You are on page 1of 15

Journal of Insect Physiology 95 (2016) 51–65

Contents lists available at ScienceDirect

Journal of Insect Physiology


journal homepage: www.elsevier.com/locate/jinsphys

Review

A look inside odorant-binding proteins in insect chemoreception


Nathália F. Brito a, Monica F. Moreira a,b, Ana C.A. Melo a,b,⇑
a
Universidade Federal do Rio de Janeiro, Instituto de Química, 21941-909 Rio de Janeiro, RJ, Brazil
b
Instituto Nacional de Ciência e Tecnologia em Entomologia Molecular, Rio de Janeiro, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Detection of chemical signals from the environment through olfaction is an indispensable mechanism for
Received 13 May 2016 maintaining an insect’s life, evoking critical behavioral responses. Among several proteins involved in the
Received in revised form 13 September 2016 olfactory perception process, the odorant binding protein (OBP) has been shown to be essential for a nor-
Accepted 14 September 2016
mally functioning olfactory system. This paper discusses the role of OBPs in insect chemoreception. Here,
Available online 14 September 2016
structural aspects, mechanisms of action and binding affinity of such proteins are reviewed, as well as
their promising application as molecular targets for the development of new strategies for insect popu-
Keywords:
lation management and other technological purposes.
Odorant-binding protein
Insect
Ó 2016 Elsevier Ltd. All rights reserved.
Review
Olfactory system
Chemical communication

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2. Molecular organization of insect peripheral olfactory system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3. Detection of semiochemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4. Expression in non-sensory tissues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5. Physicochemical properties and structural aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6. Three-dimensional structure-based mechanism of action. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7. Activation of ORs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
8. Binding affinity and selectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
9. Applied research. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
10. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Funding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

1. Introduction

Interpretation of chemical signals from the environment is an


essential mechanism for maintaining insect life. It is well
Abbreviations: OBP, odorant-binding protein; OR, odorant-receptor; PBP, established that insects use a variety of semiochemicals – such
pheromone-binding protein; PR, pheromone receptor; GOBP, general odorant- as pheromones, plant volatiles and animal odorants – to find one
binding protein; SNMP, sensory neuron membrane protein; IR, ionotropic receptor;
CSP, chemosensory protein; ODE, odorant-degrading enzyme; GPCR, G protein-
another, besides locating food sources, oviposition sites, mating
coupled receptor; ORCo, odorant receptor co-receptor; cVA, cis-vaccenyl acetate; partners and suitable hosts, in addition to identifying predators,
DEET, diethyltoluamide; AMA, 1-aminoanthracene; 1-NPN, N-phenyl-1- among other functions (Zwiebel and Takken, 2004). Semiochemi-
naphthylamine; ANS, 1-anilino-8-naphthalenesulfonic acid. cals are perceived through olfaction, the sensory modality which
⇑ Corresponding author at: Universidade Federal do Rio de Janeiro, Instituto de
is responsible for the transduction of the olfactory signal, evoking
Química, 21941-909, Rio de Janeiro, RJ, Brazil.
E-mail address: anamelo@iq.ufrj.br (A.C.A. Melo).
behavioral responses that are critical for survival and reproduction

http://dx.doi.org/10.1016/j.jinsphys.2016.09.008
0022-1910/Ó 2016 Elsevier Ltd. All rights reserved.
52 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

of insects. In order to initiate signal transduction, it is necessary 2. Molecular organization of insect peripheral olfactory system
that these hydrophobic molecules get across the hydrophilic sen-
sillum lymph surrounding sensory neurons and stimulate specific Upon encountering the antennae, odorants are able to penetrate
dendrites (Fan et al., 2011), where olfactory receptors (ORs) are the cuticle wall through multiple pores present at the surface of
located on the membrane surface (Breer, 2003). Inside olfactory sensilla and reach the aqueous environment of the sensillar lymph,
sensilla (delicate hair-like porous cuticular structures located on rich in soluble proteins. Odorant-binding proteins (OBPs), odorant
antennae), odorant-binding proteins are synthesized by special- receptors (ORs), ionotropic receptors (IRs), sensory neuron
ized accessory cells and secreted in the sensillar lymph, where they membrane proteins (SNMPs), chemosensory proteins (CSPs) and
are found in high concentrations (Klein, 1987; Vogt et al., 1985). odorant-degrading enzymes (ODEs) are the main proteins of the
Although olfaction is not yet completely understood at molecular peripheral olfactory system involved in the odorant reception pro-
level, many studies indicate these proteins participate in the chem- cess (Leal, 2012).
ical communication process and constitute the first line of olfac- OBPs are responsible for the connection between the external
tory proteins involved in odorant recognition. Several functions environment and ORs (Leal, 2005). Once odorants penetrate the
have been proposed for odorant-binding proteins, among which pore tubules of sensilla, they are bound and solubilized by OBPs,
protection of odorant molecules from the action of odorant- transported through the sensillar lymph and finally reach sensory
degrading enzymes and transport of hydrophobic semiochemicals dendrites, where they activate membrane-bound ORs (Fig. 1).
through the sensillum lymph to the odorant receptors are the most The number of genes encoding putative OBPs is very variable
studied. Odorant-binding proteins are also found in vertebrates; across different insect orders, with some species having up to sev-
these, however, belong to the lipocalin family and show no struc- eral tens of them (Gong et al., 2009a; Hekmat-Scafe et al., 2002;
tural similarity to the insect OBPs (Bianchet et al., 1996). Manoharan et al., 2013). The OBP-family also includes proteins that
Even though information regarding insect odorant-binding specifically bind and transport pheromones to pheromone recep-
proteins have been summarized in excellent reviews over the last tors (PRs), called pheromone-binding proteins (PBPs). Regarding
years (Fan et al., 2011; Leal, 2013; Pelosi et al., 2006, 2014; Venthur OBPs mode of action, two not necessarily mutually exclusive
et al., 2014; Vieira and Rozas, 2011; Zhou, 2010), knowledge about hypothesis stand out: although there is evidence in moths and,
this subject has been constantly increasing. Therefore, this review mainly, in mosquitoes, that OBPs act as passive carriers and the
aims to provide the reader an up-to-date integrated scenario on ligand per se activates its OR (Damberger et al., 2007; Horst et al.,
what has been described for insect OBPs so far. Also, this paper 2001; Sandler et al., 2000; Wojtasek and Leal, 1999), in some cases
dedicates an entire section to discussing current ideas and perspec- OBPs seem to play a more direct role, where the formation of a
tives for applied research involving insect OBPs, an important topic specific OBP-ligand complex is necessary for receptor activation
which seems not to be taken into account in the majority of (Laughlin et al., 2008; Ronderos and Smith, 2010; Xu et al., 2005).
published reviews. CSPs belong to another family of small soluble proteins identi-
fied across multiple insect orders. They are smaller than OBPs

Fig. 1. Schematic view of the odorant perception process in insects. Once odorants penetrate pore tubules of the sensillum, they are bound and solubilized by OBPs,
transported through the sensillum lymph and finally reach sensory dendrites, where they activate membrane-bound ORs. Signal transduction evokes a response behavior to
the detected stimulus and stray odorants are rapidly degraded. For further information, see Sanchez-Gracia et al. (2009). Details regarding different mechanisms of ligand
binding and release by OBPs are not represented in this scheme and are discussed in another section.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 53

and can bind a wide range of chemicals. The immunocytochemical et al., 1998). It had already been demonstrated three years before
localization of these proteins in the hemolymph of olfactory sen- that, in Phormia regina, the electrophysiological answer of gusta-
silla suggests they could be involved in insect chemical signaling. tive cells was completely abolished in the presence of anti-OBP
In ants, CSPs are highly expressed in antennae, where they might antibody (Ozaki et al., 1995).
play a role in nest mate recognition signals (Gonzalez et al., Although OBPs binding affinity for different ligands has been
2009; Ozaki et al., 2008). However, CSPs have a larger tissue distri- determined for several species since 1981, evidence that these pro-
bution when compared with OBPs and several members of this teins are essential for a normally functioning olfactory system was
group are expressed in non-sensory organs where they seem to only obtained much more recently. Knockdown studies have
be involved in different functions such as development and differ- demonstrated that OBP DmelOBP76a (or LUSH) is necessary in
entiation (Jacquin-Joly et al., 2001; Maleszka et al., 2007; Nomura the olfactory process of D. melanogaster (Laughlin et al., 2008; Xu
et al., 1992; Zhou et al., 2013). In cockroaches, for instance, p10, a et al., 2005). In experiments where pheromones were solubilized
protein of this class, was reported to promote the regeneration of with PBPs, threshold responses from corresponding receptors were
legs (Kitabayashi et al., 1998; Nomura et al., 1992). two to three orders of magnitude lower than when pheromones
In contrast to soluble OBPs and CSPs, SNMPs genes encode a were solubilized with DMSO (Forstner et al., 2009; Grosse-Wilde
transmembrane protein which represents a subclade of the CD36 et al., 2006, 2007). Similar results were obtained in vivo by express-
gene family in insects (Nichols and Vogt, 2008) and several studies ing the bombykol receptor from Bombyx mori, BmorOR1, in the
support a role for SNMPs in pheromone-based chemical communi- ‘‘empty neuron” of D. melanogaster: sensitivity to bombykol was
cation, where an PBP passes the pheromone to the SNMP, which significantly enhanced when BmorOR1 was co-expressed with a
then delivers it to the receptor. SNMP1 from Drosophila melanoga- silkworm PBP, BmorPBP1 (Syed et al., 2010). Furthermore, behav-
ster was demonstrated to be essential for detecting the sex phero- ioral assays with Drosophila mutants (Matsuo et al., 2007;
mone Z11-18OAc (also known as cis-vaccenyl acetate or simply Swarup et al., 2011) and aphids (Qiao et al., 2009; Sun et al.,
cVA) in T1 sensilla (Benton et al., 2007; Jin et al., 2008), while, in 2012a) have also indicated OBPs are, indeed, engaged in the semio-
moths, SNMP1s are associated with sex pheromone-detecting neu- chemical detection process. Results of a study involving different
rons in antennae (Forstner et al., 2008; Rogers et al., 2001; Thode combinations of PBPs and pheromone receptors (PRs) from Chilo
et al., 2008). suppressalis implicate PBPs are crucial for the olfactory response
The OR consists of a seven transmembrane domain protein and that PRs sensitivity to pheromones is enhanced by up to four
which is not homologous to vertebrate GPCR receptors. Insect orders of magnitude in the presence of PBPs. The authors suggest
ORs adopt an inverse membrane topology when compared to clas- the correct combination between OBP and OR is fundamental for
sical GPCRs, with the N-terminal region located intracellularly the detection of a specific odorant (Chang et al., 2015).
(Benton et al., 2006; Buck and Axel, 1991). ORs are usually part Functions assigned to OBPs have been described by biochemi-
of a heteromeric complex comprising at least two subunits: a cal, biophysical, structural biology and kinetic studies (Horst
ligand-specific OR and a highly conserved co-receptor (ORCo) et al., 2001; Leal et al., 2005a,b; Mao et al., 2010; Sandler et al.,
(Vosshall and Hansson, 2011). Molecules specifically bind the OR, 2000; Wogulis et al., 2006; Wojtasek and Leal, 1999; Xu and
so the ORCo is not directly involved in odorant recognition Leal, 2008), as well as using interference RNA and electrophysiol-
(Guidobaldi et al., 2014; Larsson et al., 2004). Heteromeric ORCo ogy approaches (Biessmann et al., 2010; Pelletier et al., 2010). It
complexes form ion channels induced by odorant binding to the is currently accepted that OBPs assist the passage of odorants
OR (Sato et al., 2008; Wicher et al., 2009), which leads to signal through the aqueous environment of the sensillum lymph to the
transduction. ORs, aside from contributing to the sensitivity of the olfactory sys-
IRs are related to ionotropic glutamate receptors, even though tem (Leal, 2012).
both families are divergent given the fact that important residues
for glutamate binding are not present in insect IRs. These receptors
are expressed along with a co-receptor in sensory neurons that do 4. Expression in non-sensory tissues
not express ORs or ORCo and, in contrast to ORs, there are different
co-receptors for IRs (Rytz et al., 2013). Although functional proper- Although OBPs fundamental role in olfaction is supported by
ties of IRs have not yet been studied in heterologous expression several studies demonstrating that the majority of these proteins
systems, their localization, and structural features suggest they is specifically expressed in antennae (Pelosi et al., 2006;
are chemosensory receptors that might function as ligand- Shanbhag et al., 2001), as the list of identified members of the
induced ion channels (Benton et al., 2009). OBP-gene family grows, it has become clear that not all of them
For insects whose navigation is guided by odorants, kinetics of are associated with sensory organs, indicating OBPs might exhibit
the olfactory system requires stray odorant molecules to be quickly different functions.
inactivated, on a millisecond timescale (Ishida and Leal, 2005). In Phormia regina, an OBP is responsible for solubilization of
There are two competing hypotheses currently trying to explain dietary fatty acids (Ishida et al., 2013). Bloodsucking insects such
signal termination. The first one indicates ODEs participate in as Culex nigripalpus and Rhodnius prolixus had OBPs found in their
ligand degradation in a way so fast that it terminates the signal gut transcriptome (Ribeiro et al., 2014; Smartt and Erickson,
(Ishida and Leal, 2005, 2008). The second hypothesis defends the 2009) which indicates these proteins could be associated with
existence of still unknown ‘‘scavengers” (Kaissling, 2009; Leal the transport of nutrients or other molecules involved in intestinal
and Kaissling, 2004) responsible for signal termination. function. OBPs have also been observed in pheromone producing
glands and reproductive organs, as in the case of OBP10 from Heli-
coverpa armigera (Sun et al., 2012b) and Aedes aegypti OBP22 (Li
3. Detection of semiochemicals et al., 2008), where they might participate in the controlled release
of semiochemicals into the environment.
OBPs role in the transport of molecules in insect antennae was Nevertheless, it is highly unlikely that all OBPs predicted from
described for the first time in Lepidoptera, using male Antheraea an insect’s genome are indeed olfactory proteins. In D. melanoga-
polyphemus antennae (Vogt and Riddiford, 1981). A few years later, ster, for instance, the OBP-gene family comprises as many as 51
a study with D. melanogaster mutants provided evidence of the par- putative OBPs, but only seven of them have been demonstrated
ticipation of OBPs in the presentation of odorants to ORs (Kim to be expressed specifically in olfactory organs of adults (Galindo
54 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

and Smith, 2001). The genome of the silkmoth B. mori contains 44 The labeling ‘‘odorant-binding protein” then should be restricted
genes encoding OBPs, but to date only one pheromone-binding to olfactory proteins of this family. It is possible that a large
protein and a few general odorant-binding proteins have been number of genes annotated from insect genomes as putative OBPs
identified (Dani et al., 2011; Krieger et al., 1996; Maida et al., are merely encapsulins (Leal, 2006).
1997). The Anopheles gambiae genome has 66 genes encoding pro- However, the majority of insect OBPs studied so far is highly
teins that have been classified as OBPs based on sequence similar- expressed in chemosensory structures and not found in non-
ity (Biessmann et al., 2005; Zhou et al., 2008). However, studies olfactory tissues (such as legs and wings, for instance), thus being
have already demonstrated that several of these genes encode pro- presumably associated with the detection of chemical signals.
teins that can actually be found in non-sensory organs (Li et al.,
2005; Biessmann et al., 2005; Pitts et al., 2011) and in association
with other structures, such as the mosquito eggshell (Amenya 5. Physicochemical properties and structural aspects
et al., 2010). A proteomic investigation of soluble olfactory proteins
suggests that female antennae express a larger number and higher Insect OBPs are small (13–17 kDa) water-soluble proteins, gen-
quantities of OBPs than males, with 24 OBPs being found in female erally consisting of chains of 130–150 amino acids (Leal, 2012;
antennae while only 19 were identified in male antennae Vieira and Rozas, 2011). These proteins are expressed with a signal
(Mastrobuoni et al., 2013). peptide that is removed during processing and is no longer found
Olfactory and non-olfactory proteins from the OBP-gene family in their mature form. OBPs usually present acidic features,
present the same structural properties. Their helix-rich structures however, it has already been discussed that, in Diptera, OBPs have
suggest these proteins encapsulate hydrophobic odorants and a wider range of isoelectric points than the acidic pIs reported for
other ligands, with the ability to transport them through aqueous lepidopteran OBPs. Thus, OBPs in dipteran species can be positively
environments. Hence, it is reasonable to presume that such stable or negatively charged at the physiological pH in insect antenna
proteins could be used in a variety of tissues where there is need (Zhou et al., 2008). A large range of pIs has also been found for
for transportation of hydrophobic molecules in aqueous media or OBP candidates in the genome of the hemipteran R. prolixus
protection of chemicals from degradation, as well as to assure a (Mesquita et al., 2015).
gradual release of semiochemicals in the environment. For this rea- According to the number of amino acid residues found in their
son, Leal (2006) mentions these proteins should be named ‘‘encap- primary structure, odorant-binding proteins are classified as long-
sulins”, to imply the common role of encapsulating small ligands. chain OBPs (160 residues), medium-chain OBPs (120 residues)

Fig. 2. Representative alignment of OBPs from different insect species. NvitOBP78 (CCD17847.1); AlinOBP13 (ACZ58084.1); BmorPBP1 (2FJY_A); BmorGOBP2 (2WC5_A);
AlinOBP10 (ACZ58081.1); TcasOBP25 (EFA04747.2); LmigOBP1 (4PT1_A); AaegOBP1 (3K1E_A); CquiOBP1 (3OGN_A); AgamOBP1 (2ERB_A); RproOBP17 (VectorBase ID:
RPRC000118); AlinOBP4 (ACZ58030.1); LlinLAP (AAC43033.1); AlinOBP6 (ACZ58032.1); AlucOBP5 (AEP95759.1); AlinOBP1 (ACZ58027.1); ApisOBP1 (CAR85628.1); ApisOBP8
(CAR85635.1); DmelLUSH (NP_524162.1); AmelOBP5 (3R72_A). The pattern of six cysteines is well conserved. The conserved cysteines are marked with asterisks. Nvit –
Nasonia vitripennis; Alin – Adelphocoris lineolatus; Bmor – Bombyx mori; Aluc – Apolygus lucorum; Apis – Acyrthosiphon pisum; Dmel – Drosophila melanogaster; Tcas – Tribolium
castaneum; Lmig – Locusta migratoria; Aaeg – Aedes aegypti; Cqui – Culex quinquefasciatus; Agam – Anopheles gambiae; Rpro – Rhodnius prolixus; Llin – Lygus lineolaris; Amel –
Apis mellifera.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 55

and short-chain OBPs (100 residues). Primary sequences of this Later, the structure of unligated BmorPBP1 was determined by
type of proteins are highly divergent and amino acid identity two studies involving solution NMR. The first one, attempted at
between OBPs from the same species might be lower than 10%. acidic pH (4.5), revealed the existence of a seventh helix occupying
The correct assignment of a protein into the OBP-family is mainly a position corresponding to the binding pocket observed in the
based on a pattern of six cysteines whose relative positions are crystal structure of the BmorPBP1-bombykol complex (Horst
well-conserved across all insect orders (Fig. 2) (Briand et al., et al., 2001) (Fig. 3b), providing a reasonable explanation for the
2001a; Leal et al., 1999). observation that PBPs generally do not bind ligands below pH 5.0
OBPs are usually comprised of six a-helical domains folded into (Briand et al., 2001b; Prestwich, 1993). The next study, performed
very compact and stable globular structures (Sandler et al., 2000; at the physiological pH of the sensillum lymph (6.5), reported a
Tegoni et al., 2004). Experiments have demonstrated that the six quite similar structure to the one obtained for the complex by
cysteines form three interlocked disulfide bridges in an arrange- X-ray crystallography that exhibits a hydrophobic cavity with suf-
ment that provides great stability to the protein’s three- ficient volume to accommodate a bombykol molecule, concluding
dimensional structure and also seems to be a conserved feature this should be the reactive conformation (Lee et al., 2002).
among OBPs. This cysteine pattern has become a ‘‘signature” for The occurrence of pH-dependent conformational changes was
insect classical OBPs (Pelosi et al., 2006). evidenced for the first time in Lepidoptera. Studies involving
Based on their amino acid sequences, OBPs initially described in BmorPBP1 had already described two distinct pH-dependent
Lepidoptera were subdivided into five classes: PBPs (pheromone- conformational states (Damberger et al., 2007) and also that the
binding proteins), GOBP1s (type 1 general odorant-binding pro- protein binds bombykol with high affinity at the sensillum lymph
teins), GOBP2s (type 2 general odorant-binding proteins), ABP1s pH, but shows no affinity for the pheromone at low pH, so it was
(type 1 antennal binding proteins) and ABP2s (type 2 antennal postulated that a transition from the high or near-neutral pH
binding proteins), previously known as ABPX (antennal binding (B-form, also called open form) to the low pH (A-form, also called
proteins) (Gong et al., 2009a; Krieger et al., 1996). However, since closed form) of the protein would be physiologically relevant
this classification is not well defined for other insect orders, OBPs (Wojtasek and Leal, 1999). In such context, the comparison
were then categorized according to their conserved cysteine pat- between structures BmorPBP1A and BmorPBP1B, resolved by
terns. The so-called classical OBPs present six conserved cysteines NMR, and the structure of the complex BmorPBP1-bombykol,
in their sequences (Fig. 2), while C-Minus OBPs exhibit four or five. determined by X-ray crystallography, suggests a biological func-
OBP14 from Apis mellifera was the first C-Minus OBP to have its tion for the observed conformational change. The pheromone
structure determined and the study demonstrated the presence would bind the active form of the protein, BmorPBP1B, found at
of only two disulfide bonds (Spinelli et al., 2012). The Plus-C type the physiological pH of the sensilla lymph and be transported
is distinguished by the presence of at least two additional cysteines through the aqueous environment. When the BmorPBP1-
and a conserved proline after the sixth cysteine (Xu et al., 2003; bombykol complex approaches the OR, located on the membrane
Zhou et al., 2004b). Crystal structure of AgamOBP47, a Plus-C of sensory neurons, where the local pH is reduced, the OBP
OBP from A. gambiae, revealed the existence of as many as six undergoes a conformational change to its acidic form, BmorPBP1A,
disulfide bridges (Lagarde et al., 2011a). The family referred to as triggering ligand release by the formation of a seventh a-helix that
dimer OBPs, described in Drosophila (Hekmat-Scafe et al., 2002; occupies the binding pocket.
Zhou et al., 2004b), seems to be formed by two classical six- The C-terminus of BmorPBP1 is composed mostly of nonpolar
cysteine motifs fused together (Zhou et al., 2004a). Finally, the amino acid residues, except for three acid residues, Asp132,
atypical subfamily was first described as displaying a classical Glu137, and Glu141, which lie on the surface of the helix and are
domain in the N-terminal region elongated by a not well character- well-conserved in Lepidoptera. Of these, Asp132 (or Glu132 in
ized extended C-terminal region. It has more recently been demon- Amyelois transitella) and Glu141 form salt bridges with protonated
strated that atypical OBPs, which have so far been identified only in histidine residues: His70 in A. polyphemus and B. mori (Damberger
mosquitoes (Xu et al., 2003; Zhou et al., 2008), actually present two et al., 2007; Xu and Leal, 2008) or His80 in A. transitella (Xu et al.,
domains that are homologous to those of classical OBPs, and then 2010c), and His95, respectively, at low pH. Xu et al. (2010c)
were renamed to two-domain OBPs and considered as belonging to demonstrated that in AtraPBP1, two salt bridges induced by acidic
the dimer OBPs subfamily (Vieira and Rozas, 2011). conditions promote the formation of a seventh helix at the C-
terminal region. This mechanism of salt bridges formation was
named histidine protonation switch. The importance of histidine
6. Three-dimensional structure-based mechanism of action residues protonation had already been demonstrated in previous
studies (Damberger et al., 2007; Zubkov et al., 2005) and it was
Crystal X-ray diffraction analysis and solution nuclear magnetic proposed that such mechanism provides stabilizing forces for the
resonance (NMR) are powerful tools for determining the three- insertion of the new helix into the binding cavity, taking into con-
dimensional structure of a protein, whether in its free form or in sideration that both ends of the helix form salt links (His80-Glu132
a complex with its ligand. To date, several insect OBP structures and His95-Glu141) that position the C-terminal helix inside the
have been solved and further details can be found in Table 1. pheromone binding site. Deprotonation of histidine residues at
The first OBP to have its structure determined both alone and neutral pH disrupts the salt bridges, promoting the ejection of
bound to its ligand bombykol, was BmorPBP1 from B. mori. The the seventh helix outward from the binding pocket. Extrusion of
structure of a BmorPBP1-bombykol crystal grown at pH 8.2 was the now unstructured C-terminus allows exposure of hydrophobic
initially resolved by X-ray crystallography (Fig. 3a). In this com- residues in the protein core which then can interact with the pher-
plex, BmorPBP1 consists of a compact structure formed by a omone to enable binding (Xu et al., 2010c). Mutation of charged
roughly conical arrangement of six a-helices knitted together by residues involved in the switch (His95Ala, Asp132Asn, and
three disulfide bridges. Bombykol is bound in a completely Glu141Ala) dramatically affect the pH-dependent binding of bom-
enclosed hydrophobic cavity, allowing its transport through hydro- bykol to BmorPBP1 (Xu and Leal, 2008). In AtraPBP1, mutation of
philic environments. The surface of BmorPBP1, however, is covered His80 and His95 residues inhibit the formation of salt bridges at
with charged residues, which is not surprising for such a soluble both ends of the C-terminus helix, allowing the OBP to bound the
protein (Sandler et al., 2000). pheromone at low pH (Xu et al., 2010b).
56 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

Table 1
List of published odorant-binding protein three-dimensional structures.

OBP Name Species Determined by Complexed with Reference


OBP1 Aedes aegypti X-ray (1.85 Å) Serendipitous ligand (consistent with PEG) Leite et al. (2009)
PBP1 Amyelois transitella NMR Free Xu et al. (2010)
OBP1 Anopheles gambiae X-ray PEG Wogulis et al. (2006)
(1.5 Å)
OBP7 Anopheles gambiae X-ray Serendipitous ligand (consistent with a 16-carbon-long fatty Lagarde et al. (2011b)
(1.85 Å, 2.19 Å) acid), AZO
OBP47 Anopheles gambiae X-ray Free Lagarde et al. (2011a)
(1.8 Å)
OBP1 Anopheles gambiae X-ray DEET, PEG Tsitsanou et al. (2012)
(1.60 Å)
OBP20 Anopheles gambiae X-ray (1.80 Å) Free, indole, heptene-2-one, DEET, others Ziemba et al. (2013)
NMR
OBP1 Anopheles gambiae X-ray 6-Methyl-5-hepten-2-one Murphy et al. (2013)
(2.20 Å)
OBP48 Anopheles gambiae X-ray PEG Tsitsanou et al. (2013)
(3.30 Å)
PBP1 Antheraea NMR Free Mohanty et al. (2004)
polyphemus
PBP1 Antheraea NMR Free Zubkov et al. (2005)
polyphemus
PBP1A Antheraea NMR Free Damberger et al. (2007)
polyphemus
ASP2 Apis mellifera NMR Free Lescop et al. (2001)
ASP1 Apis mellifera X-ray Serendipitous ligand (n-butyl-benzene-sulfonamide) Lartigue et al. (2004)
(1.6 Å)
ASP1 Apis mellifera X-ray Free, serendipitous ligand (nBBSA), 9-ODA, QMP blend, HDOA Pesenti et al. (2008)
(2.6 Å, 1.60 Å, 2.15 Å, 2.25 Å,
2.00 Å)
ASP1 Apis mellifera X-ray Free, serendipitous ligand (nBBSA), 9-ODA Pesenti et al. (2009)
(2.56 Å, 1.54 Å, 1.85 Å)
OBP14 Apis mellifera X-ray Free, 1-NPN, citralva, eugenol Spinelli et al. (2012)
(1.88 Å, 1.22 Å, 1.24 Å, 1.6 Å)
PBP1 Bombyx mori X-ray Bombykol Sandler et al. (2000)
(1.8 Å)
PBP1A Bombyx mori NMR Free Horst et al. (2001)
PBP1B Bombyx mori NMR Free Lee et al. (2002)
PBP1 Bombyx mori X-ray Iodohexadecane, bell pepper odorant Lautenschlager et al.
(1.9 Å, 2.0 Å) (2007)
GOBP2 Bombyx mori X-ray Free, bombykol, bombykal, other structural analogues Zhou et al. (2009)
(1.60 Å , 1.95 Å, 1.74 Å)
OBP1 Culex X-ray (1.30 Å) MOP Mao et al. (2010)
quinquefasciatus NMR
OBP76a Drosophila X-ray Ethanol, n-propanol, n-butanol Kruse et al. (2003)
(LUSH) melanogaster (1.49 Å, 1.45 Å, 1.25 Å)
OBP76a Drosophila X-ray Ethanol, butanol Thode et al. (2008)
(LUSH) melanogaster (1.8 – 2.0 Å)
OBP76a Drosophila X-ray cVA Laughlin et al. (2008)
(LUSH) melanogaster (1.4 Å)
PBP Leucophaea maderae X-ray Free, ANS, H3B2 Lartigue et al. (2003)
(1.7 Å, 1.6 Å, 1.7 Å)
OBP1 Locusta migratoria X-ray Free Zheng et al. (2015)
(1.65 Å)
OBP3 Megoura viciae X-ray Free Northey et al. (2016)
(1.3 Å)
OBP3 Nasovonia ribisnigri X-ray Free Northey et al. (2016)
(2.02 Å)

Abbreviations used are: PEG, polyethylene glycol; DEET, N,N-diethyl-m-toluamide; AZO, 4-hydroxy-40 isopropyl-azobenzene; 1-NPN, N-phenyl-1-naphthylamine; nBBSA, n-
butyl benzenesulfonamide; 9-ODA, 9-keto-2(E)-decenoic acid; QMP, queen mandibular pheromone; HDOA, hexadecanoic acid; MOP, mosquito oviposition pheromone
((5R,6S)-6-acetoxy-5-hexadecanolide); cVA, 11-cis-vaccenyl acetate; ANS, 1-anilino-8-naphthalenesulfonic acid; H3B2, 3-hydroxy-butan-2-one.

Even though BmorPBP1 is reported to bind equally well to both This seems to be a good explanation for the fact that BmorGOBP2
bombykol and bombykal (Grater et al., 2006; He et al., 2010; Zhou presents greater affinity towards bombykol than bombykal. The
et al., 2009) by the mechanism described above, BmorGOBP2 latter is not able to form the additional hydrogen bond to Glu98
seems to be able to discriminate between these two compounds. since its polar group is an aldehyde and, as such, it is only able
Zhou et al. (2009) demonstrated that bombykol adopts different to accept hydrogen bonds, in contrast to the hydroxyl group from
conformations when it binds to each one of these proteins. The bombykol, which is able to both accept and donate, explaining the
hydroxyl group of bombykol forms a hydrogen bond with the side- ligand discrimination by BmorGOBP2 at the structural level. These
chain of Ser56 when bound to BmorPBP1, while binding to Bmor- authors also reported that in the ligand-free structure of Bmor-
GOBP2 involves a hydrogen bond with Arg110 via a water GOBP2, the C-terminus does not occupy the binding site as seen
molecule. The extended interaction via a water molecule makes in BmorPBP1 and that there is no conformational change among
the formation of an additional hydrogen bond to Glu98 possible. the protein upon binding different sex pheromone components
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 57

Fig. 3. Three-dimensional structures of OBPs. a-Helices are shown in red, loops in green and disulfide linkages are highlighted in blue. (a) Bombyx mori pheromone-binding
protein, BmorPBP1, complexed with its specific ligand, the pheromone bombykol, at pH 6.5. The C-terminus is not the structure at this pH (Sandler et al., 2000). PDB ID Code:
1DQE. (b) The same protein at pH 4.5. The C-terminus is folded into a seventh a-helix that occupies the binding pocket, triggering ligand release (Horst et al., 2001). PDB ID
Code: 1GM0. (c) Aedes aegypti odorant-binding protein, AaegOBP1. The C-terminus is pulled to the core of the protein to form part of the binding pocket wall, which can
function as a ‘‘lid” for the release of ligands (Leite et al., 2009). PDB ID Code: 3K1E. (d) Culex quinquefasciatus odorant-binding protein, CquiOBP1 complexed with a mosquito
oviposition pheromone (MOP). In contrast to previously published ligand-bound OBP structures, a large part of MOP is not located inside the central cavity. Instead, MOP has
its long lipid ‘‘tail” bound to a hydrophobic tunnel formed by helices 4 and 5 and only its lactone/acetyl ester ‘‘head” sticks into the central cavity (Mao et al., 2010). PDB ID
Code: 3OGN. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

and analogues. However, it is important to notice that the seventh unfolding of a1a, similar to ApolPBP, but since Ser56 does not
helix formed by the C-terminus was only observed in BmorPBP1 at move away, helix a7 can form and competitively displace the
low pH and, in this study, BmorGOBP2 crystals were grown at pH ligand from the open binding cavity, Details of the pH-dependent
8.5, underlining the need for further studies in order to fully eluci- conformational switch in PBPs seem to vary for different moth spe-
date the binding and release mechanism of BmorGOBP2 (Zhou cies and appear to depend on the structure of the binding site and,
et al., 2009). in particular, the amino acid residues responsible for the direct
Nonetheless, internalization of the C-terminus in form of a helix interaction with the ligands. Studies regarding ligand-interaction
into the binding cavity at acidic pH has already been shown not to kinetics of a pheromone-binding protein from the gypsy moth
be a general feature of all lepidopteran PBPs. Zubkov and col- Lymantria dispar provided evidence for the existence of yet another
leagues in 2005 observed that PBP1 from Antheraea polyphemus mechanism of ligand binding for lepidopteran OBPs. LdisPBP2
undergoes a reorientation of helices a1, a3, and a4 at acidic pH seems to bind hydrophobic ligands in two steps: one rapid, the
caused by protonation of His69, His70 and His95, which provides other slow (Gong et al., 2009b). Authors suggest that the rapid
the driving force behind the opening of the ligand binding cavity phase involves the uptake of the ligand by the C-terminus and
and the release of the pheromone molecule from its carrier protein binding to an external site located near this region, where the ini-
near the membrane (Zubkov et al., 2005). The authors postulate tial interactions between ligand and OBP occur. By contrast, the
that these structural differences are likely caused by the difference slow step corresponds to the specific reorientation and embedding
in the position of the residue which forms a hydrogen bond with of the ligand to the internal binding pocket. It appears that both
the ligand (Ser56 for BmoPBP1 and Asn53 for ApolPBP1). In the steps are necessary in order to obtain a final complex. First, LdisPBP
neutral pH structure of ApolPBP1, there is a kink in the middle of and the ligand form an intermediate complex through diffusion-
a3 that causes the side-chain of Asn53 to be a part of the phero- controlled collision. Then, the ligand is relocated from an external
mone binding pocket. At acidic pH, the kink disappears and the binding site to a different internal one, resulting in a more stable
helix straightens at this point with Asn53 moving away from the ligand-OBP complex (Gong et al., 2009c).
protein core. Once helix a3 loses the kink at Asn53, the interaction Unlike what has been described for Lepidoptera, most dipteran
between a1b and a3 is also disrupted and helix a1 straightens out, OBPs do not possess a C-terminal region long enough to form a
opening the cavity entrance and allowing the ligand to be released. new helix that would compete with the ligand for the binding
In the case of BmorPBP1, helix a3 is straight and the residue cavity. However, AgamOBP1, AaegOBP1 and CquiOBP1 also
responsible for pheromone headgroup interaction (Ser56) is found undergo pH-dependent conformational changes associated with
inside the cavity at both neutral and acidic pH. Therefore, it is loss of binding affinity (Leite et al., 2009; Mao et al., 2010;
possible that at low pH, the cavity in B. mori PBP opens due to Wogulis et al., 2006), suggesting the molecular mechanism
58 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

underlying pH-dependent affinity, in this case, would be different neurons in moths (Forstner et al., 2008; Rogers et al., 2001) and
from the observed in BmorPBP1. Leite et al. (2009) described that genes encoding such proteins have already been identified in mos-
the C-terminal region of these proteins acts as a ‘‘lid” over the quitoes, however, their function remains intriguing since no sex
binding cavity (Fig. 3c), in a feature that is observed only in Diptera pheromone communication has been characterized yet for these
(Wogulis et al., 2006). This ‘‘lid” is held in place by a triad of well- insects. It is possible that pheromone signaling is more specialized
conserved hydrogen bonds (Mao et al., 2010) that seems to work as compared with general odor detection, requiring additional factors
a pH-sensitive hinge. C-terminal ‘‘lids” of these OBPs differ only in such as SNMPs, supporting the idea that a SNMP with local pH
the last residue, where Val125 is found in AgamOBP1 and might be involved in pheromone release by the domain-swapped
CquiOBP1, and Ile125 in AaegOBP1. Carboxylate oxygens of such dimer of AmelASP1 at neutral pH. In 2013, AgamOBP48 was
residues form hydrogen bonds with the hydroxyl of Tyr54 and reported as the first example of a domain-swapped dimer OBP in
the d-nitrogen of His23 in AgamOBP1 and CquiOBP1 (Mao et al., dipteran species (Tsitsanou et al., 2013).
2010; Wogulis et al., 2006), and Arg23 in AaegOBP1 (Leite et al., Lartigue et al. (2004) demonstrated that LmaPBP binding cavity,
2009). It is likely that when the OBP-odorant complex reaches although containing a majority of hydrophobic side-chains, con-
the proximity of dendritic membrane, where the pH is reduced, tains also a significant amount of polar or charged residues. Such
some of these interactions are disrupted due to carboxylate proto- information is in contrast with the cavities of other OBPs, which
nation, leading the C-terminus to move away from the binding are almost exclusively formed by hydrophobic residues. However,
cavity, thus ‘‘opening the lid’’ and promoting ligand release. the author highlights the existence of these polar/charged residues
Contrary to other OBPs described in the literature, CquiOBP1 is consistent with the hydrophilic nature of Leucophaea madere
from Culex quinquefasciatus was observed to bind the oviposition pheromone, so the protein is able to bind small hydrophilic ligands
pheromone MOP in a different fashion than the usual, where the such as H3B2 (3-hydroxy-butan-2-one), the main component of
ligand is completely ‘‘buried” in the binding cavity. In this case, this species pheromonal blend. The absence of the seventh helix
the pheromone long lipid ‘‘tail” binds to a hydrophobic tunnel together with the evidence of a hydrophilic cavity binding a hydro-
formed between helices 4 and 5, with only its lactone/acetate head philic ligand suggests that a more classical mechanism of direct
sticking into the central cavity (Mao et al., 2010) (Fig. 3d). The ligand release might be used. Also, the fact that LmaPBP binds to
authors hypothesize that CquiOBP1 does not recognize the specific a compound highly soluble in water raises doubt about the role
functional group of MOP but rather recognizes the length of lipid of OBPs as passive transport proteins and reinforces the hypothesis
chains that fit into its hydrophobic tunnel. The same hydrophobic of a mechanism in which the OBP-odorant complex would activate
channel has been reported in AgamOBP1 (Wogulis et al., 2006) and the receptors. OBP conformational changes induced by pheromone
AaegOBP1 (Leite et al., 2009), and it was described later that binding have been reported to be required for the activation of T1
AgamOBP1 binds DEET through this hydrophobic region neurons in D. melanogaster and details regarding this mechanism
(Tsitsanou et al., 2012). are discussed in the next section.
For the ASP1, from Apis mellifera, ligand affinity is pH- Even though there are over 300 hemipteran OBPs annotated in
dependent in an opposite way to the observed for other OBPs. NCBI GenBank, crystal structures of aphid OBPs were reported only
Experiments have shown that, at pH 4.0, the protein exhibits high very recently (Northey et al., 2016). Analysis of OBPs from the
binding affinity for 9-keto-2-(E)-decenoic acid (9-ODA), the main vetch aphid Megoura viciae and from the currant-lettuce aphid
component of the queen bee pheromone, while at pH 7.0 the Nasonovia ribisnigri revealed similar structures to other classical
OBP presents itself as a domain-swapped dimer with ligand affinity OBPs: six helices connected by extended loops, with three disulfide
ten times lower (Pesenti et al., 2008). Authors postulate that in bridges formed by the six conserved cysteine residues. However,
solution at pH 7.0, the loose C-terminus can recruit the the two OBPs also have some unique features which are different
N-terminus of another ASP1 molecule to form a more stable from other classical OBPs. There is a significant groove in the sur-
domain-swapped dimeric form. This form is still able to bind face of aphid OBPs formed primarily from the N-terminal region of
ligands, but with lower affinity when compared to the acidic the chains and parts of helices a1 and a2, with no significant inter-
ASP1 monomeric form. Dimerization seems to be strongly related nal binding pockets being predicted. Besides the absence of an
to residue Asp35, giving that mutants Asp35Asn and Asp35Ala, internal binding cavity, OBP3 also differs from the vast majority
which favor the formation of hydrogen bonds at neutral pH, exhibit of OBPs with respect to the location of the N-terminus. In aphid
exactly the same monomeric structure in every pH tested and OBP3, this region is in an unusual place close to the center of the
increased binding affinity at pH 7.0 (Pesenti et al., 2009). Although molecule, where the C-terminus is normally found. Molecular
their findings were not conclusive, authors postulate that different docking studies indicated that known ligands were mainly docked
OBPs very likely exhibit different mechanisms of ligand release, in the groove and Tyr30, a residue which lies close to the predicted
which might occur at low pH or at neutral or high pH, such as surface binding site, may play an important role in ligand binding.
the case of AmelASP1, indicating there might be a factor triggering Molecular dynamics showed that Tyr30 is displaced towards the
ligand release other than the pH drop at the vicinity of the mem- protein in MvicOBP3, while in NribOBP3, Tyr30 is not significantly
brane, as reported for most OBPs. In this context, the discovery that displaced, located towards the binding site. Authors discuss that
OBPs may interact with SNMPs (Jin et al., 2008) offers an appealing the most likely explanation is that Tyr30 is involved in the crystal
alternative explanation. The trigger for ligand release could be the packing/dimerization in MvicOBP3 and is relaxing towards its
ionization state of the SNMP ectodomain, which should emerge out probable monomeric conformation, similar to what is seen in
of the membrane surface and impose its own electrostatic influ- NribOBP3. However, a detail that could easily go unnoticed might
ence on PBPs to capture them and lead to pheromone release at provide valuable information: NribOBP3 has been crystallized at
the vicinity of the OR. This ionization state may create a local pH, pH 5.0 whereas MvicOBP3 at pH 7.5. Different conformations
in some cases acidic and in other cases neutral or basic, triggering adopted by Tyr30 in such structures may be directly related to a
the conformational changes of PBPs and the passage of ligands to pH-dependent binding and release mechanism, a vastly reported
SNMPs. At least for olfactory neurons of T1 trichoid sensilla of D. feature among insect OBPs, indicating that the unique features
melanogaster, SNMPs are reported to be necessary for normal func- described for the two aphid OBP3s (no internal binding pocket, sur-
tioning detection of 11-cis-vaccenyl acetate, a pheromone which face binding site and involvement of the N-terminus in forming the
mediates social behavior in this species (Jin et al., 2008). SNMPs binding pocket) need further evaluation in order to fully elucidate
were also shown to be associated with sex pheromone detecting their role in OBPs biological function.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 59

On the whole, ligand binding and release by OBPs seems to pre- from the LUSH-cVA structure, indicating the salt bridge possibly
sent specific features depending on the organism in question, maintains LUSH in the inactive state. Mutant Asp118Ala is able
which is not exactly surprising given the fact that a unique to stimulate T1 neurons in the absence of cVA. In fact, T1 sensilla
mechanism would hardly be consistent with insect diversity. It is infused with the mutant show no additional increase in activity
not very likely that selective pressures that led moths, for instance, when stimulated with cVA, possibly indicating that the protein
to have a very sensitive and dynamic olfactory system have acted has adopted an activated conformation that is not further activated
in the same way in the fruit fly. Analyses of OBPs from 12 by cVA (Laughlin et al., 2008). Data obtained in these experiments
Drosophila genomes suggest that OBPs play an important role in indicates that conformational changes in LUSH mediate T1 neuron
ecological diversification (Vieira et al., 2007), therefore, it is activation through OR67d and SNMP and that LUSH is not a passive
possible that behavioral and ecological differences between insect carrier of cVA. The pheromone, in turn, stimulates the conversion
species may affect not only the types of semiochemicals they are from the inactive form of LUSH to an active conformation that is
able to perceive, but also how they detect and process these chem- capable of triggering the activation of T1 neurons. These observa-
ical signals. tions are consistent with the model in which an OBP-odorant com-
plex is required for the activation of the OR. However, it should not
go unmentioned that this model was contested by a study indicat-
7. Activation of ORs ing that cVA directly activates ORs, where authors conclude that
neurons can be activated in the absence of LUSH, but not SNMP
Up to this point, two models have been proposed in order to or OR67d (Gomez-Diaz et al., 2013).
explain ORs activation mechanism: the first one suggests the odor-
ant per se activates the receptor while the second model supports
the idea that the OR is activated by a specific OBP-odorant 8. Binding affinity and selectivity
complex.
Considerations mentioned above regarding moths and mosqui- pH-dependent conformational change associated with change
toes support the hypothesis that odorants are solubilized upon in binding affinity is reported to be common to OBPs from a variety
being encapsulated in the binding cavity of OBPs and transported of insects. Moreover, it also seems to show a special character to
through sensillar lymph, therefore being protected from degrada- different groups. Binding activity of OBPs to small organic
tion by ODEs. Eventually, when the complex reaches the dendrite compounds, such as pheromones and other semiochemicals, is
region, the decrease in local pH triggers ligand release, which then the key feature of these proteins and provides essential informa-
activates the OR. In such cases, OBP-ligand complexes have a short tion for understanding their physiological function (Fan et al.,
life at low pH values, corroborating the idea that odorants stimu- 2011; Pelosi et al., 2006).
late olfactory receptors in an independent fashion. This model is The ligand specificity of insect OBPs has not yet been demon-
consistent with the experimental observation that receptors are strated conclusively. PBPs (at least from Lepidoptera) present great
activated by their specific ligands in the absence of OBPs binding specificity towards pheromones, with the pheromone
(Nakagawa et al., 2005). However, contrary to this direct ligand bombykol (IUPAC name (10E,12Z)-hexadeca-10,12-dien-1-ol)
hypothesis, a whole different way of signal transduction has been being described as the specific ligand for the B. mori PBP
suggested for the odorant-binding protein LUSH, from D. melanoga- (BmorPBP1) (Sandler et al., 2000). Also, the male moth A. polyphe-
ster (Xu et al., 2005). mus has three different PBPs which interact with components of
In Drosophila antennae, there are three morphologically distinct the sex pheromone of this species: ApolPBP1 exhibits higher affin-
types of sensilla that contain odor and pheromone-sensitive recep- ity towards (E,Z)-6,11-hexadecadienal, ApolPBP3 preferentially
tor neurons: basiconic and coeloconic sensilla house neurons that binds (E,Z)-4,9-tetradecadienyl acetate and ApolPBP2 binds only
detect general odorants, while trichoid sensilla are thought to be (E,Z)-6,11-hexadecadienal (Maida et al., 2003). However, OBPs
specialized for pheromone reception (Benton, 2007; Smith, may also bind to a wide range of odorant molecules and GOBPs
2007). A subset of trichoid sensilla, the T1 sensilla, contains neu- seem to have a rather broad ligand-binding affinity in comparison
rons that are activated specifically by the cVA pheromone with specificity. A good example is the LstiGOBP2, a general
(Kurtovic et al., 2007; Xu et al., 2005). The perception of cVA is odorant-binding protein from the meadow moth Loxostege
described as complex and requires the participation of a specific sticticalis. This OBPs has been shown to have high affinity to plant
odorant receptor, OR67d, and, at least two more factors: a SNMP volatiles from essential oils such as (E)-2-hexenal and (Z)-3-hexen-
and the OBP LUSH. OR67d is exclusively expressed in T1 neurons 1-ol, and to the pheromone component (E)-11-tetradecen-1-yl
and participates in cVA response (Ha and Smith, 2006; Kurtovic acetate (Yin et al., 2012). Nonetheless, there is also evidence
et al., 2007). The SNMP is located along with the olfactory receptor showing that PBPs can bind to general odorants in addition to
complex in dendrites of T1 neurons (Benton et al., 2007) and infu- sex pheromone components and analogues (Jin et al., 2014; Song
sion of a SNMP antiserum in T1 sensillar lymph results in loss-of- et al., 2014) and that, in some cases, GOBPs exhibit preference to
function mutants (Jin et al., 2008), suggesting SNMP is directly special sex pheromone components besides host-plant odorants
required in pheromone-sensitive neurons for cVA sensitivity. (Gong et al., 2009c; He et al., 2010).
A study conducted with LUSH-deficient mutants revealed this Insect OBPs were analyzed for the first time by native gel
protein is necessary for typical cVA-induced behavior and normal electrophoresis of an antenna extract previously incubated with
sensitivity of T1 neurons (Xu et al., 2005). Structural studies point the tritium-labeled pheromone (E,Z)-6,11-hexadecadienyl acetate,
out unique conformational changes mediated by residue Phe121 followed by autoradiography (Vogt et al., 1985). Later, this method
are induced upon binding of LUSH with cVA pheromone was widely used either as described in the original paper or with
(Laughlin et al., 2008). Interaction of cVA with residue Phe121 from modifications. In some studies, radioactive photoaffinity labels with
LUSH seems to mediate the conformational change into the active structural similarity to the pheromone were used (Feng and
form of the OBP through disruption of a salt bridge between resi- Prestwich, 1997), however, the preparation of radioactive pho-
dues Asp118 and Lys87. Experiments in which Asp118 was substi- toaffinity labels is often complicated and expensive. In early work
tuted by Ala were able to generate an active conformation of the to identify pheromone-binding proteins, radiolabeled pheromones
protein. Disruption of the salt bridge by mutagenesis results in a were employed in qualitative binding assays (Plettner et al., 2000).
conformational shift at the C-terminus virtually indistinguishable With the availability of recombinant proteins, radiolabeled
60 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

pheromones can be used in quantitative assays in which free ligands one which binds a hydrophilic ligand, 3-hydroxy-butan-2-one, a
are separated by gel filtration from bound ligands. Although a valu- component of the pheromonal blend for this species (Riviere
able tool for studies of pheromone-PBP interactions, this type of et al., 2003). The specific binding of A. mellifera OBP3 to (E)-b-
binding assay has limited application in screening programs because farnesene was also discovered by competitive binding experi-
radioactive test ligands are required. Leal et al. has described a bind- ments, this time using N-phenyl-1-naphthylamine (1-NPN), which
ing protocol for OBPs that does not require radioactive ligands is also extensively used as a fluorescent probe (Qiao et al., 2009).
named ‘‘the cold binding assay” (Leal et al., 2005b). After incubation The binding affinity of AfunOBP1 from Anopheles funestus was
of test compounds with an OBP, the free ligand is removed by filtra- tested against different compounds that elicited significant elec-
tion, whereas the protein-bound ligand is retained in the centrifugal troantennographic (EAG) responses, and 2-undecanone was identi-
device and extracted with an organic solvent containing an internal fied as the best ligand (Xu et al., 2010a). In a similar study, A.
standard. Binding is quantified by gas chromatography and the iden- gambiae AgamOBP1 was demonstrated to specifically bind indole
tity of the recovered ligand is confirmed by mass spectrometry. This (Biessmann et al., 2010). ANS (1-anilino-8-naphthalenesulfonic
method presents the advantage that it can be applied to mixtures of acid) is too a fluorescent probe widely used in this type of assays
organic compounds, all incubated at the same time with the protein, and was employed by Wojtasek and Leal (1999) in a remarkable
allowing the identification of the best ligands in a single experiment. study regarding BmorPBP1 pH-dependent structural flexibility.
Nowadays, fluorescence binding assays represent the most used In 2014, a possible involvement of the protein CcapOBP83a-2,
technique in studies regarding binding properties of OBP to from Ceratitis capitata, in olfactory processes and its specificity to
putative ligands. Such method uses a fluorescent probe that, upon (E,E)-a-farnesene were exposed by this kind of assay (Siciliano
protein binding, has its emission spectrum modified due to the less et al., 2014). More recently, using a fluorescent probe and a panel
polar environment inside the binding cavity, resulting in a blue shift of 34 physiologically relevant compounds, Yin and collaborators
accompanied by a marked increase in intensity. Competitive bind- investigated binding affinities of three OBPs from C. quinquefascia-
ing experiments using a fluorescent probe (Fig. 4) are widely used tus, concluding CquiOBP2 is a carrier for the oviposition attractant
to investigate OBPs selectivity by comparing their binding affinities skatole, and CquiOBP1 and CquiOBP5 might transport the oviposi-
to physiologically relevant compounds, surveying their ability to tion pheromone MOP, a human-derived attractant nonanal, and
preferentially transport some ligands over others. One limitation the insect repellent picaridin (Yin et al., 2015). Not long ago, fluores-
of the method is the need of a fluorescent compound that exhibits cence ligand-binding assays revealed that, in the oriental fruit moth
some affinity for the protein of interest. However, when applicable, Grapholita molesta, GOBP1 has dual functions in recognition of host
it offers the unique advantage of allowing binding measurements at plant volatiles and sex pheromone components, while GOBP2 is
the equilibrium, besides being simple, fast and safe. mainly involved in the perception of dodecanol, a minor sex phero-
Fluorescence assays were used for the first time in insects to mone (Li et al., 2016). Fluorescence competition assays with OBP11
demonstrate the unspecific binding capacity of OBPs from from the alfalfa plant bug Adelphocoris lineolatus provided insights
Antheraea polyphemus and Mamestra brassicae to a series of ligands into understanding the physiological role of an OBP highly
in a competitive experiment using the probe 1-aminoanthracene expressed in mouthparts of this hemipteran species (Sun et al.,
(AMA) (Campanacci et al., 2001). This kind of experiment also 2016). Results clearly showed that AlinOBP11 displayed preferential
revealed that, in contrast with other OBPs described so far, the binding abilities to non-volatile host plant secondary metabolites
PBP of the cockroach L. maderae, LmaPBP, seems to be the only than to volatile compounds, suggesting this protein could act as a

Fig. 4. Schematic representation of binding assays using a fluorescent probe. (a) This technique uses probes that fluoresce weakly in aqueous media, but become very strongly
fluorescent in non-polar or hydrophobic environments, such as the binding cavity of an OBP. Modified from Zhuang et al. (2013). (b) In order to investigate binding affinities
of OBPs to different compounds, test ligands are added to the OBP-probe complex. Competition for the binding site displaces probe molecules from the hydrophobic cavity,
resulting in a decrease in overall fluorescence proportional to the amount of bound ligand.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 61

carrier in gustatory system for non-volatile compounds detection It is reasonable to assume that computational methods will be
when plant bugs begin to search suitable food substrates by using increasingly used in screenings for potential semiochemicals. In sil-
their mouthparts to rub or to tap plant surfaces. ico studies are able to provide a fine selection of ligands from
In some cases, when a proper probe cannot be found, it is pos- extensive libraries and their association with in vitro biochemical
sible to measure the intrinsic fluorescence of a tryptophan residue analysis, such as the widely used fluorescence assays described
in order to monitor the presence of a ligand, since it has been above, is undoubtedly becoming a very important tool in the field
proposed that Trp37 is positioned in the binding pocket and that of chemical ecology, with the purpose of providing valuable infor-
its interaction with the ligand appreciably affects its fluorescence mation for the development of new pest control strategies based
properties (Campanacci et al., 2001; Leal et al., 2005b). Observation on the manipulation of insect behavior and the prevention of
of intrinsic fluorescence changes revealed, for instance, the exis- human diseases transmitted by insect vectors.
tence of three different kinds of interactions between ApolPBP1,
from A. polyphemus, and the components of its pheromonal blend, 9. Applied research
(E6,Z11)-hexadecadienal, (E6,Z11)-hexadecadienyl-1-acetate and
(E4,Z9)-tetradecadienyl-1-acetate (Bette et al., 2002). The comprehension of behaviors of various agricultural pests
The binding potential between a characterized OBP and differ- and pathogenic insects strongly depends on a better understanding
ent ligands can also be estimated by computational methods. of the molecular mechanism underlying the olfactory system.
Although most studies assess binding affinity using structural Moreover, such information is crucial for the development of
information and/or biochemical measurements, the use of molecu- new strategies of vector control capable of interfering in
lar docking and molecular dynamics simulations as virtual meth- olfaction-mediated behavioral responses. Prospecting for novel
ods to predict best ligands for insect OBPs is gaining eminence insect attractants or repellents can be based on molecular interac-
(Venthur et al., 2014), in an approach that has already been widely tions of candidate compounds with olfactory proteins.
used for drug design research (Okimoto et al., 2009). Docking stud- In this context, since no odorant comes to ORs except through
ies include the search for the most favorable conformation for the OBPs, functional OBPs can be utilized as molecular targets for the
protein-ligand complex in three-dimensional space and the score screening of insect behaviorally active compounds (attractants or
of resulting geometries is reflected in the binding energy. Thermo- repellents) in an approach similar to receptor-based drug discovery
dynamically, a ligand binds tightly to the active site of the protein named ‘‘reverse chemical ecology”. Besides, when compared to ORs,
when the free binding energy is low. Therefore, compounds with OBPs are more accessible targets for research, considering they are
lower free binding energy are predicted to be better ligands small, soluble, stable and easier to manipulate and modify. The
(Pagadala Damodaram et al., 2014). reverse chemical ecology approach narrows down the number of
Jiang et al. (2009) used molecular docking simulations to eval- odorant candidate compounds based on their binding affinity to
uate the strong binding affinity between LmigOBP1 and pentade- olfactory proteins, saving time and lowering research costs in rela-
canol, which suggested that Asn74 and is a key residue in the tion to the conventional trial-and-error screening performed in the
binding site of this particular OBP (Jiang et al., 2009). This result field.
was then validated by site-directed mutagenesis and fluorescence This strategy was successfully employed for the first time in the
assays. A recent study has provided a methodology for rapid development of better lures for the Navel Orangeworm moth, Amye-
screening of behaviorally active semiochemicals for the oriental lois transitella (Leal et al., 2005c). Later, using CquiOBP1 as a molec-
fruit fly Bactrocera dorsalis based on computational techniques. A ular target in binding assays, a study combining reverse and
general odorant-binding protein of the insect was analyzed for conventional chemical ecology approaches to identify unnatural
its binding potential against a panel of different semiochemicals. ligands has contributed to the development of the commercially
The compounds predicted to be best ligands were then subjected available oviposition attractant for C. quinquefasciatus (Leal et al.,
to behavioral bioassays and were found to be highly attractive to 2008).
insects (Pagadala Damodaram et al., 2014). Molecular docking The concept of reverse chemical ecology has been recently
studies performed between OBPs from the scarab beetle Hylamor- updated and computational methods are starting to be incorpo-
pha elegans and volatiles from Nothofagus obliqua revealed better rated to this approach in order to reduce even more the number
interaction energy values with sesquiterpenic compounds and of candidate compounds and accelerate the discovery of new
molecular dynamics reported that hydrophobic amino acids would semiochemicals (Fig. 5). The combination of molecular docking
take part of these relationships (Gonzalez-Gonzalez et al., 2016). and molecular dynamics was shown to be reliable for the

Fig. 5. Updated concept of the reverse chemical ecology approach for the prospection of behaviorally active ligands. The first screening happens in silico, where computational
methods are able to rapidly select compounds from extensive libraries. In the next step, the binding affinity of the selected compounds for olfactory proteins (in this case,
OBPs) is evaluated in vitro by biochemical techniques such as fluorescence assays or NMR. Finally, the final stage involves bioassays in olfactometers and field tests to
determine if the affinity observed in vitro is actually capable of evoking a behavioral response in vivo and, in the positive case, the nature of the observed effect (attraction or
repellency).
62 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

prediction of behaviorally active compounds for the agricultural Funding


pest B. dorsalis (Pagadala Damodaram et al., 2014). Thus, the very
first screening is now performed in silico, followed by the in vitro This study was supported by grants from the Conselho Nacional
assessment of binding affinities between OBPs and the preselected de Desenvolvimento Científico e Tecnológico (CNPq), Instituto
ligands by different techniques such as fluorescence, calorimetry, Nacional de Ciência e Tecnologia em Entomologia Molecular
NMR or even by measuring the amount of bound ligand, for (INCT-EM/CNPq), Fundação Carlos Chagas Filho de Amparo à Pes-
instance. Finally, insects’ behavioral response to the presence of quisa do Estado do Rio de Janeiro (FAPERJ) and Coordenação de
the best ligands is verified in vivo by bioassays and field tests. Aperfeiçoamento de Pessoal de Nível Superior (CAPES). No addi-
Even though their use as molecular targets for the development tional external funding was received for this study.
of new products for insect population management is the most
commonly discussed biotechnological application of insect OBPs, References
a few research groups have already discussed their potential for
Amenya, D.A., Chou, W., Li, J., Yan, G., Gershon, P.D., James, A.A., Marinotti, O., 2010.
other technological purposes. A few studies have used odorant-
Proteomics reveals novel components of the Anopheles gambiae eggshell. J.
binding proteins in biosensors for the detection of odorants. By Insect Physiol. 56, 1414–1419.
mimicking biological olfaction, these artificial devices can be used Benton, R., 2007. Sensitivity and specificity in Drosophila pheromone perception.
for the detection of important ligands in complex environments. Trends Neurosci. 30, 512–519.
Benton, R., Sachse, S., Michnick, S.W., Vosshall, L.B., 2006. Atypical membrane
Larisika et al. (2015) have designed an olfactory biosensor based topology and heteromeric function of Drosophila odorant receptors in vivo. PLoS
on a reduced graphene oxide transistor functionalized with the Biol. 4, e20.
OBP14 from the honeybee A. mellifera for the in situ and Benton, R., Vannice, K.S., Vosshall, L.B., 2007. An essential role for a CD36-related
receptor in pheromone detection in Drosophila. Nature 450, 289–293.
real-time monitoring of a broad spectrum of odorants in aqueous Benton, R., Vannice, K.S., Gomez-Diaz, C., Vosshall, L.B., 2009. Variant ionotropic
solutions known to be attractants for bees (Larisika et al., 2015). glutamate receptors as chemosensory receptors in Drosophila. Cell 136, 149–
Current oscillations are reproducible and seem to linearly vary 162.
Bette, S., Breer, H., Krieger, J., 2002. Probing a pheromone binding protein of the
with odorant concentration. Despite the different specificities silkmoth Antheraea polyphemus by endogenous tryptophan fluorescence. Insect
towards several organic compounds that make OBPs interesting Biochem. Mol. Biol. 32, 241–246.
tools for building devices for odor discrimination, not many studies Bianchet, M.A., Bains, G., Pelosi, P., Pevsner, J., Snyder, S.H., Monaco, H.L., Amzel, L.
M., 1996. The three-dimensional structure of bovine odorant binding protein
have used such proteins with this intent. and its mechanism of odor recognition. Nat. Struct. Biol. 3, 934–939.
Another interesting property of insect OBPs that can be exploited Biessmann, H., Nguyen, Q.K., Le, D., Walter, M.F., 2005. Microarray-based survey of a
for analytical purposes is their affinity and specificity for small subset of putative olfactory genes in the mosquito Anopheles gambiae. Insect
Mol. Biol. 14, 575–589.
organic molecules. Margaryan et al. (2006) reported the synthesis
Biessmann, H., Andronopoulou, E., Biessmann, M.R., Douris, V., Dimitratos, S.D.,
and characterization of affinity columns derivatised with PBP1 from Eliopoulos, E., Guerin, P.M., Iatrou, K., Justice, R.W., Krober, T., Marinotti, O.,
B. mori and OBP1 from C. quinquefasciatus (Margaryan et al., 2006). Tsitoura, P., Woods, D.F., Walter, M.F., 2010. The Anopheles gambiae odorant
Binding capacity was not affected by immobilization and OBPs also binding protein 1 (AgamOBP1) mediates indole recognition in the antennae of
female mosquitoes. PLoS One 5, e9471.
retained their pH-dependent ligand affinity, which can be used for Breer, H., 2003. Olfactory receptors: molecular basis for recognition and
gradual release of bound ligand from the column. Also, proteins discrimination of odors. Anal. Bioanal. Chem. 377, 427–433.
would be ideal substrates to differentiate enantiomers, but their Briand, L., Nespoulous, C., Huet, J.C., Pernollet, J.C., 2001a. Disulfide pairing and
secondary structure of ASP1, an olfactory-binding protein from honeybee (Apis
usually poor stability to temperature and solvents makes their mellifera L). J. Pept. Res. 58, 540–545.
application in analytical columns very difficult. However, it has Briand, L., Nespoulous, C., Huet, J.C., Takahashi, M., Pernollet, J.C., 2001b. Ligand
already been suggested that OBPs exceptional stability to thermal binding and physico-chemical properties of ASP2, a recombinant odorant-
binding protein from honeybee (Apis mellifera L.). Eur. J. Biochem. 268, 752–760.
denaturation and proteolytic degradation could provide the solution Buck, L., Axel, R., 1991. A novel multigene family may encode odorant receptors: a
for this matter (Pelosi et al., 2014). In fact, it is reasonable to imagine molecular basis for odor recognition. Cell 65, 175–187.
that, if properly employed as active chromatographic phases, OBPs Campanacci, V., Krieger, J., Bette, S., Sturgis, J.N., Lartigue, A., Cambillau, C., Breer, H.,
Tegoni, M., 2001. Revisiting the specificity of Mamestra brassicae and Antheraea
could, indeed, be directly used for resolving racemic mixtures of polyphemus pheromone-binding proteins with a fluorescence binding assay. J.
pheromones and/or other natural compounds. Biol. Chem. 276, 20078–20084.
Chang, H., Liu, Y., Yang, T., Pelosi, P., Dong, S., Wang, G., 2015. Pheromone binding
proteins enhance the sensitivity of olfactory receptors to sex pheromones in
10. Concluding remarks
Chilo suppressalis. Sci. Rep. 5, 13093.
Damberger, F.F., Ishida, Y., Leal, W.S., Wuthrich, K., 2007. Structural basis of ligand
Although the progress in the field of insect olfaction in the past binding and release in insect pheromone-binding proteins: NMR structure of
decade has been remarkable, further knowledge of insect olfactory Antheraea polyphemus PBP1 at pH 4.5. J. Mol. Biol. 373, 811–819.
Dani, F.R., Michelucci, E., Francese, S., Mastrobuoni, G., Cappellozza, S., La Marca, G.,
perception is still lacking in order to unravel the molecular mecha- Niccolini, A., Felicioli, A., Moneti, G., Pelosi, P., 2011. Odorant-binding proteins
nism whereby OBPs participate in chemoreception. How OBPs and chemosensory proteins in pheromone detection and release in the silkmoth
release odorants to ORs and whether they play a role in the selection Bombyx mori. Chem. Senses 36, 335–344.
Fan, J., Francis, F., Liu, Y., Chen, J.L., Cheng, D.F., 2011. An overview of odorant-
of these molecules are still a matter of debate. However, taking into binding protein functions in insect peripheral olfactory reception. Genet. Mol.
consideration the substantial diversity in insect species, it is not Res. 10, 3056–3069.
exactly surprising that available information regarding the roles Feng, L., Prestwich, G.D., 1997. Expression and characterization of a lepidopteran
general odorant binding protein. Insect Biochem. Mol. Biol. 27, 405–412.
and modes of action of olfactory proteins sometimes seems to go Forstner, M., Gohl, T., Gondesen, I., Raming, K., Breer, H., Krieger, J., 2008. Differential
in different directions. When it comes to the physiological function expression of SNMP-1 and SNMP-2 proteins in pheromone-sensitive hairs of
of OBPs, current understanding designates a role to these proteins as moths. Chem. Senses 33, 291–299.
Forstner, M., Breer, H., Krieger, J., 2009. A receptor and binding protein interplay in
carriers for odorants and pheromones across the sensillar lymph, the detection of a distinct pheromone component in the silkmoth Antheraea
keeping them protected from degradation. Such concept was first polyphemus. Int. J. Biol. Sci. 5, 745–757.
suggested about 30 years ago for OBPs of both vertebrates (Pelosi Galindo, K., Smith, D.P., 2001. A large family of divergent Drosophila odorant-
binding proteins expressed in gustatory and olfactory sensilla. Genetics 159,
and Tirindelli, 1989; Pevsner et al., 1990) and insects (Vogt et al.,
1059–1072.
1985) and remains the most widely accepted hypothesis. Neverthe- Gomez-Diaz, C., Reina, J.H., Cambillau, C., Benton, R., 2013. Ligands for pheromone-
less, odorant-binding proteins appear to be feasible molecular tar- sensing neurons are not conformationally activated odorant binding proteins.
gets to disrupt important behaviors in novel strategies for insect PLoS Biol. 11, e1001546.
Gong, D.P., Zhang, H.J., Zhao, P., Xia, Q.Y., Xiang, Z.H., 2009a. The odorant binding
population management, as well as other biotechnological protein gene family from the genome of silkworm, Bombyx mori. BMC Genomics
applications. 10, 332.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 63

Gong, Y., Pace, T.C., Castillo, C., Bohne, C., O’Neill, M.A., Plettner, E., 2009b. Ligand- Larisika, M., Kotlowski, C., Steininger, C., Mastrogiacomo, R., Pelosi, P., Schutz, S.,
interaction kinetics of the pheromone- binding protein from the gypsy moth, L. Peteu, S.F., Kleber, C., Reiner-Rozman, C., Nowak, C., Knoll, W., 2015. Electronic
dispar: insights into the mechanism of binding and release. Chem. Biol. 16, 162– olfactory sensor based on A. mellifera odorant-binding protein 14 on a reduced
172. graphene oxide field-effect transistor. Angew. Chem. Int. Ed. Engl. 54, 13245–
Gong, Y., Tang, H., Bohne, C., Plettner, E., 2009c. Binding conformation and kinetics 13248.
of two pheromone-binding proteins from the Gypsy moth Lymantria dispar with Larsson, M.C., Domingos, A.I., Jones, W.D., Chiappe, M.E., Amrein, H., Vosshall, L.B.,
biological and nonbiological ligands. Biochemistry 49, 793–801. 2004. Or83b encodes a broadly expressed odorant receptor essential for
Gonzalez, D., Zhao, Q., McMahan, C., Velasquez, D., Haskins, W.E., Sponsel, V., Cassill, Drosophila olfaction. Neuron 43, 703–714.
A., Renthal, R., 2009. The major antennal chemosensory protein of red imported Lartigue, A., Gruez, A., Spinelli, S., Riviere, S., Brossut, R., Tegoni, M., Cambillau, C.,
fire ant workers. Insect Mol. Biol. 18, 395–404. 2003. The crystal structure of a cockroach pheromone-binding protein suggests
Gonzalez-Gonzalez, A., Palma-Millanao, R., Yanez, O., Rojas, M., Mutis, A., Venthur, a new ligand binding and release mechanism. J. Biol. Chem. 278, 30213–30218.
H., Quiroz, A., Ramirez, C.C., 2016. Virtual screening of plant volatile compounds Lartigue, A., Gruez, A., Briand, L., Blon, F., Bezirard, V., Walsh, M., Pernollet, J.C.,
reveals a high affinity of Hylamorpha elegans (Coleoptera: Scarabaeidae) Tegoni, M., Cambillau, C., 2004. Sulfur single-wavelength anomalous diffraction
odorant-binding proteins for sesquiterpenes from its native host. J. Insect Sci. crystal structure of a pheromone-binding protein from the honeybee Apis
16. mellifera L. J. Biol. Chem. 279, 4459–4464.
Grater, F., Xu, W., Leal, W., Grubmuller, H., 2006. Pheromone discrimination by the Laughlin, J.D., Ha, T.S., Jones, D.N., Smith, D.P., 2008. Activation of pheromone-
pheromone-binding protein of Bombyx mori. Structure 14, 1577–1586. sensitive neurons is mediated by conformational activation of pheromone-
Grosse-Wilde, E., Svatos, A., Krieger, J., 2006. A pheromone-binding protein binding protein. Cell 133, 1255–1265.
mediates the bombykol-induced activation of a pheromone receptor in vitro. Lautenschlager, C., Leal, W.S., Clardy, J., 2007. Bombyx mori pheromone-binding
Chem. Senses 31, 547–555. protein binding nonpheromone ligands: implications for pheromone
Grosse-Wilde, E., Gohl, T., Bouche, E., Breer, H., Krieger, J., 2007. Candidate recognition. Structure 15, 1148–1154.
pheromone receptors provide the basis for the response of distinct antennal Leal, W.S., 2005. Pheromone reception. In: Schulz, S. (Ed.), The Chemistry of
neurons to pheromonal compounds. Eur. J. Neurosci. 25, 2364–2373. Pheromones and Other Semiochemicals II. Springer, Berlin Heidelberg, pp. 1–36.
Guidobaldi, F., May-Concha, I.J., Guerenstein, P.G., 2014. Morphology and Leal, W.S., 2006. Molecular-based chemical propecting of mosquito attractants and
physiology of the olfactory system of blood-feeding insects. J. Physiol. Paris repellents. In: Debboun, M.F., Strickman, D. (Eds.), Insect Repellents: Principles,
108, 96–111. Methods, and Uses. CRC Press, pp. 229–244.
Ha, T.S., Smith, D.P., 2006. A pheromone receptor mediates 11-cis-vaccenyl acetate- Leal, W.S., 2012. Odorant reception in insects: roles of receptors, binding proteins,
induced responses in Drosophila. J. Neurosci. 26, 8727–8733. and degrading enzymes. Annu. Rev. Entomol. 58, 373–391.
He, X., Tzotzos, G., Woodcock, C., Pickett, J.A., Hooper, T., Field, L.M., Zhou, J.J., 2010. Leal, W.S., 2013. Odorant reception in insects: roles of receptors, binding proteins,
Binding of the general odorant binding protein of Bombyx mori BmorGOBP2 to and degrading enzymes. Annu. Rev. Entomol. 58, 373–391.
the moth sex pheromone components. J. Chem. Ecol. 36, 1293–1305. Leal, W.S., Kaissling, K.-E., 2004. Biologische Nanokapseln fur Duftstoffe. Naturwiss.
Hekmat-Scafe, D.S., Scafe, C.R., McKinney, A.J., Tanouye, M.A., 2002. Genome-wide Rundschau 668, 66–71.
analysis of the odorant-binding protein gene family in Drosophila melanogaster. Leal, W.S., Nikonova, L., Peng, G., 1999. Disulfide structure of the pheromone
Genome Res. 12, 1357–1369. binding protein from the silkworm moth, Bombyx mori. FEBS Lett. 464, 85–90.
Horst, R., Damberger, F., Luginbuhl, P., Guntert, P., Peng, G., Nikonova, L., Leal, W.S., Leal, W.S., Chen, A.M., Erickson, M.L., 2005a. Selective and pH-dependent binding of
Wuthrich, K., 2001. NMR structure reveals intramolecular regulation a moth pheromone to a pheromone-binding protein. J. Chem. Ecol. 31, 2493–
mechanism for pheromone binding and release. Proc. Natl. Acad. Sci. U.S.A. 2499.
98, 14374–14379. Leal, W.S., Chen, A.M., Ishida, Y., Chiang, V.P., Erickson, M.L., Morgan, T.I., Tsuruda, J.
Ishida, Y., Leal, W.S., 2005. Rapid inactivation of a moth pheromone. Proc. Natl. M., 2005b. Kinetics and molecular properties of pheromone binding and release.
Acad. Sci. U.S.A. 102, 14075–14079. Proc. Natl. Acad. Sci. U.S.A. 102, 5386–5391.
Ishida, Y., Leal, W.S., 2008. Chiral discrimination of the Japanese beetle sex Leal, W.S., Parra-Pedrazzoli, A.L., Kaissling, K.E., Morgan, T.I., Zalom, F.G., Pesak, D.J.,
pheromone and a behavioral antagonist by a pheromone-degrading enzyme. Dundulis, E.A., Burks, C.S., Higbee, B.S., 2005c. Unusual pheromone chemistry in
Proc. Natl. Acad. Sci. U.S.A. 105, 9076–9080. the navel orangeworm: novel sex attractants and a behavioral antagonist.
Ishida, Y., Ishibashi, J., Leal, W.S., 2013. Fatty acid solubilizer from the oral disk of Naturwissenschaften 92, 139–146.
the blowfly. PLoS One 8, e51779. Leal, W.S., Barbosa, R.M., Xu, W., Ishida, Y., Syed, Z., Latte, N., Chen, A.M., Morgan, T.
Jacquin-Joly, E., Vogt, R.G., Francois, M.C., Nagnan-Le Meillour, P., 2001. Functional I., Cornel, A.J., Furtado, A., 2008. Reverse and conventional chemical ecology
and expression pattern analysis of chemosensory proteins expressed in approaches for the development of oviposition attractants for Culex mosquitoes.
antennae and pheromonal gland of Mamestra brassicae. Chem. Senses 26, PLoS One 3, e3045.
833–844. Lee, D., Damberger, F.F., Peng, G., Horst, R., Guntert, P., Nikonova, L., Leal, W.S.,
Jiang, Q.Y., Wang, W.X., Zhang, Z., Zhang, L., 2009. Binding specificity of locust Wuthrich, K., 2002. NMR structure of the unliganded Bombyx mori pheromone-
odorant binding protein and its key binding site for initial recognition of binding protein at physiological pH. FEBS Lett. 531, 314–318.
alcohols. Insect Biochem. Mol. Biol. 39, 440–447. Leite, N.R., Krogh, R., Xu, W., Ishida, Y., Iulek, J., Leal, W.S., Oliva, G., 2009. Structure
Jin, X., Ha, T.S., Smith, D.P., 2008. SNMP is a signaling component required for of an odorant-binding protein from the mosquito Aedes aegypti suggests a
pheromone sensitivity in Drosophila. Proc. Natl. Acad. Sci. U.S.A. 105, 10996– binding pocket covered by a pH-sensitive ‘‘Lid”. PLoS One 4, e8006.
11001. Lescop, E., Briand, L., Pernollet, J.C., Van Heijenoort, C., Guittet, E., 2001. 1H, 13C and
Jin, J.Y., Li, Z.Q., Zhang, Y.N., Liu, N.Y., Dong, S.L., 2014. Different roles suggested by 15N chemical shift assignment of the honeybee odorant-binding protein ASP2.
sex-biased expression and pheromone binding affinity among three pheromone J. Biomol. NMR 21, 181–182.
binding proteins in the pink rice borer, Sesamia inferens (Walker) (Lepidoptera: Li, Z.X., Pickett, J.A., Field, L.M., Zhou, J.J., 2005. Identification and expression of
Noctuidae). J. Insect Physiol. 66, 71–79. odorant-binding proteins of the malaria-carrying mosquitoes Anopheles
Kaissling, K.E., 2009. Olfactory perireceptor and receptor events in moths: a kinetic gambiae and Anopheles arabiensis. Arch. Insect Biochem. Physiol. 58, 175–189.
model revised. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 195, Li, S., Picimbon, J.F., Ji, S., Kan, Y., Chuanling, Q., Zhou, J.J., Pelosi, P., 2008. Multiple
895–922. functions of an odorant-binding protein in the mosquito Aedes aegypti. Biochem.
Kim, M.S., Repp, A., Smith, D.P., 1998. LUSH odorant-binding protein mediates Biophys. Res. Commun. 372, 464–468.
chemosensory responses to alcohols in Drosophila melanogaster. Genetics 150, Li, G., Chen, X., Li, B., Zhang, G., Li, Y., Wu, J., 2016. Binding properties of general
711–721. odorant binding proteins from the oriental fruit moth, Grapholita molesta
Kitabayashi, A.N., Arai, T., Kubo, T., Natori, S., 1998. Molecular cloning of cDNA for (Busck) (Lepidoptera: Tortricidae). PLoS One 11, e0155096.
p10, a novel protein that increases in the regenerating legs of Periplaneta Maida, R., Proebstl, T., Laue, M., 1997. Heterogeneity of odorant-binding proteins in
americana (American cockroach). Insect Biochem. Mol. Biol. 28, 785–790. the antennae of Bombyx mori. Chem. Senses 22, 503–515.
Klein, U., 1987. Sensillum-lymph proteins from antennal olfactory hairs of the moth Maida, R., Ziegelberger, G., Kaissling, K.E., 2003. Ligand binding to six recombinant
Antheraea polyphemus (Saturniidae). Insect Biochem. 17, 1193–1204. pheromone-binding proteins of Antheraea polyphemus and Antheraea pernyi. J.
Krieger, J., von Nickisch-Rosenegk, E., Mameli, M., Pelosi, P., Breer, H., 1996. Binding Comp. Physiol. B 173, 565–573.
proteins from the antennae of Bombyx mori. Insect Biochem. Mol. Biol. 26, 297– Maleszka, J., Foret, S., Saint, R., Maleszka, R., 2007. RNAi-induced phenotypes
307. suggest a novel role for a chemosensory protein CSP5 in the development of
Kruse, S.W., Zhao, R., Smith, D.P., Jones, D.N., 2003. Structure of a specific alcohol- embryonic integument in the honeybee (Apis mellifera). Dev. Genes Evol. 217,
binding site defined by the odorant binding protein LUSH from Drosophila 189–196.
melanogaster. Nat. Struct. Biol. 10, 694–700. Manoharan, M., Ng Fuk Chong, M., Vaitinadapoule, A., Vaitinadapoule, A., Frumence,
Kurtovic, A., Widmer, A., Dickson, B.J., 2007. A single class of olfactory neurons E., Sowdhamini, R., Offmann, B., 2013. Comparative genomics of odorant
mediates behavioural responses to a Drosophila sex pheromone. Nature 446, binding proteins in Anopheles gambiae, Aedes aegypti, and Culex quinquefasciatus.
542–546. Genome Biol. Evol. 5, 163–180.
Lagarde, A., Spinelli, S., Qiao, H., Tegoni, M., Pelosi, P., Cambillau, C., 2011a. Crystal Mao, Y., Xu, X., Xu, W., Ishida, Y., Leal, W.S., Ames, J.B., Clardy, J., 2010. Crystal and
structure of a novel type of odorant binding protein from Anopheles gambiae, solution structures of an odorant-binding protein from the southern house
belonging to the C+ class. Biochem. J. mosquito complexed with an oviposition pheromone. Proc. Natl. Acad. Sci. U.S.
Lagarde, A., Spinelli, S., Tegoni, M., He, X., Field, L., Zhou, J.J., Cambillau, C., 2011b. A. 107, 19102–19107.
The crystal structure of odorant binding protein 7 from Anopheles gambiae Margaryan, A., Moaddel, R., Aldrich, J.R., Tsuruda, J.M., Chen, A.M., Leal, W.S.,
exhibits an outstanding adaptability of its binding site. J. Mol. Biol. 414, 401– Wainer, I.W., 2006. Synthesis of an immobilized Bombyx mori pheromone-
412. binding protein liquid chromatography stationary phase. Talanta 70, 752–755.
64 N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65

Mastrobuoni, G., Qiao, H., Iovinella, I., Sagona, S., Niccolini, A., Boscaro, F., Caputo, B., Pevsner, J., Hou, V., Snowman, A.M., Snyder, S.H., 1990. Odorant-binding protein.
Orejuela, M.R., Della Torre, A., Kempa, S., Felicioli, A., Pelosi, P., Moneti, G., Dani, Characterization of ligand binding. J. Biol. Chem. 265, 6118–6125.
F.R., 2013. A proteomic investigation of soluble olfactory proteins in Anopheles Pitts, R.J., Rinker, D.C., Jones, P.L., Rokas, A., Zwiebel, L.J., 2011. Transcriptome
gambiae. PLoS One 8, e75162. profiling of chemosensory appendages in the malaria vector Anopheles gambiae
Matsuo, T., Sugaya, S., Yasukawa, J., Aigaki, T., Fuyama, Y., 2007. Odorant-binding reveals tissue- and sex-specific signatures of odor coding. BMC Genomics 12,
proteins OBP57d and OBP57e affect taste perception and host-plant preference 271.
in Drosophila sechellia. PLoS Biol. 5, e118. Plettner, E., Lazar, J., Prestwich, E.G., Prestwich, G.D., 2000. Discrimination of
Mesquita, R.D., Vionette-Amaral, R.J., Lowenberger, C., Rivera-Pomar, R., Monteiro, F. pheromone enantiomers by two pheromone binding proteins from the gypsy
A., Minx, P., Spieth, J., Carvalho, A.B., Panzera, F., Lawson, D., Torres, A.Q., Ribeiro, moth Lymantria dispar. Biochemistry 39, 8953–8962.
J.M., Sorgine, M.H., Waterhouse, R.M., Montague, M.J., Abad-Franch, F., Alves- Prestwich, G.D., 1993. Bacterial expression and photoaffinity labeling of a
Bezerra, M., Amaral, L.R., Araujo, H.M., Araujo, R.N., Aravind, L., Atella, G.C., pheromone binding protein. Protein Sci. 2, 420–428.
Azambuja, P., Berni, M., Bittencourt-Cunha, P.R., Braz, G.R., Calderon-Fernandez, Qiao, H., Tuccori, E., He, X., Gazzano, A., Field, L., Zhou, J.J., Pelosi, P., 2009.
G., Carareto, C.M., Christensen, M.B., Costa, I.R., Costa, S.G., Dansa, M., Daumas- Discrimination of alarm pheromone (E)-beta-farnesene by aphid odorant-
Filho, C.R., De-Paula, I.F., Dias, F.A., Dimopoulos, G., Emrich, S.J., Esponda- binding proteins. Insect Biochem. Mol. Biol. 39, 414–419.
Behrens, N., Fampa, P., Fernandez-Medina, R.D., da Fonseca, R.N., Fontenele, M., Ribeiro, J.M., Genta, F.A., Sorgine, M.H., Logullo, R., Mesquita, R.D., Paiva-Silva, G.
Fronick, C., Fulton, L.A., Gandara, A.C., Garcia, E.S., Genta, F.A., Giraldo-Calderon, O., Majerowicz, D., Medeiros, M., Koerich, L., Terra, W.R., Ferreira, C.,
G.I., Gomes, B., Gondim, K.C., Granzotto, A., Guarneri, A.A., Guigo, R., Harry, M., Pimentel, A.C., Bisch, P.M., Leite, D.C., Diniz, M.M., Da S.G.V.Junior, J.L., Da
Hughes, D.S., Jablonka, W., Jacquin-Joly, E., Juarez, M.P., Koerich, L.B., Latorre- Silva, M.L., Araujo, R.N., Gandara, A.C., Brosson, S., Salmon, D., Bousbata, S.,
Estivalis, J.M., Lavore, A., Lawrence, G.G., Lazoski, C., Lazzari, C.R., Lopes, R.R., Gonzalez-Caballero, N., Silber, A.M., Alves-Bezerra, M., Gondim, K.C., Silva-
Lorenzo, M.G., Lugon, M.D., Majerowicz, D., Marcet, P.L., Mariotti, M., Masuda, Neto, M.A., Atella, G.C., Araujo, H., Dias, F.A., Polycarpo, C., Vionette-Amaral,
H., Megy, K., Melo, A.C., Missirlis, F., Mota, T., Noriega, F.G., Nouzova, M., Nunes, R.J., Fampa, P., Melo, A.C., Tanaka, A.S., Balczun, C., Oliveira, J.H., Goncalves, R.
R.D., Oliveira, R.L., Oliveira-Silveira, G., Ons, S., Pagola, L., Paiva-Silva, G.O., L., Lazoski, C., Rivera-Pomar, R., Diambra, L., Schaub, G.A., Garcia, E.S.,
Pascual, A., Pavan, M.G., Pedrini, N., Peixoto, A.A., Pereira, M.H., Pike, A., Azambuja, P., Braz, G.R., Oliveira, P.L., 2014. An insight into the transcriptome
Polycarpo, C., Prosdocimi, F., Ribeiro-Rodrigues, R., Robertson, H.M., Salerno, A. of the digestive tract of the bloodsucking bug, Rhodnius prolixus. PLoS Negl.
P., Salmon, D., Santesmasses, D., Schama, R., Seabra-Junior, E.S., Silva-Cardoso, Trop. Dis. 8, e2594.
L., Silva-Neto, M.A., Souza-Gomes, M., Sterkel, M., Taracena, M.L., Tojo, M., Tu, Z. Riviere, S., Lartigue, A., Quennedey, B., Campanacci, V., Farine, J.P., Tegoni, M.,
J., Tubio, J.M., Ursic-Bedoya, R., Venancio, T.M., Walter-Nuno, A.B., Wilson, D., Cambillau, C., Brossut, R., 2003. A pheromone-binding protein from the
Warren, W.C., Wilson, R.K., Huebner, E., Dotson, E.M., Oliveira, P.L., 2015. cockroach Leucophaea maderae: cloning, expression and pheromone binding.
Genome of Rhodnius prolixus, an insect vector of Chagas disease, reveals unique Biochem. J. 371, 573–579.
adaptations to hematophagy and parasite infection. Proc. Natl. Acad. Sci. U.S.A. Rogers, M.E., Krieger, J., Vogt, R.G., 2001. Antennal SNMPs (sensory neuron
112, 14936–14941. membrane proteins) of Lepidoptera define a unique family of invertebrate
Mohanty, S., Zubkov, S., Gronenborn, A.M., 2004. The solution NMR structure of CD36-like proteins. J. Neurobiol. 49, 47–61.
Antheraea polyphemus PBP provides new insight into pheromone recognition by Ronderos, D.S., Smith, D.P., 2010. Activation of the T1 neuronal circuit is necessary
pheromone-binding proteins. J. Mol. Biol. 337, 443–451. and sufficient to induce sexually dimorphic mating behavior in Drosophila
Murphy, E.J., Booth, J.C., Davrazou, F., Port, A.M., Jones, D.N.M., 2013. Interactions of melanogaster. J. Neurosci. 30, 2595–2599.
Anopheles gambiae odorant-binding proteins with a human-derived repellent: Rytz, R., Croset, V., Benton, R., 2013. Ionotropic Receptors (IRs): chemosensory
implications for the mode of action of n, n-diethyl-3-methylbenzamide (DEET). ionotropic glutamate receptors in Drosophila and beyond. Insect Biochem. Mol.
JBC 288 (6), 4475–4485. Biol. 43, 888–897.
Nakagawa, T., Sakurai, T., Nishioka, T., Touhara, K., 2005. Insect sex-pheromone Sanchez-Gracia, A., Vieira, F.G., Rozas, J., 2009. Molecular evolution of the major
signals mediated by specific combinations of olfactory receptors. Science 307, chemosensory gene families in insects. Heredity 103, 208–216.
1638–1642. Sandler, B.H., Nikonova, L., Leal, W.S., Clardy, J., 2000. Sexual attraction in the
Nichols, Z., Vogt, R.G., 2008. The SNMP/CD36 gene family in Diptera, Hymenoptera silkworm moth: structure of the pheromone-binding-protein-bombykol
and Coleoptera: Drosophila melanogaster, D. pseudoobscura, Anopheles gambiae, complex. Chem. Biol. 7, 143–151.
Aedes aegypti, Apis mellifera, and Tribolium castaneum. Insect Biochem. Mol. Biol. Sato, K., Pellegrino, M., Nakagawa, T., Vosshall, L.B., Touhara, K., 2008. Insect
38, 398–415. olfactory receptors are heteromeric ligand-gated ion channels. Nature 452,
Nomura, A., Kawasaki, K., Kubo, T., Natori, S., 1992. Purification and localization of 1002–1006.
p10, a novel protein that increases in nymphal regenerating legs of Periplaneta Shanbhag, S.R., Hekmat-Scafe, D., Kim, M.S., Park, S.K., Carlson, J.R., Pikielny, C.,
americana (American cockroach). Int. J. Dev. Biol. 36, 391–398. Smith, D.P., Steinbrecht, R.A., 2001. Expression mosaic of odorant-binding
Northey, T., Venthur, H., De Biasio, F., Chauviac, F.X., Cole, A., Ribeiro, K.A.J., Grossi, proteins in Drosophila olfactory organs. Microsc. Res. Tech. 55, 297–306.
G., Falabella, P., Field, L.M., Keep, N.H., Zhou, J.J., 2016. Crystal structures and Siciliano, P., He, X.L., Woodcock, C., Pickett, J.A., Field, L.M., Birkett, M.A., Kalinova, B.,
binding dynamics of odorant-binding protein 3 from two aphid species Megoura Gomulski, L.M., Scolari, F., Gasperi, G., Malacrida, A.R., Zhou, J.J., 2014.
viciae and Nasonovia ribisnigri. Sci. Rep. 6, 24739. Identification of pheromone components and their binding affinity to the
Okimoto, N., Futatsugi, N., Fuji, H., Suenaga, A., Morimoto, G., Yanai, R., Ohno, Y., odorant binding protein CcapOBP83a-2 of the Mediterranean fruit fly, Ceratitis
Narumi, T., Taiji, M., 2009. High-performance drug discovery: computational capitata. Insect Biochem. Mol. Biol. 48, 51–62.
screening by combining docking and molecular dynamics simulations. PLoS Smartt, C.T., Erickson, J.S., 2009. Expression of a novel member of the odorant-
Comput. Biol. 5, e1000528. binding protein gene family in Culex nigripalpus (Diptera: Culicidae). J. Med.
Ozaki, M., Morisaki, K., Idei, W., Ozaki, K., Tokunaga, F., 1995. A putative lipophilic Entomol. 46, 1376–1381.
stimulant carrier protein commonly found in the taste and olfactory systems. A Smith, D.P., 2007. Odor and pheromone detection in Drosophila melanogaster.
unique member of the pheromone-binding protein superfamily. Eur. J. Biochem. Pflugers Arch. 454, 749–758.
230, 298–308. Song, Y.-Q., Dong, J.-F., Qiao, H.-L., WU, J.-X., 2014. Molecular characterization,
Ozaki, K., Utoguchi, A., Yamada, A., Yoshikawa, H., 2008. Identification and genomic expression patterns and binding properties of two pheromone-binding proteins
structure of chemosensory proteins (CSP) and odorant binding proteins (OBP) from the oriental fruit moth, Grapholita molesta (Busck). J. Integr. Agric. 13,
genes expressed in foreleg tarsi of the swallowtail butterfly Papilio xuthus. 2709–2720.
Insect Biochem. Mol. Biol. 38, 969–976. Spinelli, S., Lagarde, A., Iovinella, I., Legrand, P., Tegoni, M., Pelosi, P., Cambillau, C.,
Pagadala Damodaram, K.J., Kempraj, V., Aurade, R.M., Tapas, K.R., Shivashankara, K. 2012. Crystal structure of Apis mellifera OBP14, a C-minus odorant-binding
S., Verghese, A., 2014. Computational reverse chemical ecology: virtual protein, and its complexes with odorant molecules. Insect Biochem. Mol. Biol.
screening and predicting behaviorally active semiochemicals for Bactrocera 42, 41–50.
dorsalis. BMC Genomics 15 (209), 1–7. Sun, Y.F., De Biasio, F., Qiao, H.L., Iovinella, I., Yang, S.X., Ling, Y., Riviello, L., Battaglia,
Pelletier, J., Hughes, D.T., Luetje, C.W., Leal, W.S., 2010. An odorant receptor from the D., Falabella, P., Yang, X.L., Pelosi, P., 2012a. Two odorant-binding proteins
southern house mosquito Culex pipiens quinquefasciatus sensitive to oviposition mediate the behavioural response of aphids to the alarm pheromone (E)-ss-
attractants. PLoS One 5, e10090. farnesene and structural analogues. PLoS One 7, e32759.
Pelosi, P., Tirindelli, R., 1989. Structure/activity studies and characterization of an Sun, Y.L., Huang, L.Q., Pelosi, P., Wang, C.Z., 2012b. Expression in antennae and
odorant-binding protein. In: Brand, J.G., Cagan, T.J., Kare, R.H. (Eds.), Chemical reproductive organs suggests a dual role of an odorant-binding protein in two
senses: receptor events and transduction in taste and olfaction. Marcel Dekker, sibling Helicoverpa species. PLoS One 7, e30040.
New York, pp. 207–226. Sun, L., Wei, Y., Zhang, D.D., Ma, X.Y., Xiao, Y., Zhang, Y.N., Yang, X.M., Xiao, Q., Guo,
Pelosi, P., Zhou, J.J., Ban, L.P., Calvello, M., 2006. Soluble proteins in insect chemical Y.Y., Zhang, Y.J., 2016. The mouthparts enriched odorant binding protein 11 of
communication. Cell. Mol. Life Sci. 63, 1658–1676. the alfalfa plant bug Adelphocoris lineolatus displays a preferential binding
Pelosi, P., Mastrogiacomo, R., Iovinella, I., Tuccori, E., Persaud, K.C., 2014. Structure behavior to host plant secondary metabolites. Front Physiol 7, 201.
and biotechnological applications of odorant-binding proteins. Appl. Microbiol. Swarup, S., Williams, T.I., Anholt, R.R., 2011. Functional dissection of odorant
Biotechnol. 98, 61–70. binding protein genes in Drosophila melanogaster. Genes Brain Behav. 10, 648–
Pesenti, M.E., Spinelli, S., Bezirard, V., Briand, L., Pernollet, J.C., Tegoni, M., Cambillau, 657.
C., 2008. Structural basis of the honey bee PBP pheromone and pH-induced Syed, Z., Kopp, A., Kimbrell, D.A., Leal, W.S., 2010. Bombykol receptors in the
conformational change. J. Mol. Biol. 380, 158–169. silkworm moth and the fruit fly. Proc. Natl. Acad. Sci. U.S.A. 107, 9436–9439.
Pesenti, M.E., Spinelli, S., Bezirard, V., Briand, L., Pernollet, J.C., Campanacci, V., Tegoni, M., Campanacci, V., Cambillau, C., 2004. Structural aspects of sexual
Tegoni, M., Cambillau, C., 2009. Queen bee pheromone binding protein pH- attraction and chemical communication in insects. Trends Biochem. Sci. 29,
induced domain swapping favors pheromone release. J. Mol. Biol. 390, 981–990. 257–264.
N.F. Brito et al. / Journal of Insect Physiology 95 (2016) 51–65 65

Thode, A.B., Kruse, S.W., Nix, J.C., Jones, D.N., 2008. The role of multiple hydrogen- Xu, W., Cornel, A.J., Leal, W.S., 2010a. Odorant-binding proteins of the malaria
bonding groups in specific alcohol binding sites in proteins: insights from mosquito Anopheles funestus sensu stricto. PLoS One 5, e15403.
structural studies of LUSH. J. Mol. Biol. 376, 1360–1376. Xu, W., Xu, X., Leal, W.S., Ames, J.B., 2010b. Extrusion of the C-terminal helix in
Tsitsanou, K.E., Thireou, T., Drakou, C.E., Koussis, K., Keramioti, M.V., Leonidas, D.D., navel orangeworm moth pheromone-binding protein (AtraPBP1) controls
Eliopoulos, E., Iatrou, K., Zographos, S.E., 2012. Anopheles gambiae odorant pheromone binding. Biochem. Biophys. Res. Commun. 404, 335–338.
binding protein crystal complex with the synthetic repellent DEET: implications Xu, X., Xu, W., Rayo, J., Ishida, Y., Leal, W.S., Ames, J.B., 2010c. NMR structure of navel
for structure-based design of novel mosquito repellents. Cell. Mol. Life Sci. 69, orangeworm moth pheromone-binding protein (AtraPBP1): implications for
283–297. pH-sensitive pheromone detection. Biochemistry 49, 1469–1476.
Tsitsanou, K.E., Drakou, C.E., Thireou, T., Vitlin Gruber, A., Kythreoti, G., Azem, A., Yin, J., Feng, H., Sun, H., Xi, J., Cao, Y., Li, K., 2012. Functional analysis of general
Fessas, D., Eliopoulos, E., Iatrou, K., Zographos, S.E., 2013. Crystal and solution odorant binding protein 2 from the meadow moth, Loxostege sticticalis L.
studies of the ‘‘Plus-C” odorant-binding protein 48 from Anopheles gambiae: (Lepidoptera: Pyralidae). PLoS One 7, e33589.
control of binding specificity through three-dimensional domain swapping. J. Yin, J., Choo, Y.M., Duan, H., Leal, W.S., 2015. Selectivity of odorant-binding proteins
Biol. Chem. 288, 33427–33438. from the southern house mosquito tested against physiologically relevant
Venthur, H., Mutis, A., Zhou, J.-J., Quiroz, A., 2014. Ligand binding and homology ligands. Front Physiol. 6, 56.
modelling of insect odorant-binding proteins. Physiol. Entomol. 39, 183–198. Zheng, J., Li, J., Han, L., Wang, Y., Wu, W., Qi, X., Tao, Y., Zhang, L., Zhang, Z., Chen, Z.,
Vieira, F.G., Rozas, J., 2011. Comparative genomics of the odorant-binding and 2015. Crystal structure of the Locusta migratoria odorant binding protein.
chemosensory protein gene families across the Arthropoda: origin and Biochem. Biophys. Res. Commun. 456, 737–742.
evolutionary history of the chemosensory system. Genome Biol. Evol. 3, 476– Zhou, J.J., 2010. Odorant-binding proteins in insects. Vitam. Horm. 83, 241–272.
490. Zhou, J.J., Huang, W., Zhang, G.A., Pickett, J.A., Field, L.M., 2004a. ‘‘Plus-C” odorant-
Vieira, F.G., Sanchez-Gracia, A., Rozas, J., 2007. Comparative genomic analysis of the binding protein genes in two Drosophila species and the malaria mosquito
odorant-binding protein family in 12 Drosophila genomes: purifying selection Anopheles gambiae. Gene 327, 117–129.
and birth-and-death evolution. Genome Biol. 8, R235. Zhou, J.J., Zhang, G.A., Huang, W., Birkett, M.A., Field, L.M., Pickett, J.A., Pelosi, P.,
Vogt, R.G., Riddiford, L.M., 1981. Pheromone binding and inactivation by moth 2004b. Revisiting the odorant-binding protein LUSH of Drosophila melanogaster:
antennae. Nature 293, 161–163. evidence for odour recognition and discrimination. FEBS Lett. 558, 23–26.
Vogt, R.G., Riddiford, L.M., Prestwich, G.D., 1985. Kinetic properties of a sex Zhou, J.J., He, X.L., Pickett, J.A., Field, L.M., 2008. Identification of odorant-binding
pheromone-degrading enzyme: the sensillar esterase of Antheraea polyphemus. proteins of the yellow fever mosquito Aedes aegypti: genome annotation and
Proc. Natl. Acad. Sci. U.S.A. 82, 8827–8831. comparative analyses. Insect Mol. Biol. 17, 147–163.
Vosshall, L.B., Hansson, B.S., 2011. A unified nomenclature system for the insect Zhou, J.J., Robertson, G., He, X., Dufour, S., Hooper, A.M., Pickett, J.A., Keep, N.H.,
olfactory coreceptor. Chem. Senses 36, 497–498. Field, L.M., 2009. Characterisation of Bombyx mori Odorant-binding proteins
Wicher, D., Schafer, R., Bauernfeind, R., Stensmyr, M.C., Heller, R., Heinemann, S.H., reveals that a general odorant-binding protein discriminates between sex
Hansson, B.S., 2009. DOr83b–receptor or ion channel? Ann. N. Y. Acad. Sci. 1170, pheromone components. J. Mol. Biol. 389, 529–545.
164–167. Zhou, X.H., Ban, L.P., Iovinella, I., Zhao, L.J., Gao, Q., Felicioli, A., Sagona, S., Pieraccini,
Wogulis, M., Morgan, T., Ishida, Y., Leal, W.S., Wilson, D.K., 2006. The crystal G., Pelosi, P., Zhang, L., Dani, F.R., 2013. Diversity, abundance, and sex-specific
structure of an odorant binding protein from Anopheles gambiae: evidence for a expression of chemosensory proteins in the reproductive organs of the locust
common ligand release mechanism. Biochem. Biophys. Res. Commun. 339, Locusta migratoria manilensis. Biol. Chem. 394, 43–54.
157–164. Zhuang, Y.-D., Chiang, P.-Y., Wang, C.-W., Tan, K.-T., 2013. Environment-sensitive
Wojtasek, H., Leal, W.S., 1999. Conformational change in the pheromone-binding fluorescent turn-on probes targeting hydrophobic ligand-binding domains for
protein from Bombyx mori induced by pH and by interaction with membranes. J. selective protein detection. Angew. Chem. Int. Ed. 52, 8124–8128.
Biol. Chem. 274, 30950–30956. Ziemba, B.P., Murphy, E.J., Edlin, H.T., Jones, D.N., 2013. A novel mechanism of ligand
Xu, W., Leal, W.S., 2008. Molecular switches for pheromone release from a moth binding and release in the odorant binding protein 20 from the malaria
pheromone-binding protein. Biochem. Biophys. Res. Commun. 372, 559–564. mosquito Anopheles gambiae. Protein Sci. 22, 11–21.
Xu, P.X., Zwiebel, L.J., Smith, D.P., 2003. Identification of a distinct family of genes Zubkov, S., Gronenborn, A.M., Byeon, I.J., Mohanty, S., 2005. Structural consequences
encoding atypical odorant-binding proteins in the malaria vector mosquito, of the pH-induced conformational switch in A. polyphemus pheromone-binding
Anopheles gambiae. Insect Mol. Biol. 12, 549–560. protein: mechanisms of ligand release. J. Mol. Biol. 354, 1081–1090.
Xu, P., Atkinson, R., Jones, D.N., Smith, D.P., 2005. Drosophila OBP LUSH is required Zwiebel, L.J., Takken, W., 2004. Olfactory regulation of mosquito-host interactions.
for activity of pheromone-sensitive neurons. Neuron 45, 193–200. Insect Biochem. Mol. Biol. 34, 645–652.

You might also like