergent (Robertson et
al. , 1999; this chapter). Different types of CSP with multiple isoforms are also found in the
acridid species L. migratoria and S. gregaria (Angeli et al. , 1999; Picimbon et al. , 2000a;
Picimbon, 2001; see Fig. 5).
Analyzing the cysteine arrangements in a CSP from S. gregaria, Angeli et al. have shown
that the two adjacent cysteines form two small loops along the main protein chain in a fashion
similar to thioredoxins. Thus, CSP may well play a role not only in binding lipid molecules but
also in CO2 sensing (Bogner et al. , 1986; Stange, 1992, 1996; Maleszka and Stange, 1997;
Angeli et al. , 1999).
17
E- Multiple functions of CSPs
1. Tissue-distribution
CSPs are found not only in external sensory organs, but also in external and internal nonsensory tissues, further suggesting the proteins may have diverse functions. In moths the
distributions of CSPs have been characterized by Northern blot analysis. A DIG-RNA probe
encoding the BmorCSP1 hybridized not only with mRNAs from male and female antennae but
also with mRNAs from legs and other parts of the insect body. Northern blot analysis using
BmorCSP2 showed a very similar distribution pattern with no tissue specificity, as did the three
CSPs of H. virescens which showed high levels of expression in legs (Picimbon et al. , 2000a,
2001). These results are consistent with the isolation of CSP cDNAs and proteins. The CSP of
C. cactorum was as a cDNA from labial palps. The CSPs of M. Brassicae were isolated as
cDNAs from antennae and pheromone glands, and as proteins ( N-terminal sequences) from
proteins antennae, proboscis and leg (Malezska and Stange, 1997; Bobhot et al. , 1998; NagnanLe Meillour et al. , 2000; Jacquin-Joly et al. , 2001).
CSPs are also broadly distributed in tissues of insects other than moths. In Drosophila
melanogaster, OS-D, also called A10, is extremely abundant in the antennal appendages, and the
CSP EBSP-III, is expressed in the ejaculatory bulb (McKenna et al. , 1994; Pickielny et al. ,
1995; Dyanov and Dzitoeva, 1995). In locusts, CSPs are expressed in male and female from
adults as well as 5th instar larvae, and are found in many different tissues including antennae,
legs, mouth organs, thorax, abdomen, head and wings (Picimbon et al. , 2000b; in preparation).
In the phasmid Eurycantha calcarata, a CSP has been isolated from the cellular layer
underlying the cuticle (Marchese et al. , 2000). In the cockroach Periplaneta americana, the CSP
p10 is expressed in legs and antennae (Nomura et al. , 1992; Kitabayashi et al. , 1998; Picimbon
18
et al. , 2001), and other CSPs have been detected in tissues including legs, brain and cerci
(Picimbon and Leal, 1999; unpublished).
Thus, in various insect species, the expression of CSP occurs in many different tissues,
especially in the legs and contact sensory organs. The differences between CSPs and OBPs in
tissue distribution and developmental expression suggest that the roles of these proteins in
chemoreception may also be different. CSPs may have a broader function than OBPs,
functioning in more systems than olfaction and taste, perhaps as general molecule carriers but
especially involved in the transport of contact sensory molecules.
2. Role in contact chemosensation
Immunocytochemistry experiments performed with rabbit antiserum generated against a
CSP isoform from Schistocerca gregaria (sgCSP-5) showed selective labeling of the outer
lymph of the peg lumen and in the cavity below the peg base of contact sensilla from the
antennae and mouth organs, such as maxillary palps and tarsi. Olfactory sensilla were not labeled
(Angeli et al. , 1999). This pioneer work showed the localization of CSPs in the lymphatic space
surrounding the contact chemosensory receptor neurons and suggested an OBP-like function for
CSPs in contact chemosensation. In the context of OBP-like function, the CSPs could well
mediate the clearance of hydrophobic odorant molecules absorbed haphazardly by the cuticle as
well as the delivery of these odorant molecules to degradative enzymes (Ferkovitch et al. , 1982;
Lonergan, 1986; Prestwich et al. , 1989). Clearance and degradation of odorant molecules has to
occur, e.g. while the insect flies in a pheromone plume or feeds on nectar. More data regarding
19
the specific localization of CSPs in chemosensory and non-chemosensory organs would provide
a better understanding of the OBP-like function of these proteins.
3. Function and evolutionary history
In the cockroach Periplaneta americana, CSP have been found that differentially express
between the sexes, suggesting a role in delivering conspecific pheromones to olfactory neurons
(Picimbon and Leal, 1999). In the phasmid Carausius morosius, a CSP is differentially
expressed in only a subset of olfactory sensilla (Monteforti et al. , 2002). Multiple CSPs have
been detected in phasmids, locusts and cockroaches, but no proteins related to known PBPs have
been found (Tuccini et al. , 1996; Mameli et al. , 1996; Picimbon and Leal, 1999; Picimbon et al.
, unpublished). These observations do not exclude the presence of RNA encoding PBPs in these
insects, but RNA may not be necessarily translated into proteins (Segal et al. , 2001). Since
Phasmatodea, Acridoidea and Blattoidea are among the most primitive insect Orders, one could
hypothesize that CSPs perform PBP functions in ancient insects. Later in evolution, CSPs may
have developed into general chemosensory proteins in flies and moths, where their original
function was replaced by more efficient pheromone-binding proteins. This hypothesis could be
tested by analyzing the diversity of CSP genes across a variety of insect species and by
comparing the CSP and OBP genes.
The genes underlying olfaction and general chemosensation are, indeed, most likely
under different evolutionary pressures that may act selectively through defined regulatory
elements. Specific intron structures, like the intron 2 of PBPs, may be key targets for regulatory
mechanisms that control the differential expression or loss of specific OBPs and CSPs. What
20
gene structures will be found in the various insect species that use different pheromone and
chemosensory components?
IV- Concluding remarks and perspectives
Binding of hydrophobic molecules by specific protein carriers appears to be a very
efficient mechanism to increase both solubility and transport these molecular messengers in a
hydrophilic medium. OBPs and CSPs may represent a successful application of this principle.
In particular, the molecular mechanisms of transport of hydrophobic molecules may be more
ancient than that most ancient of senses, olfaction. The olfactory system may have developed to
extract the hydrophobic odorants from the air environment and optimize their transport and
delivery to sensory cells.
In recent years, studies of diverse insect species have revealed the heterogeneity of the
family of binding proteins, and thus challenged the dogmatic concept that odorant-binding
proteins are expressed only in olfactory structures. In particular, an increasing number of
binding proteins related to OBPs have been identified in various non-sensory organs, such as the
hemolymph, brain, accessory and salivary glands (TH12 proteins, sericotropin, B proteins and
D7 gene products) (Paesen and Happ, 1995; Kõdrick et al. , 1995; Thymianou et al. , 1998;
Rothemund et al. , 1999; Graham et al. , 2001). Such distributions suggest that this large family
of carrier proteins may perform diverse functions throughout the insect body, paralleling the
distribution and functional breadth of lipocalins from vertebrates (Flower, 2000; Ganfornica et
al., 2000). OBPs may have arisen as general transporters of hydrophobic molecules, and later
developed to bind and solubilize odorant molecules. The CSPs, which are expressed all over the
body, have certainly conserved a functional polyvalence. In contrast, the specific expression of
21
many OBPs in the antennae strongly support an adaptation of these proteins to the reception of
hydrophobic odorants.
Efforts should be made to utilize the most modern techniques to analyze the binding
properties and tissue specificity of all identified proteins, and eventually to rename the proteins
on the basis of specific groupings and conserved motifs of amino acid. This suggestion is
intended to be neither provocative nor inflammatory, but certainly the most reasonable way to
define function of diverse binding proteins with respect to the pheromone systems and life
history of the different insect species.
Genetic studies will undoubtedly lead to the elucidation of how specific transporter
molecules have developed to a fine-tuned function in olfaction. Specific and non-specific
transporter molecules might have evolved differently over the course of evolutionary history and
the diversification of species. If there is one thing future research should accomplish, it would
be the unveiling of the evolution of the families of CSPs and OBPs. In addition, it must be
considered that other families of binding proteins may well exist and participate to the reception
of the extremely large repertoire of odor and chemosensory molecules. Investigations into these
matters would have a strong impact not only in fundamental genetics underlying chemosensation
and olfaction but also in applied industry, assuming that gene manipulation is accessible and
permits control of the expression of OBP and CSP, and thereby control of the sensory abilities of
insect pests (Picimbon, 2001, 2002).
Acknowledgements: My heartfelt thanks to Profs. R.G. Vogt and P. Pelosi who discovered the
sensory binding proteins simultaneously in insects and vertebrates.
22
REFERENCES
Abraham, D. , Löfstedt, C. and Picimbon, J.F. (2002). Molecular evolution and gene
characterization of Grp1 and Grp2 Pheromone Binding Proteins in moths. Genetics (Manuscript
submitted for publication).
Adams, M.D. , Celniker, S.E. , Holt, R.A. et al. (2000). The genome sequence of Drosophila
melanogaster. Science 287, 2185-2195.
Angeli, S. , Ceron, F. , Scaloni, A. , Monti, M. , Monteforti, G. , Minnocci, A. , Petacchi, R. ,
Pelosi, P. (1999). Purification, structural characterization, cloning and immunocytochemical
localization of chemoreception proteins from Schistocerca gregaria. Eur. J. Biochem. 262, 745754.
Arca, B. , Lombardo, F. , Lanfrancotti, A. , Spanos, L. , Veneri, M. , Louis, C. and Coluzzi, M.
(2002). A cluster of four D7-related genes is expressed in the salivary glands of the African
malaria vector Anopheles gambiae. Insect Mol. Biol. 11, 47-55.
Attygale, A.B. , Herrig, M. , Vostrowsky, O. and Bestmann H.J. (1987). Technique for injecting
intact glands for analysis of sex pheromones of Lepidoptera by capillary gas chromatography:
Reinvestigation of pheromone complex of Mamestra brassicae. J. Chem. Ecol. 13, 1299-1311.
Bestmann, H.J. , Attygale, A.B. , Brosche, T. , Erler, J. , Platz, H. , Schwarz, J. , Vostrowsky, O.
, Cai-Hong, W. , Kaissling, K.E. and Te-Ming, C.Z. (1987). Identification of three sex
23
pheromone components of the female Saturniid moth Antheraea pernyi (Lepidoptera:
Saturniidae). Z. Naturforsch. 42c, 631-636.
Bette, S. , Breer, H. and Krieger, J. (2002). Probing a pheromone binding protein of the
silkworm moth Antheraea polyphemus by endogenous tryptophan fluorescence. Insect Biochem.
Mol. Biol. 32, 241-246.
Biessmann, H. , Walter, M.F. , Dimitratos, S. and Woods, D. (2002). Isolation of cDNA clones
encoding putative odourant binding proteins from the antennae of the malaria-transmitting
mosquito, Anopheles gambiae. Insect Mol. Biol. 11, 123-132.
Bogner, F. , Boppre, M. , Ernst, K.D. et Boeckh, J. (1986). CO2-sensitive receptors on labial
palps of Rhodogastria moths (Lepidoptera: Arctiidae): physiology, fine structure and projection.
J. Comp. Physiol. A 158, 741-749.
Bohbot, J. , Sobrio, F. , Lucas, P. , Nagnan-Le Meillour, P. (1998). Functional characterization
of a new class of Odorant Binding Proteins in the moth Mamestra brassicae. Biochem. Biophys.
Res. Comm. 253,489-494.
Breer, H. , Boekhoff, I. , Krieger, J. , Raming, K. , Strotmann, J. and Tareilus, E. (1992).
Molecular Mechanism of Olfactory Signal Transduction. In: “Sensory transduction” (Corey D.P.
and Roper S.D. , eds), pp. 94-108. The Rockfeller University Press, New York.
24
Campanacci, V. , Longhi, S. , Nagnan-Le Meillour, P. , Cambillaud, C. and Tegoni M. (1999).
Recombinant pheromone binding protein 1 from Mamestra brassicae (MbraPBP1). Functional
and structural characterization. Eur. J. Biochem. 264, 707-716.
Campanacci, V. , Krieger, J. , Bette, S. , Sturgis, J.N. , Lartigue, A. , Cambillau, C. , Breer, H.
and Tegoni, M. (2001a). Revisiting the specificity of Mamestra brassicae and Antheraea
polyphemus pheromone binding proteins with a fluorescence binding assay. J. Biol. Chem. 276,
20078-20084.
Campanacci, V. , Mosbah, A. , Bornet, O. , Wechselberger, R. , Jacquin-Joly, E. , Cambillaud, C.
Darbon, H. and Tegoni, M. (2001b). Chemosensory protein (CSP2) from the moth Mamestra
brassicae: Expression and secondary structure from 1H and 15N NMR. Eur. J. Biochem. 268,
4731-4739.
Danty. E. , Arnold, G. , Huet, J.-C. , Masson, C. , Pernollet, J.-C. (1998). Separation,
characterization and sexual heterogeneity of multiple putative odorant-binding proteins in the
honey bee Apis mellifera L. (Hymenoptera: Apidea). Chem. Senses 23, 83-91.
Du, G. , Ng, C.S. and Prestwich, G.D. (1994). Odorant binding by a pheromone binding proteinactive site mapping by photoaffinity labeling. Biochemistry 33, 4812-4819.
Du, G. and Prestwich, G.D. (1995). Protein structure encodes the ligand binding specificity in
pheromone binding proteins. Biochemistry 34, 8726-8732.
25
Dyanov, H.M. and Dzitoeva, S.G. (1995). Method for attachment of microscopic preparations on
glass for in situ hybridization, PRINS and in situ PCR studies. BioTechniques 18, 822-824.
Feixas, J. , Prestwich, G. and Guerrero, A. (1995). Ligand specificity of pheromone binding
proteins of the processionary moth. Eur. J. Biochem. 234, 521-526.
Felsenstein, J. (1985). Confidence limits on phylogenies: an approach using the bootstrap.
Evolution 39, 783-791.
Ferkovich, S.M. , Oliver, J.E. et Dillard, C. (1982). Pheromone hydrolysis by cuticular and
interior esterases of the antennae, legs, and wings of the cabbage looper moth Trichoplusia ni
(Hubner). J. Chem. Ecol. 8: 859-866.
Flower, D.R. (2000). Beyond the superfamily: the lipocalin receptors. Biochem. Biophys. Acta
18, 327-336.
Gadenne, C. , Picimbon, J.F. , Bécard, J.M. , Lalanne-Cassou, B. et Renou, M. (1997).
Development and pheromone communication systems in hybrids of Agrotis ipsilon and Agrotis
segetum (Lepidoptera: Noctuidae). J. Chem. Ecol. 23, 191-209.
Ganfornica, M.D. , Guttierrez, G. , Bastiani, M. , Sanchez, D. (2000). A phylogenetical analysis
of the lipocalin protein family. Mol. Biol. Evol. 27, 405-412.
26
Gavillet, B. and Picimbon, J.F. (2002). Endocrine regulation of olfaction and chemosensation.
Proc. Eur. Comp. Endocrinol. in press.
Gemeno, C. and Haynes, K. (1998). Chemical and behavioral evidence for a third pheromone
component in a North American population of the black cutworm moth, Agrotis ipsilon. J.
Chem. Ecol. 26, 329-342.
Graham, L.A. , Tang, W. , Baust, J.G. , Liou, Y.-C. , Reid, T.S. , Davies, P.L. (2001).
Characterization and cloning of a Tenebrio molitor hemolymph protein with sequence similarity
to insect odorant-binding proteins. Insect Biochem. Mol. Biol. 31, 691-702.
Gries, G. , Gries, R. , Khashin, G. , Slessor, K.N. , Grant, G.G. , Liska, J. and Kapitola, P.
(1996). Specificity of nun and gypsy moth sexual communication through multiple-component
pheromone blends. Naturwissenschaften 83, 382-385.
Hekmat-Scafe, D.S. , Steinbrecht, R.A. and Carlson, J.R. (1997). Coexpression of two odorantbinding protein homologs in Drosophila: implications for olfactory coding. J. Neurosci. 17,
1616-1624.
Hekmat-Scafe, D.S. , Dorit, R.L. and Carlson, J.R. (2000). Molecular evolution of odorantbinding protein genes OS-E and OS-F in Drosophila. Genetics 155, 117-127.
27
Horst, R. , Damberger, F. , Luginbühl, P. , Güntert, P. , Peng, G. , Nikonova, L. , Leal, W.S. and
Wüthrich, K. (2001). NMR structure reveals intramolecular regulation mechanism for
pheromone binding and release. Proc. Natl. Acad. Sci. USA 98, 14374-14379.
Jacquin-Joly, E. , Vogt, R.G. , Francois, M.C. and Nagnan-Le Meillour, P. (2001). Functional
and expression pattern analysis of chemosensory proteins expressed in the antennae and
pheromonal gland of Mamestra brassicae. Chem. Senses 26, 833-844.
Kaissling, K.E., Klein, U., De Kramer, J.J., Keil, T.A., Kanaujia, S. and Hemberger, J. (1985).
Insect olfactory cells: Electrophysiological and biochemical studies. In "Molecular Basis of
Nerve Activity" (Proc. Int. Symp. in Memory of D. Nachmansohn, Oct. 1984) (J.P. Changeux, F.
Hucho, E. Maelicke, and E. Neumann, eds. ), pp. 173-183. de Gruyter, Berlin.
Kanaujia, L. et Kaissling, K.E. (1985). Interactions of pheromone with moth antennae:
adsorption, desorption and transport. J. Insect Physiol. 31, 71-81.
Keil T.A. (1996). Sensilla on the maxillary palps of Helicoverpa armigera caterpillars: in search
of the CO2-receptor. Tissue cell 28, 703-717.
Kim, M.S. , Repp, A. and Smith, D.P. (1998). LUSH Odorant-Binding Protein mediates
chemosensory responses to alcohols in Drosophila melanogaster. Genetics 150, 711-721.
28
Kitabayashi, A.N. , Arai, T. , Kubo, T. and Natori, S. (1998). Molecular cloning of cDNA for
p10, a novel protein that increases in the regenerating legs of Periplaneta americana (American
cockroach). Insect Biochem. Mol. Biol. 28, 785-790.
Klun, J.A. , Chapman, O.L. , Mattes, K.C. , Wojtkowski, P.W. , Beroza, M. and Sonnet, P.E.
(1973). Insect sex pheromones: minor amounts of opposite geometrical isomer critical to sex
attraction. Science 162, 661-663.
Kõdrick, D. , Filippov, V.A. , Filippova, M.A. , Sehnal, F. (1995). Sericotropin: an insect
neurohormonal factor affecting RNA transcription. J. Zool. 45, 68-70.
Kowcun, A. , Honson, N. and Plettner, E. (2001). Olfaction in the gypsy moth, Lymantria
dispar: effect of pH, ionic strength, and reductants on pheromone transport by pheromonebinding proteins. J. Biol. Chem. 276, 44770-44776.
Krieger, J. , Raming, K. and Breer, H. (1991). Cloning of genomic and complementary DNA
encoding insect pheromone binding proteins: evidence for microdiversty. Biochim. Biophys.
Acta 1088, 277-284.
Krieger, J. , Ganssle, K. , Raming, K. , Breer, H. (1993). Odorant binding proteins of Heliothis
virescens. Insect Biochem. Molec. Biol. 23, 449-456.
Krieger, J. , von Nickisch-Roseneck, E.V. , Mameli, M. , Pelosi, P. et Breer, H. (1996). Binding
proteins from the antennae of Bombyx mori. Insect Biochem. Molec. Biol. 26, 297-307.
29
Krieger, J. , Mameli, M. and Breer, H. (1997). Elements of the olfactory signaling pathways in
insect antennae. Invert. Neurosci. 3, 137-144.
Koganezawa, M. and Shimada, I. (2002). Novel odorant-binding proteins expressed in the taste
tissue of the fly. Chem. Senses 27, 319-332.
Laforest, S. Prestwich, G.D. and Löfstedt, C. (1999). Intraspecific nucleotide variation at the
Pheromone Binding Protein locus in the turnip moth, Agrotis segetum. Insect Mol. Biol. 8, 481490.
Lartigue, A. , Campanacci, V. , Roussel, A. , Larsson, A.M. , Jones, T.A. , Tegoni, M. and
Cambillaud, C. (2002). X-Ray structure and ligand binding study of a moth chemosensory
protein. J. Biol. Chem. in press.
Laue, M. , Steinbrecht, R.A. and Ziegelberger, G. (1994). Immunocytochemical localization of
general odorant binding protein in olfactory sensilla of the silkworm Antheraea polyphemus.
Naturwissenschaften 81, 178-180.
Leal, W.S. , Nikonova, L. and Peng, G. (1999). Disulfide structure of the pheromone binding
protein from the silkworm moth, Bombyx mori. FEBS Lett. 464, 85-90.
30
Löfstedt, C. , Löfsqvist, J. , Lanne, B.S. , Van Der Pers, J.C.N. and Hansson, B.S. (1982). Sex
pheromone components of the turnip moth, Agrotis segetum, chemical identification,
electrophysiological evaluation and behavioral activity. J. Chem. Ecol. 8, 1305-1321.
Lonergan, G.C. (1986). Metabolism of pheromone components and analogs by cuticular
enzymes of Choristoneura fumiferana. J. Chem. Senses 12, 483-496.
Maïbèche-Coisné, M. , Sobrio, F. , Delaunay, T. , Lettere, M. , Dubroca, J. , Jacquin-Joly, E. et
Nagnan-Le Meillour, P. (1997). Pheromone binding proteins of the moth Mamestra brassicae:
specificity of ligand binding. Insect Biochem. Molec. Biol. 27, 213-221.
Maïbèche-Coisné, M. , Jacquin-Joly, E. , Francois, M.C. et Nagnan-Le Meillour, P. (1998).
Molecular cloning of two Pheromone Binding Proteins in Mamestra brassicae. Insect Biochem.
Molec. Biol. 28, 815-818.
Maida, R. , Steinbrecht, A. , Ziegelberger, G. and Pelosi, P. (1993). The pheromone binding
protein of Bombyx mori: purification, characterization and immunocytochemical localization.
Insect Biochem. Molec. Biol. 23, 243-253.
Maida, R. , Krieger, J. , Gebauer, T. , Lange, U. and Ziegelberger, G. (2000). Three pheromonebinding proteins in olfactory sensilla of the two silkmoth species Antheraea polyphemus and
Antheraea pernyi. Eur. J. Biochem. 267, 2899-2908.
31
Maleszka, R. and Stange, G. (1997). Molecular cloning by a novel approach, of a cDNA
encoding a putative olfactory protein in the labial palps of the moth Cactoblastis cactorum. Gene
202, 39-43.
Mameli, M. , Tuccini, A. , Mazza, M. , Petacchi, R. and Pelosi, P. (1996). Soluble proteins in
chemosensory organs of Phasmids. Insect Biochem. Mol. Biol. 26, 875-882.
Marchese, S., Angeli, S. , Andolfo, A. , Scaloni, A. , Brandazza, A. , Mazza, M. , Picimbon, J.F.
, Leal, W.S. and Pelosi P. (2000). Soluble proteins from chemosensory organs of Eurycantha
calcarata (Insects, Phasmatodea). Insect Biochem. Molec. Biol. . 30, 1091-1098.
McKenna, M.P. , Hekmat-Scafe, D.S. , Gaines, P. and Carlson, J.R. (1994). Putative Drosophila
pheromone-binding-proteins expressed in a subregion of the olfactory system. J. Biol. Chem.
269, 16340-16347.
Merritt, T.J.S. , Laforest, S. , Prestwich, G.D. , Quattro, J.M. and Vogt, R.G. (1998). Patterns of
gene duplication in lepidopteran pheromone binding protein. J. Mol. Evol. 46, 272-276.
Monteforti, G. , Angeli, S. , Petacchi, R. and Minnocci, A. (2002). Ultrastructural
characterization of antennal sensilla and immunocytochemical localization of a chemosensory
protein in Carausius morosus Bruner (Phasmida: Phasmatidae). Arthrop. Struct. Dev. 30, 195205.
32
Nagnan-Le Meillour, P. , Cain, A.H. , Jacquin-Joly, E. , Francois, M.C. , Ramashadran, S. ,
Maida, R. and Steinbrecht, R.A. (2000). Chemo-Sensory proteins from the proboscis of
Mamestra brassicae. Chem. Senses 25, 541-553.
Nomura, A. , Kawasaki, K. , Kubo, T. and Natori, S. (1992). Purification and localization of p10,
a novel protein that increases in nymphal regenerating legs of Periplaneta americana (American
cockroach). Int. J. Dev. Biol. 36, 391-398.
Ozaki, M. , Morizaki, K. , Idei, W. , Ozaki, K. and Tokunaga, F. (1995). A putative lipophilic
stimulant carrier protein commonly found in the taste and olfactory systems A unique member of
the pheromone binding protein superfamily. Eur. J. Biochem. 230, 298-308.
Paesen, G.C. and Happ, G.M. (1995). The B proteins secreted by the tubular accessory sex
glands of the male mealworm beetle, Tenebrio molito, have sequence similarity to moth
pheromone-binding proteins. Insect Biochem. Mol. Biol. 25, 401-408.
Park, S.-K. , Shanbag, S. , Wang, Q. , Yu, P. , Hasan, G. , Steinbrecht, A. and Pikielny, C.W.
(2002). Inactivation of olfactory sensilla of a single morphological type differentially affects the
response of Drosophila to odors. J. Neurobiol. in press.
Picimbon, J.F. (2001). Les protéines liant les odeurs (OBPs) et les protéines chimiosensorielles
(CSPs): cibles moléculaires de la lutte intégrée. In: “Biopesticides d'Origine Végétale”
(Philogène B. , Regnault-Roger, C. and Vincent, C. , eds. ), pp. 265-284, Lavoisier Tech and
Doc, Paris.
33
Picimbon, J.F. (2002). Les périrécepteurs chimiosensoriels des insectes. Med. Sci. in press.
Picimbon, J.F. , and Gadenne, C. (2002). Evolution and noctuid pheromone binding proteins:
identification of PBP in the black cutworm moth, Agrotis ipsilon. Insect Biochem. Molec. Biol.
32, 839-846.
Picimbon, J.F. , and Leal, W.S. (1999). Olfactory soluble proteins of cockroaches. Insect
Biochem. Molec. Biol. 29, 973-978.
Picimbon, J.F. , Gadenne, C. , Becard, J.M. , Clement, J.L. and Sreng, L. (1997). Sex pheromone
of the French black cutworm moth Agrotis ipsilon (Lepidoptera: Noctuidae): identifcation and
regulation of a multicomponent blend. J. Chem. Ecol. 23, 211-230.
Picimbon, J.F. , Dietrich, K. , Breer, H. and Krieger, J. (2000a). Chemosensory proteins of
Locusta migratoria (Orthoptera, Acriididae). Insect Biochem. Molec. Biol. 30, 233-241.
Picimbon, J.F. , Dietrich, K. , Angeli, S. , Scaloni, A. , Krieger, J. , Breer, H. and Pelosi, P.
(2000b). Purification and molecular cloning of chemosensory proteins from Bombyx mori. Arch.
Insect Biochem. Physiol. 44, 120-129.
Picimbon, J.F. , Dietrich, K. , Krieger, J. and Breer, H. (2001). Identity and expression pattern of
chemosensory proteins in Heltiothis virescens. Insect Biochem. Mol. Biol. 31, 1173-1181.
34
Picone, D. , Crescenzi, O. , Angeli, S. , Marchese, S. , Brandazza, A. , Pelosi, P. and Scaloni, A.
(2001). Bacterial expression and conformational analysis of a chemosensory protein from
Schistocerca gregaria. Eur. J. Biochem. 268, 4794-4801.
Pikielny, C.W. , Hasan, G. , Rouyer, F. and Rosbach, M. (1994). Members of a family of
Drosophila putative odorant-binding proteins are expressed in different subsets of olfactory
hairs. Neuron 12, 35-49.
Plettner, E. , Lazar, J. , Prestwich, E. and Prestwich, G.D. (2000). Discrimination of pheromone
enantiomers by two pheromone binding proteins from the gypsy moth, Lymantria dispar.
Biochemistry 30, 8953-8962.
Prestwich, G.D. , Graham, S. , Handley, M. , Latli, B. , Streinz, L. and Tasayco, M.J. (1989).
Enzymatic processing of pheromones and pheromone analogs. Experientia 45, 263-270.
Prestwich, G.D. , Du, G. and Laforest, S. (1995). How is pheromone specificity encoded in
proteins. Chem. Senses 20, 461-469.
Raming, K. , Krieger, J. and Breer, H. (1989). Molecular cloning of an insect pheromonebinding-protein. FEBS Lett. 256, 215-218.
Raming, K. , Krieger, J. and Breer, H. (1990). Primary structure of a pheromone-binding protein
from Antheraea pernyi: homologies with other ligand-cerrying proteins. J. Comp. Physiol. B
160, 503-509.
35
Robertson, H.M. , Martos, R. , Sears, C.R. , Todres, E.Z. , Walden, K.K.O. and Nardi, J.B.
(1999). Diversity of odourant binding proteins revealed by an expressed sequence tag project on
male Manduca sexta moth antennae. Insect Mol. Biol. 8, 501-518.
Rothemund, S. , Liou, Y.-C. , Davies, P.L. , Krause, E. , Sonnischen, F.D. (1999). A new class of
hexahelical insect proteins revealed as putative carriers of small hydrophobic ligands. Structure
7, 143-151.
Sandler, B.H. , Nikonova, L. , Leal, W.S. and Clardy J. (2000). Sexual attraction in the silkworm
moth: structure of the pheromone-binding-protein-bombykol complex. Chem. Biol. 7, 143-151.
Schneider, D. (1969). Insect olfaction: deciphering system for chemical messages. Science 163,
1031-1036.
Segal, S.P. , Graves, L.E. , Verheyden, J. and Goodwin, E.B. (2001). RNA-regulated TRA-1
nuclear export controls sexual fate. Dev. Cell 1, 539-551.
Saitou, N. and Nei, M. (1987). The neighbor joining method: a new method for reconstructing
phylogenetic trees. Mol. Biol. Evol. 8, 501-518.
Scaloni, A. , Monti, M. , Angeli, S. and Pelosi, P. (1999). Structural analysis and disulfide bridge
pairing of two odorant binding proteins from Bombyx mori. Biochem. Biophys. Res. Comm. 266,
386-391.
36
Shanbhag, S.R. , Hekmat-Scafe, D. , Kim, M.-S. , Park, S.-K. , Carlson, J.R. , Pikielny, C. ,
Smith, D.P. and Steinbrecht, R.A. (2001). Expression mosaic of odorant-binding proteins in
Drosophila olfactory organs. Microsc. Res. Tech. 55, 297-306.
Shanbhag, S.R. , Park, S.-K. , Pikielny, C. and Steinbrecht, R.A. (2001). Gustatory organs of
Drosophila melanogaster : fine structure and expression of the putative odorant-binding protein
PBPRP2. Cell Tissue Res. 304, 423-437.
Stange, G. (1992). High resolution measurements of atmospheric carbon dioxide changes by the
labial palp organ of the moth Heliothis armigera (Lepidoptera: Noctuidae). J. Comp. Physiol. A
171, 317-324.
Stange, G. (1996). Sensory and behavioural responses of terrestrial invertebrates to biogenic
carbon dioxide gradients. Dans: Advances in Bioclimatology Vol. 4 (ed. G.E. Stanhill), Springer
Verlag, Berlin. pp. 223-253.
Steinbrecht, R.A. (1996). Are Odorant-binding proteins in odorant discimination ? Chem. Senses
21, 719-727.
Steinbrecht, R.A. , Ozaki, M. and Ziegelberger, G. (1992). Immunocytochemical localization of
pheromone-binding-protein in moth antennae. Cell Tissue Res. 270, 287-302.
37
Steinbrecht, R.A. , Laue, M. and Ziegelberger, G. (1995). Immunolocalization of pheromonebinding protein and general odorant-binding protein in olfactory sensilla of the silk moths
Antheraea and Bombyx. Cell Tiss. Res. 282, 203-217.
Swofford, D.L. (1999). PAUP*. Phylogenetic Analysis Using Parsimony (*and other methods).
Sinauer Associates, Sunderland, MA.
Teal, P.E.A. , Tumlinson, J.H. and Heath, R.R. (1986). Chemical and behavioral analyses of
volatile sex pheromone components released by calling Heliothis virescens (F.) females
(Lepidoptera: Noctuidae). J. Chem. Ecol. 12, 1071-26.
Thompson, J.D. , Gibson, T.J. , Plevniak, F. , Jeanmougin, F. , Higgins, D.G. (1997). The Clustal
X windows interface: flexible strategies for multiple sequence alignment aided by quality
analysis tools. Nucleic Acid Res. 24, 4876-4882.
Toth, M. , Lofstedt, C. , Blair, B.W. , Cabello, T. , Farag, A.I. , Hansson, B.S. , Kovalev, B.G. ,
Maini, S. , Nesterov, E.A. , Pajor, I. , Sazonov, A.P. , Shamsev, I.V. , Subchev, M. and Szocs, G.
(1992). Attraction of male turnip moths Agrotis segetum (Lepidoptera: Noctuidae) to sex
pheromone components and their mixtures at 11 sites in Europe, Asia and Africa. J. Chem. Ecol.
18, 1337-1347.
Tuccini, A. , Maida, R. , Rovero, P. , Mazza, M. et Pelosi, P. (1996). Putative odorant-binding
protein in antennae and legs of Carausius morosus (Insecta, Phasmatodea). Insect Biochem.
Molec. Biol. 26, 19-24.
38
Tumlinson, J.H. , Brennan, M.M. , Doolittle, R.E. , Mitchell, E.R., Brabham, A. , Mazomemos,
B.E. , Baumhover, A.H. and jackson, D.M. (1989). Identification of a pheromone blend
attractive to Manduca sexta (L. ) males in a wind tunnel. Arch. Insect. Biochem. Physiol. 10,
255-271.
Thymianou, S. , Mavroidis, M. , Kokolakis, G. , Komitopoulos, K. , Zacharopoulou, A. ,
Mintzas, A.C. (1998). Cloning and characterization of a cDNA encoding a male-specific serum
protein of the Mediterranean fruit fly, Ceratitis capitata, with sequence similarity to odourant
binding proteins. Insect Mol. Biol. 7, 345-353.
Vogt, R.G. (1987). The molecular basis of pheromone reception: its influence on behavior. In:
Pheromone biochemistry, (Prestwich G.D. and Blomquist G.L. , eds. ), pp. 385-431, Acad. Press,
Orlando.
Vogt, R.G and Riddiford, L.M. (1981). Pheromone binding and inactivation by moth antennae.
Nature 293, 161-163.
Vogt, R.G. and Riddiford, L.M. (1986). Pheromone reception: a kinetic equilibrium. In:
“Seminar on mechanisms in perception and orientation to insect olfactory signals” (Payre T.L. ,
Birch M.C. and Kennedy C.E.J. , eds. ), pp. 201-208, Clarendon Press, Oxford.
39
Vogt, R.G. , Riddiford, L.M. and Prestwich, G.D. (1985). Kinetic properties of a pheromone
degrading enzyme: the sensillar esterase of Antheraea polyphemus. Proc. Natl. Acad. Sci. USA
82: 8827-8831.
Vogt, R.G. , Köhne, A.C. , Dubnau, J.T. and Prestwich G.D. (1989). Expression of Pheromone
Binding Proteins during antennal development in the gypsy moth Lymantria dispar. J. Neurosci.
9, 332-3346.
Vogt, R.G. , Prestwich, G.D. and Lerner, M.R. (1991a). Odorant binding protein subfamilies
associate with distinct classes of olfactory receptor neurons in insects. J. Neurobiol. 22, 74-84.
Vogt, R.G. , Rybczynski, R. and Lerner, M.R. (1991b). Molecular cloning and sequencing of
general odorant binding proteins GOBP1 and GOBP2 from the tobacco hawk moth Manduca
sexta: comparison with other insect OBPs and their signal peptides. J. Neurosci. 11, 2972-2984.
Vogt, R.G. , Rybczynski, R. , Cruz, M. and Lerner, M.R. (1993). Ecdysteroid regulation of
olfactory protein in the developing antenna of the tobacco hawk moth, Manduca sexta. J.
Neurobiol. 24, 581-597.
Vogt, R.G. , Callahan, F.E. , Rogers, M.E. and Dickens J.C. (1999). Odorant Binding Proteins
diversity and distribution among the insect orders, as indicated by LAP, an OBP-related protein
of the true bug Lygus lineolaris (Hemiptera, Heteroptera). Chem. Senses 24, 481-495.
40
Vogt, R.G. , Rogers, M.E. , Franco, M.D. and Sun, M. (2002). A comparative study of odorant
binding protein genes: differential expression of the PBP1-GOBP2 gene cluster in M. sexta
(Lepidoptera) and the organization of OBP genes in Drosophila melanogaster (Diptera). J. Exp.
Biol. 205, 719-744.
Willett, C.S. (2000). Do pheromone binding proteins converge in amino acid sequence when
pheromones converge ? J. Mol. Evol. 50, 175-183.
Willett, C.S. and Harrison, R.G. (1999). Pheromone binding proteins in the European and Asian
corn borers: No protein change associated with pheromone differences. Insect Biochem. Mol.
Biol. 29, 277-284.
Wojtasek, H. , and Leal W.S. (1999). Conformational change in the pheromone-binding protein
from Bombyx mori induced by pH and by interaction with membranes. J. Biol. Chem. 274,
30950-30956.
Wojtasek, H. , Hansson, B.H. , Leal W.S. (1998).Attracted or repelled? A matter of two neurons,
one pheromone binding protein and a chiral center. Biochem. Biophys. Res. Comm. 250, 217222.
Wojtasek, H. , Picimbon, J.F. , Leal W.S. (1999). Identification and cloning of odorant binding
proteins from the scarab beetle Phyllopertha diversa. Biochem. Biophys. Res. Comm. 263, 832837.
41
Wu, W.Q. , Cottrell, C.B. , Hansson, B.S. and Lofstedt, C. (1999). Comparative study of
pheromone production and response in Swedish and Zimbabwean populations of turnip moth,
Agrotis segetum. J. Chem. Ecol. 25, 177-196.
Xu, A. , Park, S.-K. , Domello, D. , Kim, E. , Wang, Q. , and Pikielny, C. (2002). Novel genes
expressed in subsets of chemosensory sensilla on the front legs of male Drososphila. Cell Tiss.
Res. in press.
42
Table 1: The PBP related family of Odorant Binding Proteins.
Protein
Insect
PBP
Aips-1
Aips-2
Aseg-2
Aseg-1
Aper-1
Aper-2
Apol-1
Bmor-1
Hvir-1
Hvir-2
Hzea-1
Ldis-2
Ldis-1
Mbra-2
Mbra-1
Msex-1
ABPX
AipsABPX-1
AperABPX
BmorABPX
HvirABPX-1
MsexABPX
DmelPBPRP
DmelPBPRP-1
Lepidoptera
Agrotis ipsilon
A. ipsilon
Agrotis segetum
A. segetum
Antheraea pernyi
A. pernyi
Antheraea polyphemus
Bombyx mori
Heliothis virescens
H. virescens
Heliothis zea
Lymantria dispar
L. dispar
Mamestra brassicae
M. brassicae
Manduca sexta
A. ipsilon
A. pernyi
B. mori
H. virescens
M. sexta
Diptera
Drosophila melanogaster
GenBANK Access Number
References
AF090191
AF007868
AF007867
AF051143
AF05051142
AF323972
Picimbon and Gadenne, 2002
Picimbon and Gadenne, 2002
Abraham et al., 2002
Prestwich et al., 1995; Laforest et al., 1999
Raming et al., 1990
Krieger et al., 1991
Raming et al., 1989
Krieger et al., 1996
Krieger et al., 1993
Abraham et al., 2002
Callahan et al., 2000
Prestwich et al., 1995
Merritt et al., 1998
Maïbèche-Coisne et al., 1998
Maïbèche-Coisne et al., 1998
Gyorgyi et al., 1988; Robertson et al., 1999
CAA05509
CAA64446
CAA05508
AF117577_1/AF117575_1
Picimbon et al., unpublished
Krieger et al., 1997
Krieger et al., 1996
Krieger et al., 1997
Robertson et al., 1999
NP_524039/P54191/AAC46474
Pikielny et al., 1994
AF134253-AF134294
X96773
X96860
X17559
X94987
X96861
Table 2: The OS-D-related family of ChemoSensory Proteins.
Protein
Insect
GenBANK Access Number
References
agCG50175
agCG50200
agCG50208
agCG50210
agCG50220
agCP11484
AgSAP1
AipsCSP
BmorCSP1
BmorCSP2
CLP1
DmelOS-D/A10
CG30172
CG9358
PEBmeIII
RH74005/CG11390
CSPec1
CSPec2
CSPec3
CSPHarm
HvirCSP1
HvirCSP2
HvirCSP3
LmigOS-D1
LmigOS-D2
LmigOS-D3
Anopheles gambiae
A. gambiae
A. gambiae
A. gambiae
A. gambiae
A. gambiae
A. gambiae
Agrotis ipsilon
Bombyx mori
B. mori
Cactoblastis cactorum
Drosophila melanogaster
D. melanogaster
D. melanogaster
D. melanogaster
D. melanogaster
Eurycantha calcarata
E. calcarata
E. calcarata
Heliothis armigera
Heliothis virescens
H. virescens
H. virescens
Locusta migratoria
L. migratoria
L. migratoria
EAA12703
EAA12591
EAA12601
EAA12338
EAA12353
EAA12322
AAL84186
The Anopheles Genome Sequencing Consortium
TAGSC
TAGSC
TAGSC
TAGSC
TAGSC
Biessmann et al. , 2002
Picimbon, unpublished
Picimbon et al. , 2000b
Picimbon et al. , 2000b
Maleszka and Stange, 1997
McKenna et al. , 1994; Pikielny et al. , 1994
Adams et al. , 2000
Adams et al. , 2000
Dyanov and Dzitoeva, 1995
Stapelton et al. , unpublished; Adams et al. , 2000
Marchese et al. , 2000
Marchese et al. , 2000
Marchese et al. , 2000
Deyts et al. , unpublished
Picimbon et al. , 2001
Picimbon et al. , 2001
Picimbon et al. , 2001
Picimbon et al. , 2000a
Picimbon et al. , 2000a
Picimbon et al. , 2000a
AAM34276
AAM34275
AAC47827
AAA21358/NP524121
AAM68292
AAF47307
AAA87058
AAM29645/AAF47140
AAD30550
AAD30551
AAD30552
AAK53762
AAM77041
AAM77040
AAM77042
CAB65177
CAB65178
CAB65179
LmigOS-D4
LmigOS-D5
SAP1
SAP2
SAP3
SAP4
SAP5
CSPMbraA1
CSPMbraA2
CSPMbraA4
CSPMbraA5
CSPMbraA6/A3
CSPMbraB1
CSPMbraB2
CSPMbraB3
CSPMbraB4
p10
CSPsg1
CSPsg2
CSPsg3
CSPsg4
CSPsg5
SgreOS-D1
L.migratoria
L.migratoria
Manduca sexta
M. sexta
M. sexta
M. sexta
M. sexta
Mamestra brassicae
M. brassicae
M. brassicae
M. brassicae
M. brassicae
M. brassicae
M. brassicae
M. brassicae
M. brassicae
Periplaneta americana
Schistocerca gregaria
S. gregaria
S. gregaria
S. gregaria
S. gregaria
S. gregaria
CAB65180
CAB65181
AAF16696
AAF16714
AAF16707
AAF16721
AAF16716
AAF19647
AAF19648
AAF19650
AAF19651
AAF71289, AAF19649
AAF19652
AAF19653
AAF71290
AAF71291
AAB84283/AAB24286
AAC25399
AAC25400
AAC25401
AAC25402
AAC25403
Picimbon et al. , 2000a
Picimbon et al. , 2000a
Robertson et al. , 1999
Robertson et al. , 1999
Robertson et al. , 1999
Robertson et al. , 1999
Robertson et al. , 1999
Nagnan-Le Meillour et al. , 2000
Nagnan-Le Meillour et al. , 2000
Nagnan-Le Meillour et al. , 2000
Nagnan-Le Meillour et al. , 2000
Nagnan-Le Meillour et al. , 2000; Jacquin-Joly et al. , 2001
Jacquin-Joly et al. , 2001
Jacquin-Joly et al. , 2001
Jacquin-Joly et al. , 2001
Jacquin-Joly et al. , 2001
Nomura et al. , 1992; Kitabayashi et al. , 1998
Angeli et al. , 1999
Angeli et al. , 1999
Angeli et al. , 1999
Angeli et al. , 1999
Angeli et al. , 1999
Picimbon, unpublished
FIGURE CAPTIONS
Figure 1: Neighbor joining tree selected members of the PBP protein class, based on 1000
bootstrap replicates (Clustal X 1.8; Saitou and Nei, 1987; Thompson et al. , 1997). Relative
branch lengths are indicated by the scale bar. Two groups of PBPs within Noctuidae are clearly
revealed. Group 1 (Grp1): Aips-1/Aseg-1/Mbra-2/Hvir-1/Hzea-1; Group 2 (Grp2): Aips-2/Aseg2/Mbra-1/Hvir-2.
Figure 2: Exon and intron boundaries of the genes encoding the PBPs from A. ipsilon and A.
segetum. The numbers indicate the length of exons and introns (base pairs). The two groups of
PBPs (Grp1 and Grp2) correspond to different gene structures. Aseg-1/Aseg-2/Aips-1/Aips-2
(Laforest et al. , 1999; Abraham et al. , 2002; Picimbon and Gadenne, 2002).
Figure 3: Alignment of moth ABPX and DmelPBPRPs proteins. Sources of the sequences and
accession numbers are reported in table 1. The amino acids conserved in ABPXs and
DmelPBPRP-1 are represented in bold. Those amino acids found in DmelPBPRP-1 and ABPX
or AipsABPX and DmelPBPRP-1, are represented in italics. Conserved cysteines are
underlined. These specific amino acids support the classification of these sequences as OBP1
type of proteins.
43
Figure 4: A. Phylogenetic relationships of moth ABPXs and DmelPBPRPs. The primary
sequences of the proteins were aligned in Clustal X 1.8 and processed using PAUP 4.0d65. The
tree represents equally most parsimonious trees of 909 steps and consistency index 0.52. The
numbers above each branch indicates the percent bootstrap support above 50% for the supported
node using maximum parsimony (Felsenstein, 1985). The ABPX related protein PdivOBP1 from
the scarab beetle Phylloperta diversa (Acc. Num. BAA88061; Wojtasek et al. , 1999) was used
as outgroup to root the tree. Using the well-defined PBP/GOBP clade as outgroup generated the
same tree topology. B. Phylogenetic analysis of the CSP protein family. An alignment of the
sequences of different CSPs reported in table 2 was used to determine a distance matrix and
generate an unrooted tree (Clustal X 1.8; PAUP 4.0d65). This strict consensus tree is the result
of trees derived using maximum parsimony of 10,000 steps and consistency index 0.52. The
numbers above each branch indicates the percentage of bootstrap above 50% that support the
branching pattern. The proteins agCG50208 and CG30172 were declared as the outgroup to
build the phylogenetic tree of CSPs. Specific grouping of CSPs is observed, in particular in the
CSPs from moths.
Group 1 (CSP1):AipsCSP/SAP2/BmorCSP1/CLP1/HvirCSP3/CSPHarm/CSPMbraA; Group 2
(CSP2): HvirCSP1/SAP3/CSPMbraB; (CSP3): HvirCSP2/SAP5; SAP1; BmorCSP2; SAP5.
Figure 5: Structural properties of moth CSPs. The amino acid residues conserved in these
three types of CSP are shown in bold. Black triangles indicate residues in contact with water
molecules in the channel structure of CSP. The position of the six α-helices characteristic of
CSP are indicated (α1 to α6).
44
45
Picimbon Figure 1
Figure 2, Picimbon
Aseg-1
Aips-1
Aseg-2
Aips-2
Gene:
Exon1 - Intron1 - Exon2 - Intron2 - Exon3
Aseg-1: 66
318
- 180 993
- 183
Aips-1: 66
312
- 180 - 1056
- 183
Aseg-2: 66
362
- 180 513
- 180
458
- 180 564
- 180
Aips-2: 66
Picimbon Figure 3
AipsABPX-1
HvirABPX-1
MsexABPX
AperABPX
BmorABPX
DmelPBPRP-1
OBP-1
1
10
20
30
40
50
60
GVVMDEDMAELARMVRESCVDETGADVKLVEAANGGADLME--DDKLKCYIKCTMETAGMAVAMDEDMAELARMVRENCAAETGADVALVERVNAGADLMP--DDKLKCYIKCTMETAGMLALEDEEQAELARMVRENCVHEIGVDEGLLAKVDDGADLMP--DPKLKCYLKCTMEMAGMVASLDGEMAELAKMIRDNCADEIGVDVTLLEQVDAGANLMP--DEKFKCYLKCTLETAGMHGQLDDEIAELAAMVRENCADESSVDLNLVEKVNAGTDLATITDGKLKCYIKCTMETAGMVEINPTIIKQV-RKLRMRCLNQTGASVDVIDKSVKNRILPT--DPEIKCFLYCMFDMFGLI
V-----------R--R--C---TGA-V-----------L-T--D---KC-L-C-----G--
70
80
90
100
110
MSDGEVDIEAVMALLPPEMAEHNGPALKSCGTQRGADDCDTAWKTQVCWQNANKAEYFLI
MADGEVDIEAVLALLPPELAEHNAPSLRACGTVRGADHCDTAFRTQQCWQNANKADYFLI
ISDGVVDVEAVLGLLPDDVKLRTTDIVRACDTQKGADDCDTAFLTQTCWQQANRADYIFI
MSDGVVDIEIVLELLPEDLKTKNENLLRKCDTQKGSDDCDTAFLTQVCWQNGNKADYFLI
MSDGVVDVEAVLSLLPDSLKTKNEASLKKCDTQKGSDDCDTAYLTQICWQAANKADYFLI
DSQNIMHLEALLEVLPEEIYKTINGLVSSCGTQKGKDGCDTAYETVKCYIAVNGKFFIWEEIIVLLG
-S------EA-LE-LPEE-------LV-SCGTQ-G-D-CDTA--T--C--A-N-------
-118
-118
-118
-118
-120
-123
Picimbon Figure 4
Picimbon Figure 5
1
CSP1
CSP2
CSP3
10
20
40
50
-D-YTDKYD-----EIL-N-RLL--Y--CV---GKC--EGKELK--L--A----YTD-YD-V-LDEIL-N-R--VPY-KCILD-GKCAPD-KELKEHI-EALE
ASTYTDKWDNINVDEILES-RL-K-YVDCLLD-GRCTPDGKALKETLPDALE
YTD D
EIL
R
Y C
GKC
K LK
A
α2
α1
60
CSP1
CSP2
CSP3
30
70
α3
80
90
100
110
-GC-KC---QE-G-----I--LIKN----W--L----DP---WR-KYEDRA-A-GI-IP--ECGKCT--QK-GGTRRVI-HLINHE---W-EL--K-DP---KYEKEL---K--CSKCTEKQKAG~S-KVIR-LVNKR--LWKELSAKYDPNN-YQ--YKDKI---KGQ
C KC
Q G
I L
W L
DP
α4
α5
α6