18.06.2013 Views

Mechanisms of Olfaction in Insects - ResearchSpace@Auckland ...

Mechanisms of Olfaction in Insects - ResearchSpace@Auckland ...

Mechanisms of Olfaction in Insects - ResearchSpace@Auckland ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

http://researchspace.auckland.ac.nz<br />

<strong>ResearchSpace@Auckland</strong><br />

Copyright Statement<br />

The digital copy <strong>of</strong> this thesis is protected by the Copyright Act 1994 (New<br />

Zealand).<br />

This thesis may be consulted by you, provided you comply with the<br />

provisions <strong>of</strong> the Act and the follow<strong>in</strong>g conditions <strong>of</strong> use:<br />

Any use you make <strong>of</strong> these documents or images must be for<br />

research or private study purposes only, and you may not make<br />

them available to any other person.<br />

Authors control the copyright <strong>of</strong> their thesis. You will recognise the<br />

author's right to be identified as the author <strong>of</strong> this thesis, and due<br />

acknowledgement will be made to the author where appropriate.<br />

You will obta<strong>in</strong> the author's permission before publish<strong>in</strong>g any<br />

material from their thesis.<br />

To request permissions please use the Feedback form on our webpage.<br />

http://researchspace.auckland.ac.nz/feedback<br />

General copyright and disclaimer<br />

In addition to the above conditions, authors give their consent for the<br />

digital copy <strong>of</strong> their work to be used subject to the conditions specified on<br />

the Library Thesis Consent Form and Deposit Licence.<br />

Note : Masters Theses<br />

The digital copy <strong>of</strong> a masters thesis is as submitted for exam<strong>in</strong>ation and<br />

conta<strong>in</strong>s no corrections. The pr<strong>in</strong>t copy, usually available <strong>in</strong> the University<br />

Library, may conta<strong>in</strong> corrections made by hand, which have been<br />

requested by the supervisor.


The School <strong>of</strong> Biological Sciences<br />

The University <strong>of</strong> Auckland<br />

New Zealand<br />

Identification and<br />

Analysis <strong>of</strong> Olfactory<br />

Receptors from the Light<br />

Brown Apple Moth,<br />

(Epiphyas postvittana)<br />

Doreen Shabana Begum<br />

March 2011<br />

Supervisors: Richard D. Newcomb<br />

Andrew V. Kralicek<br />

David L. Christie<br />

A thesis submitted <strong>in</strong> fulfilment <strong>of</strong> the requirements for the degree <strong>of</strong> Doctor <strong>of</strong><br />

Philosophy <strong>in</strong> Biological Sciences, The University <strong>of</strong> Auckland, 2011.


Abstract<br />

Abstract<br />

<strong>Olfaction</strong> or the sense <strong>of</strong> smell is one <strong>of</strong> the major modes <strong>of</strong> communication <strong>in</strong> moths,<br />

play<strong>in</strong>g an <strong>in</strong>tegral part <strong>in</strong> the moths‟ ability to locate mates for reproduction, location<br />

<strong>of</strong> host plants and oviposition sites. Some <strong>of</strong> the key players <strong>of</strong> the moths‟ olfactory<br />

system <strong>in</strong>clude general odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s and olfactory receptors (OR). The<br />

aim <strong>of</strong> this study was to <strong>in</strong>vestigate the molecular mechanisms <strong>of</strong> olfaction <strong>in</strong> the light<br />

brown apple moth, Epiphyas postvittana. These data will contribute to the<br />

development <strong>of</strong> new pest control strategies for the moth and the development <strong>of</strong><br />

olfactory biosensors. E. postvittana OR1 (EpOR1) is closely related to the pheromone<br />

receptors (PR) <strong>of</strong> other lepidopterans, but is expressed at similar levels <strong>in</strong> male and<br />

female antennae. Functional analysis <strong>in</strong> Sf9 cells demonstrated that it b<strong>in</strong>ds important<br />

plant semiochemicals but not pheromone. EpOR1 is sensitive to methyl salicylate,<br />

recognis<strong>in</strong>g this odorant to a low concentration <strong>of</strong> 10 -15 M. EpOR1 also b<strong>in</strong>ds a range<br />

<strong>of</strong> terpenes which showed Hill slopes rang<strong>in</strong>g from 0.33–3.2.<br />

E. postvittana GOBP2 (EpGOBP2) was expressed <strong>in</strong> bacteria and purified to<br />

homogeneity. B<strong>in</strong>d<strong>in</strong>g analysis revealed it was able to b<strong>in</strong>d octanol, α-farnesene,<br />

methyl salicylate, nerol, eucalyptol, geranyl acetate and pentyl acetate, but not citral,<br />

geranial and geraniol. EpGOBP2 was able to replace DMSO as a solubilis<strong>in</strong>g agent<br />

for methyl salicylate and geranyl acetate <strong>in</strong> a functional assay <strong>of</strong> EpOR1 <strong>in</strong> Sf9 cells.<br />

The sensitivity <strong>of</strong> EpOR1 for these two ligands <strong>in</strong> vitro was enhanced <strong>in</strong> the presence<br />

<strong>of</strong> EpGOBP2 [EC50 <strong>of</strong> EpOR1 with methyl salicylate <strong>in</strong> DMSO is (1.8 ± 0.9) x 10 -<br />

12 M and EC50 <strong>of</strong> EpOR1 with methyl salicylate <strong>in</strong> the presence on EpGOBP2 is (1.19<br />

± 0.82) x 10 -13 M], suggest<strong>in</strong>g that EpGOBP2 might not only be act<strong>in</strong>g as a<br />

solubilis<strong>in</strong>g agent but also as an activated ligand.<br />

Forty-n<strong>in</strong>e new ORs were identified by deep transcriptomic sequenc<strong>in</strong>g and light<br />

coverage genome scann<strong>in</strong>g <strong>of</strong> E. postvittana br<strong>in</strong>g<strong>in</strong>g the total number <strong>of</strong> ORs<br />

identified to date for this species to 52. Phylogenetic analysis revealed that associated<br />

with l<strong>in</strong>eage, 24 ORs had one-to-one orthologs with ORs from the silkworm, Bombyx<br />

mori, while 28 E. postvittana ORs were expansions compared with B. mori. EpOR1<br />

i


Abstract<br />

and EpOR6 are the most closely related E. postvittana ORs to other moth pheromone<br />

receptors. However, EpOR1 b<strong>in</strong>ds plant semiochemicals and not pheromone<br />

components and EpOR6 rema<strong>in</strong>s to be functionally annotated. Tissue expression<br />

analysis <strong>of</strong> 26 <strong>of</strong> the ORs revealed that three (EpOR30, 33 and 34) are more than 600<br />

times more highly expressed <strong>in</strong> male than female antennae, mak<strong>in</strong>g them good<br />

candidates for be<strong>in</strong>g the pheromone receptors <strong>of</strong> E. postvittana. Functional analysis <strong>of</strong><br />

these ORs will reveal their role(s) <strong>in</strong> E. postvittana olfaction.<br />

ii


Acknowledgements<br />

Acknowledgements<br />

I would like to extend my s<strong>in</strong>cere thanks for the guidance and support I received from<br />

my supervisors: Richard Newcomb, Andrew Kralicek and David Christie. This PhD<br />

would not have been possible without your encouragement and open door policy for<br />

us students. Thank you for always be<strong>in</strong>g available for discussions from the<br />

commencement through to the completion <strong>of</strong> my research.<br />

No work is complete without the efforts <strong>of</strong> all the members <strong>of</strong> a team. My heartfelt<br />

thanks to all the members <strong>of</strong> the Molecular Sens<strong>in</strong>g Team. Thank you Melissa for<br />

do<strong>in</strong>g the groundwork on EpOR1; to Cyril for your expertise <strong>in</strong> recomb<strong>in</strong>ant prote<strong>in</strong><br />

expression and purification; to Colm and Aidan for shar<strong>in</strong>g your expertise <strong>of</strong> Sf9 cell<br />

culture; to Nadeesha for your guidance <strong>in</strong> the lab and for your friendship and to<br />

Edwige for your valuable time and guidance <strong>in</strong> deal<strong>in</strong>g with the ghosts <strong>of</strong> qRT-PCR,<br />

and your friendly support and encouragement <strong>in</strong> the highs and lows <strong>of</strong> my life. A big<br />

thank you to Pablo, Andy, Sylvia, Jeremy, Leah and Carol<strong>in</strong>e for your company <strong>in</strong> the<br />

lab and dur<strong>in</strong>g the many lunch breaks. The time spent as part <strong>of</strong> the Molecular<br />

Sens<strong>in</strong>g team would not have been enjoyable without you all!<br />

A very big thank you to Anne Barr<strong>in</strong>gton <strong>of</strong> the <strong>in</strong>sect rear<strong>in</strong>g facility at Plant and<br />

Food Research Ltd, for rear<strong>in</strong>g the thousands <strong>of</strong> sacrificial moths without whom a<br />

major part <strong>of</strong> this thesis would have been <strong>in</strong>complete. Thank you to the members <strong>of</strong><br />

the microarray facility at Plant and Food Research, especially Luke Luo for sett<strong>in</strong>g up<br />

the hybridisations and Bart Jenssen for your expertise on data analysis. I would also<br />

like to show my gratitude to the Bio<strong>in</strong>formatics department at Plant and Food<br />

Research, especially Ross Crowhurst for do<strong>in</strong>g all the raw data analysis <strong>of</strong> the tons <strong>of</strong><br />

sequenc<strong>in</strong>g with a smile! Thank you a lot Ross!<br />

I would also like to thank the Allan Wilson Centre Genome Service, especially<br />

Lorra<strong>in</strong>e Berry for your advice on sample preparations and genomic sequenc<strong>in</strong>g; and<br />

to the team at the University <strong>of</strong> Otago high-throughput DNA sequenc<strong>in</strong>g unit.<br />

iii


Acknowledgements<br />

I owe my deepest gratitude to my family, without whose support and encouragement<br />

this achievement would never have been possible. Thank you mum and dad for<br />

always encourag<strong>in</strong>g me <strong>in</strong> my studies; to my sibl<strong>in</strong>gs Shaireen, Imtiyaz, Imraan and<br />

Rizwaan for shar<strong>in</strong>g <strong>in</strong> my life. I would also like to thank my parents-<strong>in</strong>-law and<br />

Arisha for your support, and to Sohail and Shayaan for be<strong>in</strong>g the brightest stars <strong>in</strong> my<br />

life. A very big thank you to my husband Azna<strong>in</strong> for your understand<strong>in</strong>g and<br />

sacrifices dur<strong>in</strong>g the course <strong>of</strong> my project, especially for your patience and words <strong>of</strong><br />

guidance dur<strong>in</strong>g the times when writ<strong>in</strong>g seemed like an unatta<strong>in</strong>able feat. It just would<br />

not have been possible without all your love, support and encouragement.<br />

F<strong>in</strong>ally, I would like to thank School <strong>of</strong> Biological Sciences and The University <strong>of</strong><br />

Auckland for the travel grants for support<strong>in</strong>g my attendance to the Queenstown<br />

Molecular Biology Meet<strong>in</strong>g (2006), to the International Symposium <strong>of</strong> Taste and<br />

<strong>Olfaction</strong> (2008) and to the Genetics Society <strong>of</strong> Australasia (2010). This research has<br />

been made possible through grants from Foundation <strong>of</strong> Research, Science and<br />

Technology and the M<strong>in</strong>istry <strong>of</strong> Research, Science and Technology <strong>of</strong> New Zealand.<br />

iv


Contents<br />

v<br />

Contents<br />

Abstract……………………………………………………………………...…...…...i<br />

Acknowledgements…………………………………………………………….……iii<br />

List <strong>of</strong> Figures……………………………………………………………………..…ix<br />

List <strong>of</strong> Tables…………………………………………………………………...……xi<br />

List <strong>of</strong> Abbreviations…………………………………………………………...…..xii<br />

1 General Introduction ........................................................................................... 1<br />

1.1 General overview ............................................................................................ 1<br />

1.2 <strong>Olfaction</strong> <strong>in</strong> <strong>in</strong>sects .......................................................................................... 2<br />

1.3 <strong>Olfaction</strong> <strong>in</strong> moths ........................................................................................... 2<br />

1.3.1 Sex pheromones ....................................................................................... 2<br />

1.3.2 Other odorants detected by moths............................................................ 3<br />

1.4 The olfactory system ....................................................................................... 4<br />

1.5 Components <strong>of</strong> the olfactory system ............................................................... 7<br />

1.5.1 Prote<strong>in</strong>s found <strong>in</strong> the sensillum lymph..................................................... 7<br />

1.5.2 Membrane prote<strong>in</strong>s ................................................................................ 12<br />

1.5.2.1 Odorant receptor identification <strong>in</strong> moths ........................................... 13<br />

1.5.2.2 Insect odorant receptor activation ...................................................... 18<br />

1.6 Insect pest control strategies ......................................................................... 21<br />

1.7 Light brown apple moth ................................................................................ 24<br />

1.8 Aims .............................................................................................................. 32<br />

2 Functional Characterisation <strong>of</strong> Epiphyas postvittana Odorant Receptor 1 .. 34<br />

2.1 Introduction ................................................................................................... 34<br />

2.1.1 Aims ....................................................................................................... 39<br />

2.2 Materials and methods .................................................................................. 39<br />

2.2.1 Materials ................................................................................................ 39<br />

2.2.2 Insect cell culture ................................................................................... 40<br />

2.2.3 Transfection <strong>of</strong> cells ............................................................................... 40<br />

2.2.4 Detection <strong>of</strong> mRNA encod<strong>in</strong>g pIB-EpOR1 by RT-PCR ....................... 41<br />

2.2.5 Membrane Fraction Isolation and Western Blot .................................... 41<br />

2.2.6 EpOR1 functional assay ......................................................................... 43<br />

2.2.7 Data analysis .......................................................................................... 44


Contents<br />

2.3 Results ........................................................................................................... 45<br />

2.3.1 EpOR1 functional characterisation <strong>in</strong> Sf9 cells ..................................... 46<br />

2.4 Discussion ..................................................................................................... 51<br />

3 The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection by EpOR1 .... 57<br />

3.1 Introduction ................................................................................................... 57<br />

3.1.1 Aim ........................................................................................................ 60<br />

3.2 Materials and methods .................................................................................. 61<br />

3.2.1 Recomb<strong>in</strong>ant Expression <strong>of</strong> EpGOBP2 ................................................. 61<br />

3.2.1.1 Purification <strong>of</strong> recomb<strong>in</strong>ant EpGOBP2.............................................. 62<br />

3.2.1.2 Delipidation <strong>of</strong> recomb<strong>in</strong>ant EpGOBP2 ............................................ 62<br />

3.2.2 Volatile odorant b<strong>in</strong>d<strong>in</strong>g assay............................................................... 63<br />

3.2.3 Reconstituted cell assay ......................................................................... 65<br />

3.3 Results ........................................................................................................... 65<br />

3.3.1 Purification <strong>of</strong> recomb<strong>in</strong>ant GOBP2 ...................................................... 65<br />

3.3.2 Volatile Odorant B<strong>in</strong>d<strong>in</strong>g Assay ............................................................ 68<br />

3.3.3 Reconstituted EpOR1 receptor activation assays .................................. 70<br />

3.4 Discussion ..................................................................................................... 74<br />

4 Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana ....... 79<br />

4.1 Introduction ................................................................................................... 79<br />

4.1.1 Transcriptome sequenc<strong>in</strong>g – EST approach .......................................... 79<br />

4.1.2 Genome sequenc<strong>in</strong>g ............................................................................... 80<br />

4.1.3 Recent progress <strong>in</strong> sequenc<strong>in</strong>g field ...................................................... 81<br />

4.1.3.1 454 Sequenc<strong>in</strong>g .................................................................................. 82<br />

4.1.3.2 Illum<strong>in</strong>a Genome Analyzer ................................................................ 82<br />

4.1.3.3 Bio<strong>in</strong>formatics .................................................................................... 82<br />

4.1.4 Odorant Receptor Identification <strong>in</strong> Moths ............................................. 83<br />

4.1.5 Aims ....................................................................................................... 85<br />

4.2 Materials and Methods .................................................................................. 86<br />

4.2.1 Materials ................................................................................................ 86<br />

4.2.2 Total RNA Extraction ............................................................................ 86<br />

4.2.3 mRNA Isolation ..................................................................................... 86<br />

4.2.4 cDNA Synthesis ..................................................................................... 86<br />

vi


Contents<br />

4.2.5 Degenerate PCR ..................................................................................... 86<br />

4.2.6 Clon<strong>in</strong>g and Sequenc<strong>in</strong>g ........................................................................ 87<br />

4.2.7 Microarray Screen<strong>in</strong>g............................................................................. 88<br />

4.2.8 Microarray Data Analysis ...................................................................... 89<br />

4.2.9 Transcriptome Sequenc<strong>in</strong>g..................................................................... 89<br />

4.2.10 Genome Sequenc<strong>in</strong>g .............................................................................. 90<br />

4.2.11 Nuclear DNA isolation and mitochondrial DNA contam<strong>in</strong>ation test .... 90<br />

4.2.11.1 Isolation <strong>of</strong> <strong>in</strong>sect nuclei ................................................................. 90<br />

4.2.11.2 Genomic DNA extraction ............................................................... 91<br />

4.2.11.3 Mitochondrial DNA contam<strong>in</strong>ation test ......................................... 91<br />

4.2.12 qRT-PCR primer design ........................................................................ 92<br />

4.2.13 qRT-PCR primer test ............................................................................. 94<br />

4.2.14 Quantitative Real-Time PCR ................................................................. 94<br />

4.2.15 Rapid Amplification <strong>of</strong> cDNA Ends (RACE) ....................................... 96<br />

4.2.16 Bio<strong>in</strong>formatics........................................................................................ 98<br />

4.3 Results ........................................................................................................... 99<br />

4.3.1 Microarray Analysis............................................................................... 99<br />

4.3.2 Deep Transcriptomics .......................................................................... 100<br />

4.3.2.1 454 Sequenc<strong>in</strong>g Statistics ................................................................. 101<br />

4.3.3 Mitochondrial DNA contam<strong>in</strong>ation ..................................................... 102<br />

4.3.4 Illum<strong>in</strong>a Sequenc<strong>in</strong>g ............................................................................ 103<br />

4.3.5 Data M<strong>in</strong><strong>in</strong>g ......................................................................................... 103<br />

4.3.6 Quantitative Real-Time PCR ............................................................... 110<br />

4.3.7 EpOR34 – a putative PR ...................................................................... 114<br />

4.4 Discussion ................................................................................................... 119<br />

5 Conclud<strong>in</strong>g Discussion ..................................................................................... 125<br />

5.1 Introduction ................................................................................................. 125<br />

5.2 Summary <strong>of</strong> results and discussion ............................................................. 126<br />

5.2.1 EpOR1 characterisation ....................................................................... 126<br />

5.2.2 EpGOBP2 reconstitution <strong>of</strong> the Sf9 cell assay .................................... 126<br />

5.2.3 Identification <strong>of</strong> putative ORs from E. postvittana .............................. 127<br />

5.3 Current hypothesis and future directions .................................................... 128<br />

A. Appendix A – Dose response pr<strong>of</strong>ile <strong>of</strong> pIB-V5 His..................................134<br />

vii


Contents<br />

B. Appendix B – EpGOBP2 expression buffer recipe and prote<strong>in</strong> details ...... 135<br />

C. Appendix C – ClustalX multiple sequence alignment <strong>of</strong> moth PR clade .... 137<br />

D. Appendix D – Solutions for nuclear DNA isolation ...................................... 139<br />

E. Appendix E – Multiple sequence alignment <strong>of</strong> E. postvittana ORs.............. 140<br />

References ................................................................................................................. 146<br />

viii


List <strong>of</strong> Figures<br />

List <strong>of</strong> Figures<br />

Figure 1.1: Structure <strong>of</strong> a s<strong>in</strong>gle <strong>in</strong>sect olfactory sensillum ......................................... 5<br />

Figure 1.2: Odorant receptor activation ...................................................................... 10<br />

Figure 1.3: Model <strong>of</strong> <strong>in</strong>sect pheromone detection ...................................................... 13<br />

Figure 1.4: Topologies <strong>of</strong> <strong>in</strong>sect and mammalian odorant receptors .......................... 19<br />

Figure 1.5: The ion channel–GPCR model <strong>of</strong> odorant signall<strong>in</strong>g .............................. 20<br />

Figure 1.6: The sex pheromone blend <strong>of</strong> E. postvittana. ............................................ 26<br />

Figure 1.7: Scann<strong>in</strong>g electronmicrograph <strong>of</strong> adult male (left) and female (right)<br />

antennaes <strong>of</strong> E.postvittana. .......................................................................................... 27<br />

Figure 1.8: Phylogenetic tree constructed from lepidopteran odorant receptors with<br />

the three E. postvittana ORs ........................................................................................ 31<br />

Figure 2.1: Reverse transcription PCR analysis ......................................................... 45<br />

Figure 2.2: Western blot <strong>of</strong> myc-tagged EpOR1 ........................................................ 46<br />

Figure 2.3: The change <strong>in</strong> response <strong>of</strong> a s<strong>in</strong>gle Sf9 cell transfected with pIB-EpOR1<br />

over the timecourse <strong>of</strong> a calcium assay experiment. .................................................... 47<br />

Figure 2.4: Dose response curves ............................................................................... 50<br />

Figure 3.1: Schematic representation <strong>of</strong> the VOBA sett<strong>in</strong>g. ...................................... 64<br />

Figure 3.2: Purification <strong>of</strong> His6-EpGOBP2 by HiTrap chelat<strong>in</strong>g HP column. ....... 66<br />

Figure 3.3: Purification <strong>of</strong> His6-EpGOBP2 from the pooled HiTrap fractions by Q-<br />

sepharose anion exchange chromatography................................................................. 67<br />

Figure 3.4: Purification <strong>of</strong> AcTEV cleaved EpGOBP2 .............................................. 68<br />

Figure 3.5: Overnight VOBA <strong>of</strong> EpGOBP2 ............................................................... 69<br />

Figure 3.6: Dissociation constants ( ) for ten odorants aga<strong>in</strong>st EpGOBP2 ............. 69<br />

Figure 3.7: Dose response <strong>of</strong> EpOR1 express<strong>in</strong>g Sf9 cells to EpGOBP2 .................. 71<br />

Figure 3.8: Dose response curves <strong>of</strong> EpOR1 to geranyl acetate ................................. 72<br />

Figure 3.9: Dose response curves <strong>of</strong> EpOR1 to methyl salicylate.............................. 73<br />

Figure 4.1: Quantitative RT-PCR <strong>of</strong> four candidates PRs .......................................... 99<br />

Figure 4.2: E. postvittana male antennal SMART-amplified cDNA ....................... 101<br />

Figure 4.3: Semi-quantitative PCR on nuclear (Takeout 3) and mitochondrial<br />

(cytochrome oxidase I) genes .................................................................................... 103<br />

Figure 4.4: Multiple sequence alignment <strong>in</strong> ClustalX <strong>of</strong> E. postvittana ORs.......... 110<br />

Figure 4.5: Expression levels <strong>of</strong> the putative E. postvittana ORs ............................ 111<br />

ix


List <strong>of</strong> Figures<br />

Figure 4.6: Phylogram result<strong>in</strong>g from multiple sequence alignment........................ 113<br />

Figure 4.7: EpOR34 cDNA nucleotide sequence ..................................................... 115<br />

Figure 4.8: Schematic representation <strong>of</strong> the consensus <strong>of</strong> the predicted TM regions<br />

from seven TM prediction programs ......................................................................... 116<br />

Figure 4.9: Transmembrane doma<strong>in</strong> prediction <strong>of</strong> EpOR34 with TMHMM 2.0<br />

algorithms .................................................................................................................. 117<br />

Figure 4.10: Transmembrane doma<strong>in</strong> prediction <strong>of</strong> EpOR34 with TMMOD<br />

algorithms. ................................................................................................................. 117<br />

Figure 4.11: Theoretical membrane topology <strong>of</strong> EpOR34 ....................................... 118<br />

Figure A.1: Dose response <strong>of</strong> the empty pIB-V5 His...............................................134<br />

Figure B.1: Chromatogram show<strong>in</strong>g the elution pr<strong>of</strong>ile <strong>of</strong> His6-EpGOBP2 from<br />

HiTrap chelat<strong>in</strong>g HP column. ................................................................................ 136<br />

Figure B.2: Elution pr<strong>of</strong>ile <strong>of</strong> His6-EpGOBP2 on Q-sepharose HP column. ........... 136<br />

Figure C.1: ClustalX multiple sequence alignment <strong>of</strong> the moth PR clade ............... 138<br />

Figure E.1: Multiple sequence alignment <strong>in</strong> ClustalX <strong>of</strong> E. postvittana ORs with<br />

those <strong>of</strong> five other species .......................................................................................... 145<br />

x


List <strong>of</strong> Tables<br />

List <strong>of</strong> Tables<br />

Table 1.1: Summary <strong>of</strong> the moth pheromone receptors .............................................. 17<br />

Table 1.2: Plant volatiles detected by E. postvittana <strong>in</strong> electroantennogram<br />

experiments from (Suckl<strong>in</strong>g et al., 1996). ................................................................... 28<br />

Table 2.1: Change <strong>in</strong> fluorescence at 10 -5 M <strong>of</strong> the compounds tested with EpOR1 <strong>in</strong><br />

the Sf9 cell assay. ......................................................................................................... 48<br />

Table 2.2: EC50 (± standard error) and Hill slope estimates <strong>of</strong> the dose response<br />

curves <strong>of</strong> five best ligands for EpOR1 ......................................................................... 51<br />

Table 3.1: A comparison <strong>of</strong> the EC50 values ............................................................... 74<br />

Table 4.1: Outer and nested degenerate primers used <strong>in</strong> degenerate PCR <strong>of</strong> male E.<br />

postvittana antennal cDNA. ......................................................................................... 87<br />

Table 4.2: Primer pairs for qRT-PCR amplification ................................................... 93<br />

Table 4.3: Gene specific primers used for amplify<strong>in</strong>g 3‟RACE products. ................. 96<br />

Table 4.4: Primers used for amplify<strong>in</strong>g 5‟ RACE products ........................................ 98<br />

Table 4.5: Analysis <strong>of</strong> the 454 transcriptome sequence data .................................... 102<br />

Table 4.6: E. postvittana OR repertoire to date ........................................................ 104<br />

Table D.1: Reagents for mak<strong>in</strong>g 1 L <strong>of</strong> Nuclei Extraction Buffer, pH 6. ................. 139<br />

xi


List <strong>of</strong> Abbreviations<br />

ΔF Change <strong>in</strong> fluorescence<br />

°C Degrees celsius<br />

2D-GE Two dimensional gel electrophoresis<br />

ABP Antennal b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

BAC Bacterial artificial chromosome<br />

BLAST Basic local alignment search tool<br />

bp Base pair(s)<br />

cAMP Cyclic adenos<strong>in</strong>e monophosphate<br />

CCD Charge–coupled device<br />

cDNA Complementary deoxyribonucleic acid<br />

cGMP Cyclic guanos<strong>in</strong>e monophosphate<br />

CMF Chloramphenicol<br />

CSP Chemosensory prote<strong>in</strong><br />

CT Cycle threshold<br />

C–term<strong>in</strong>us Carboxy–term<strong>in</strong>us<br />

DEPC Diethyl pyrocarbonate<br />

DIG Digoxigen<strong>in</strong><br />

DMSO Dimethyl sulfoxide<br />

DNA Deoxyribonucleic acid<br />

dNTP Deoxynucleotide triphosphate<br />

dsDNA Double stranded deoxyribonucleic acid<br />

DSN Duplex-specific nuclease<br />

EAG Electroantennogram<br />

EC50<br />

Half maximal effective concentration<br />

EDTA Ethylenediam<strong>in</strong>e tetra-acetic acid<br />

EST Expressed sequence tag<br />

FID Flame ionisation detector<br />

G Gravitational constant<br />

GC Gas chromatography<br />

xii<br />

List <strong>of</strong> Abbreviations


List <strong>of</strong> Abbreviations<br />

gDNA Genomic deoxyribonucleic acid<br />

GOBP General odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

GPCR G prote<strong>in</strong>–coupled receptor<br />

G prote<strong>in</strong> Guan<strong>in</strong>e nucleotide–b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s<br />

HCl Hydrochloric acid<br />

HEK293 Human embryonic kidney–293<br />

IPM Integrated pest management<br />

IPTG Isopropyl-beta-D-1-thiogalactopyranoside<br />

kb kilo base<br />

kDa kilo Daltons<br />

LB Luria broth<br />

LDS Lithium dodecyl sulphate<br />

mRNA Messenger ribonucleic acid<br />

MWCO Molecular weight cut-<strong>of</strong>f<br />

mtDNA Mitochondrial deoxyribonucleic acid<br />

NBT/BCIP Nitro blue tetrazolium chloride/5-bromo-4-chloro-3-<strong>in</strong>dolyl<br />

phosphate<br />

NCBI National centre for biotechnology <strong>in</strong>formation<br />

N–term<strong>in</strong>us Am<strong>in</strong>o–term<strong>in</strong>us<br />

OBP Odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

OD Optical density<br />

ODE Odorant degrad<strong>in</strong>g enzyme<br />

OR Olfactory receptor<br />

ORF Open read<strong>in</strong>g frame<br />

ORN Olfactory receptor neuron<br />

OSN Olfactory sensory neuron<br />

PAGE Polyacrylamide electrophoresis<br />

PBAN Pheromone biosynthesis activat<strong>in</strong>g neuropeptide<br />

PBP Pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

PBS Phosphate–buffered sal<strong>in</strong>e<br />

PCR Polymerase cha<strong>in</strong> reaction<br />

PDE Pheromone degrad<strong>in</strong>g enzyme<br />

xiii


List <strong>of</strong> Abbreviations<br />

pH Potential <strong>of</strong> hydrogen<br />

PIPES Piperaz<strong>in</strong>e-N-N‟-bis(2-ethanesulfonic acid)<br />

PR Pheromone receptor<br />

psi Pound per square <strong>in</strong>ch<br />

PVDF Polyv<strong>in</strong>ylidene fluoride<br />

qRT-PCR Quantitative real time polymerase cha<strong>in</strong> reaction<br />

RACE Rapid amplification <strong>of</strong> cDNA ends<br />

ROI Region <strong>of</strong> <strong>in</strong>terest<br />

RNA Ribonucleic acid<br />

rpm Revolutions per m<strong>in</strong>ute<br />

S2 Schneider 2<br />

SDS Sodium dodecyl sulphate<br />

Sf9 Spodoptera frugiperda 9<br />

SNMP Sensory neuron membrane prote<strong>in</strong><br />

ssDNA S<strong>in</strong>gle stranded deoxyribonucleic acid<br />

TB Terrific broth<br />

TEV Tobacco etch virus<br />

TM Transmembrane<br />

UTR Untranslated region<br />

UV Ultra Violet<br />

VOBA Volatiles odorant b<strong>in</strong>d<strong>in</strong>g assay<br />

xiv


1.1 General overview<br />

1<br />

General Introduction<br />

<strong>Olfaction</strong> is the ability <strong>of</strong> organisms to detect and discrim<strong>in</strong>ate volatile compounds<br />

from the vast range <strong>of</strong> odours present <strong>in</strong> the environment. In animals, this sensory<br />

ability is used for such behaviours as to f<strong>in</strong>d a mate, locate food sources and detect<br />

enemies. Chemical cues are used extensively <strong>in</strong> animals as a means <strong>of</strong> communication<br />

both with<strong>in</strong> species (<strong>in</strong>traspecific) and between species (<strong>in</strong>terspecific) (Hartlieb and<br />

Anderson, 1999). Intraspecific chemical signals or pheromones are released by<br />

members <strong>of</strong> the same species and can be recognised as either releasers or primers.<br />

Releasers are pheromones that take effect immediately, for example, as <strong>in</strong> k<strong>in</strong><br />

recognition, mat<strong>in</strong>g stimulation, act<strong>in</strong>g as alarms aga<strong>in</strong>st dangers, such as predators by<br />

elicit<strong>in</strong>g aggregation, <strong>in</strong>fluenc<strong>in</strong>g oviposition, territory, trail, recruitment <strong>of</strong> new<br />

workers, nest build<strong>in</strong>g, and lead<strong>in</strong>g to food sources. Primers are pheromones that<br />

cause changes <strong>in</strong> development <strong>of</strong> the organism, for example, sexual maturation,<br />

development and physiological state (Howse, 1998). Interspecific chemical signals,<br />

which are released by one species and detected by another, can either be a kairomone<br />

whereby the signal benefits the receiver or an allomone, which is beneficial to the<br />

sender, or both. Examples <strong>of</strong> kairomones <strong>in</strong>clude pheromones, tox<strong>in</strong>s, and<br />

metabolites, used <strong>in</strong> host/prey location and floral scents (<strong>of</strong> host plants and food<br />

source). Allomones <strong>in</strong>clude defence secretions, repellents and floral scents (Jones,<br />

1998).


General Introduction 2<br />

1.2 <strong>Olfaction</strong> <strong>in</strong> <strong>in</strong>sects<br />

The olfactory system is the most widely used sensory detection method <strong>in</strong> <strong>in</strong>sects and<br />

as such is highly specific and receptive to chemical cues <strong>in</strong> the environment<br />

(Hildebrand and Shepherd, 1997). In <strong>in</strong>sects this highly developed system has<br />

warranted a great deal <strong>of</strong> focus from research groups and has become a model system<br />

for olfactory studies <strong>in</strong> general due to the relative small size <strong>of</strong> the <strong>in</strong>sect olfactory<br />

system compared with higher organisms, the organisation <strong>of</strong> olfactory receptor<br />

neurons (ORN) <strong>in</strong>to sensilla that has enabled s<strong>in</strong>gle neuron record<strong>in</strong>gs and the ability<br />

to utilise the knowledge directly <strong>in</strong> the management <strong>of</strong> horticultural and agricultural<br />

pests.<br />

1.3 <strong>Olfaction</strong> <strong>in</strong> moths<br />

Moths, belong to the order Lepidoptera and are some <strong>of</strong> the most important pests <strong>of</strong><br />

fruit and vegetable crops around the world. The detection <strong>of</strong> plant volatiles and sex<br />

pheromones by moths has been the major focus <strong>of</strong> study for the development <strong>of</strong> pest<br />

control strategies. Much focus has been on sex pheromones and their application with<br />

some other odorous signals or semiochemicals.<br />

1.3.1 Sex pheromones<br />

The first pheromones were first described <strong>in</strong> moths with the isolation <strong>of</strong> N-acetyl<br />

tyram<strong>in</strong>e (from cocoons) and bombykol (female sex pheromone), both from the<br />

silkmoth, Bombyx mori (Butenandt et al., 1959a; Butenandt et al., 1959b). S<strong>in</strong>ce then<br />

thousands <strong>of</strong> sex pheromonal compounds have been identified from hundreds <strong>of</strong><br />

<strong>in</strong>sect species, as seen <strong>in</strong> the onl<strong>in</strong>e pheromone database, (El-Sayed, 2008). Sex<br />

pheromones are produced by either the male or female or both sexes and are the most<br />

widely characterised <strong>of</strong> all pheromones. Long range sex pheromones are species-<br />

specific volatiles released from the pheromone glands found between the eighth and<br />

n<strong>in</strong>th abdom<strong>in</strong>al segments <strong>of</strong> female moths (Tamaki, 1985). The importance <strong>of</strong> sex<br />

pheromones is <strong>in</strong> mate location and courtship, where they act as chemical signals to<br />

conspecific males that the female is ready to mate and acts as a guide for the male


General Introduction 3<br />

moth <strong>in</strong> locat<strong>in</strong>g the „call<strong>in</strong>g‟ female over long distances. Lepidopteran sex<br />

pheromones are mostly composed <strong>of</strong> blends <strong>of</strong> two or more compounds (major and<br />

m<strong>in</strong>or pheromone components) <strong>in</strong> specific ratios (Butenandt et al., 1961; Arn et al.,<br />

1986). The ratios are species-specific and the pheromone may fail to elicit a response<br />

<strong>in</strong> males if altered. Sex pheromones are ma<strong>in</strong>ly composed <strong>of</strong> C10 to C18 unsaturated,<br />

straight cha<strong>in</strong> hydrocarbons together with an oxygenated functional group, for<br />

<strong>in</strong>stance an alcohol, acetate ester or aldehyde (Arn et al., 1992; Jurenka, 2003). Sex<br />

pheromone components <strong>in</strong> Lepidoptera are synthesised from fatty acids via enzymes<br />

such as acetyl-CoA, followed by desaturation whereby double bonds are placed at<br />

specific positions along the carbon cha<strong>in</strong>. Limited cha<strong>in</strong> shorten<strong>in</strong>g reactions then<br />

shorten the carbon cha<strong>in</strong> to the appropriate length and f<strong>in</strong>ally reductive modification<br />

<strong>of</strong> the carbonyl carbon generates the functional group (Foster and Roel<strong>of</strong>s, 1987;<br />

Foster and Dugdale, 1988; Jurenka, 2003).<br />

1.3.2 Other odorants detected by moths<br />

The extreme sensitivity and specificity <strong>of</strong> sex pheromone reception has made it the<br />

ideal model for study <strong>of</strong> olfaction <strong>in</strong> moths. However, other odorants such as plant<br />

volatiles <strong>in</strong>clud<strong>in</strong>g esters, aldehydes, alcohols, terpenes and carbon dioxide are also<br />

detected by the moth‟s olfactory system (Kaissl<strong>in</strong>g, 1971). These odorants are used by<br />

moths to locate host plants, oviposition sites and food sources. For example, l<strong>in</strong>alool,<br />

a plant volatile, is suggested to be used by B. mori, Heliothis virescens, Helicoverpa<br />

armigera and Spodoptera littoralis <strong>in</strong> host plant location for lay<strong>in</strong>g eggs (He<strong>in</strong>bockel<br />

and Kaissl<strong>in</strong>g, 1996; Angioy et al., 2003; Røstelien et al., 2005; Anderson et al.,<br />

2009). Herbivore-<strong>in</strong>duced volatiles, which are released by plants <strong>in</strong> response to <strong>in</strong>sect<br />

feed<strong>in</strong>g, result <strong>in</strong> the recruitment <strong>of</strong> beneficial <strong>in</strong>sects that parasitise the <strong>in</strong>fest<strong>in</strong>g<br />

<strong>in</strong>sects. One such example is methyl salicylate, which is detected by moths such as<br />

Mamestra brassica, S. littoralis, and Epiphyas postvittana as a warn<strong>in</strong>g that a plant is<br />

already occupied and thus females avoid such plants for lay<strong>in</strong>g eggs (Suckl<strong>in</strong>g et al.,<br />

1996; Jönsson and Anderson, 1999; Ulland et al., 2008). This is a survival mechanism<br />

used by the moths <strong>in</strong> order to m<strong>in</strong>imise competition once the larvae hatch. Eugenol,<br />

geraniol and citral have also been shown to act as oviposition deterrents for E.<br />

postvittana, while hexanal, l<strong>in</strong>alool, nonanol, octanol and nonanal act as attractants.<br />

Plant volatiles such as citral, nonanol, octanol and n-decylaldehyde have been shown


General Introduction 4<br />

<strong>in</strong> behavioural assays to decrease the proportion <strong>of</strong> E. postvittana lay<strong>in</strong>g eggs while<br />

some compounds have implications <strong>in</strong> decreas<strong>in</strong>g the number <strong>of</strong> fertile eggs laid per<br />

female moth (Suckl<strong>in</strong>g et al., 1996).<br />

1.4 The olfactory system<br />

Antennae are the major odorant detection appendages <strong>in</strong> moths. Hair-like projections<br />

on the surface <strong>of</strong> the antennae called sensilla house chemosensory organs that detect<br />

odorants present <strong>in</strong> the environment (Schneider, 1964). A sensillum has a porous<br />

outer wall and the <strong>in</strong>ner layer is filled with an aqueous fluid, the sensillum lymph<br />

(Breer, 1997). The bipolar olfactory receptor neuron projects a dendritic membrane<br />

<strong>in</strong>to the sensillum lymph, and the other end forms an axon project<strong>in</strong>g to the antennal<br />

lobe (Keil, 1999). The ORN cell is surrounded by three auxiliary cells, the tormogen,<br />

trichogen and thechogen (Ste<strong>in</strong>brecht et al., 1992). These cells function <strong>in</strong> sensillum<br />

structural development and <strong>in</strong> ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g the sensillum lymph. Located with<strong>in</strong> the<br />

sensillum lymph are soluble odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s (OBPs) and odorant degrad<strong>in</strong>g<br />

enzymes (ODEs) and located with<strong>in</strong> the dendritic membrane are the odorant receptors<br />

(ORs) and pheromone receptors (PRs), as depicted <strong>in</strong> Figure 1.1 (Vogt and Riddiford,<br />

1981; Ste<strong>in</strong>brecht, 1999; Vogt et al., 2005).


General Introduction 5<br />

Figure 1.1: Structure <strong>of</strong> a s<strong>in</strong>gle <strong>in</strong>sect olfactory sensillum (adapted and modified<br />

from Vogt (1987)). Odorants enter the sensillum lymph through pores <strong>in</strong> the cuticular<br />

wall and <strong>in</strong>teract with the b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s present <strong>in</strong> the lymph which confer them<br />

onto the odorant receptors present on the dendritic membrane <strong>of</strong> the bipolar olfactory<br />

receptor neuron. Degrad<strong>in</strong>g enzymes present <strong>in</strong> the sensillum lymph may help <strong>in</strong> the<br />

clear<strong>in</strong>g <strong>of</strong> odorants from the lymph.<br />

The sensillum types present on moth antennae can be classified <strong>in</strong>to six different<br />

groups. These are the sensilla trichodea, sensilla basiconica, sensilla auricillica,<br />

sensilla chaetica, sensilla styloconica and sensilla coeloconica. Sensilla trichodea are<br />

very long, th<strong>in</strong>, porous hair-like structures with sharp po<strong>in</strong>ted tips (Schneider, 1964).<br />

Unbranched dendrites <strong>of</strong> up to three neurons <strong>in</strong>nervate this sensilla type (Keil, 1999).<br />

They are present <strong>in</strong> high numbers <strong>in</strong> male moths, while they occur <strong>in</strong> low numbers or<br />

not at all <strong>in</strong> female moths and thus were implicated to be <strong>in</strong>volved <strong>in</strong> recognition <strong>of</strong><br />

sex pheromone components released by female moths (Seabrook, 1978). The specific<br />

tun<strong>in</strong>g <strong>of</strong> this sensilla type to sex pheromones was confirmed by s<strong>in</strong>gle sensillum<br />

electroantennogram (EAG) record<strong>in</strong>gs (Shields and Hildebrand, 2001). The females<br />

<strong>of</strong> some moth species also carry large numbers <strong>of</strong> this sensilla type, however, these<br />

are shorter and the neurons are responsive to general odorants (Nagy and George,<br />

1981). Sensilla basiconica are shorter with blunt rounded tips and are <strong>in</strong>nervated by<br />

the dendrites <strong>of</strong> up to 50 neurons (Schneider, 1964). This sensilla type has been<br />

implicated <strong>in</strong> the detection <strong>of</strong> plant volatiles <strong>in</strong> S. littoralis (Anderson et al., 1995) and


General Introduction 6<br />

Cactoblastic cactorum (Poph<strong>of</strong> et al., 2005) <strong>in</strong> electrophysiological studies. Sensilla<br />

auricillica is a shoehorn shaped, porous sensilla that is <strong>in</strong>nervated by the branched<br />

dendrites <strong>of</strong> three neurons. They are <strong>in</strong>volved <strong>in</strong> detect<strong>in</strong>g plant volatiles (Anderson et<br />

al., 2000) and sex pheromone components <strong>in</strong> Cydia pomonella (Ebb<strong>in</strong>ghaus et al.,<br />

1998; Ansebo et al., 2005). Sensilla chaetica is similar to the trichoid sensilla except<br />

that it has thicker walls and the base has a flexible circular membrane (Schneider,<br />

1964). This sensilla type are either devoid <strong>of</strong> cuticular pores with a function <strong>in</strong><br />

mechano-reception, or have only one pore at their tip and function <strong>in</strong> gustatory<br />

reception (Keil, 1999). Sensilla styloconica are short, devoid <strong>of</strong> pores with a peg-like<br />

shape and function as thermo– or hygro– receptors (Shields and Hildebrand, 2001).<br />

The last sensilla type, sensilla coeloconica are very short, double walled, shaped like<br />

peg and are located <strong>in</strong> pits below the antennae cuticle. The sensilla are aporous but the<br />

double wall is not fused hence odours can move through it. This sensilla type is<br />

<strong>in</strong>nervated by dendrites <strong>of</strong> five neurons (Altner et al., 1977) and are <strong>in</strong>volved <strong>in</strong><br />

detection <strong>of</strong> aliphatic acids and aldehydes <strong>in</strong> B. mori and C. cactorum (Poph<strong>of</strong>, 1997;<br />

Poph<strong>of</strong> et al., 2005).<br />

The cascade <strong>of</strong> events that leads from the detection <strong>of</strong> an odorant present <strong>in</strong> the<br />

environment and its conversion <strong>in</strong>to a neuronal signal <strong>in</strong> the antennal lobe <strong>of</strong> the moth<br />

is still unclear. A proposed mechanism <strong>of</strong> the perireceptor events that follow detection<br />

<strong>of</strong> an odorant is as follows (Kaissl<strong>in</strong>g, 2009). The hydrophobic odorant molecule<br />

enters the sensillum lymph via waxed pores <strong>in</strong> the cuticular wall <strong>of</strong> the sensilla. This<br />

hydrophobic molecule has to travel through the aqueous layer to the membrane bound<br />

receptors <strong>in</strong> order for signal transduction to occur. Soluble prote<strong>in</strong>s present <strong>in</strong><br />

abundance <strong>in</strong> the lymph may function <strong>in</strong> transport<strong>in</strong>g the odorants by form<strong>in</strong>g<br />

complexes. The b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> the odorant to the membrane receptor activates a series <strong>of</strong><br />

signall<strong>in</strong>g cascades <strong>in</strong> the ORN caus<strong>in</strong>g action potentials to be generated. This results<br />

<strong>in</strong> the conveyance <strong>of</strong> the signal to the antennal lobe and process<strong>in</strong>g <strong>in</strong> higher bra<strong>in</strong><br />

centres. Odorants need to be cleared from the olfactory system once the signal has<br />

been relayed to the ORNs. This ensures the sensitivity and specificity <strong>of</strong> the system<br />

and the detection <strong>of</strong> new odorants or different concentrations <strong>of</strong> odorant plumes that<br />

direct the moths‟ flight (towards or away from attractants and repellents) (Kaissl<strong>in</strong>g,<br />

1974; Vogt, 1987; Prestwich et al., 1989; Kaissl<strong>in</strong>g, 2009).


General Introduction 7<br />

1.5 Components <strong>of</strong> the olfactory system<br />

The identification <strong>of</strong> several prote<strong>in</strong> components <strong>of</strong> the olfactory system <strong>in</strong> <strong>in</strong>sects<br />

such as fruit flies and moths has paved the way for further detailed study <strong>of</strong> these<br />

prote<strong>in</strong>s and the peripheral events <strong>in</strong>volved <strong>in</strong> signal transduction. The identification,<br />

isolation and functional analysis <strong>of</strong> these prote<strong>in</strong>s are described below.<br />

1.5.1 Prote<strong>in</strong>s found <strong>in</strong> the sensillum lymph<br />

Abundant <strong>in</strong> the sensillum lymph are small, soluble prote<strong>in</strong>s, rang<strong>in</strong>g <strong>in</strong> size from 14–<br />

16 kDa called odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s (OBP). These prote<strong>in</strong>s are produced <strong>in</strong> the<br />

sensillum support cells and secreted <strong>in</strong>to the lymph (Vogt and Riddiford, 1981). The<br />

first OBP to be identified was a 15 kDa male antennal specific prote<strong>in</strong>, unique to the<br />

pheromone sensitive sensilla, that bound sex pheromones <strong>in</strong> the wild silkmoth<br />

Antheraea polyphemus (Vogt and Riddiford, 1981). Due to its close association with<br />

sex pheromones, it was called a pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> (PBP). PBPs are generally<br />

found at high concentrations (10 mM) <strong>in</strong> male <strong>in</strong>sect antennae (Kle<strong>in</strong>, 1987), but<br />

some PBPs have also been identified <strong>in</strong> female moth antennas (Callahan et al., 2000).<br />

Low levels <strong>of</strong> PBP have been identified <strong>in</strong> some Lepidoptera species while high<br />

expression levels <strong>of</strong> PBPs have been shown <strong>in</strong> members <strong>of</strong> the Noctuoidea (Kle<strong>in</strong>,<br />

1987; Callahan et al., 2000). It was not until 1991 when the first evidence <strong>of</strong><br />

multigene families <strong>of</strong> OBPs emerged with the identification <strong>of</strong> two groups <strong>of</strong> general<br />

odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s (GOBP), namely GOBP1 and GOBP2; prote<strong>in</strong>s that are not<br />

sex-biased and have highly conserved regions with<strong>in</strong> members <strong>of</strong> the groups (Vogt et<br />

al., 1991; Krieger et al., 1993; Krieger et al., 1996). Most PBPs were identified and<br />

characterised by their b<strong>in</strong>d<strong>in</strong>g to radio-labelled versions <strong>of</strong> known sex pheromone<br />

components. GOBPs on the other hand have been less studied <strong>in</strong> terms <strong>of</strong> their<br />

b<strong>in</strong>d<strong>in</strong>g partners. In 1997, Feng and Prestwich deduced the ligand <strong>of</strong> GOBP2 from<br />

Manduca sexta by competitive displacement <strong>of</strong> tritium-labelled diazoacetate<br />

pheromone analog, [ 3 H]-6E,11Z-16:Dza, by the plant odorants (Z)-3-hexen-1-ol,<br />

geraniol, geranyl acetate and limonene (Feng and Prestwich, 1997). In a recent study<br />

<strong>of</strong> OBPs from B. mori, GOBP2 was shown to b<strong>in</strong>d the sex pheromone <strong>of</strong> B. mori,<br />

bombykol and also to be able to discrim<strong>in</strong>ate it from its antagonist, bombykal (Zhou<br />

et al., 2009). The specific function <strong>of</strong> OBPs is still unclear but several roles <strong>of</strong> OBPs


General Introduction 8<br />

have been postulated based on experimental data and modell<strong>in</strong>g <strong>of</strong> the perireceptor<br />

and receptor events <strong>in</strong>volved <strong>in</strong> olfactory perception <strong>of</strong> moths (Kaissl<strong>in</strong>g, 2009).<br />

These <strong>in</strong>clude roles <strong>in</strong> the solubilisation and transport <strong>of</strong> ligands, protection <strong>of</strong> the<br />

ligands from enzymatic degradation by formation <strong>of</strong> OBP/ligand complex, <strong>in</strong>teraction<br />

<strong>of</strong> the complex with specific ORs, or the OBP might be <strong>in</strong>volved <strong>in</strong> the deactivation<br />

<strong>of</strong> the ligand.<br />

OBPs have been shown to be <strong>in</strong>volved <strong>in</strong> the solubilisation <strong>of</strong> odorants <strong>in</strong> <strong>in</strong> vivo and<br />

<strong>in</strong> vitro assays. van den Berg and Ziegelberger (1991) perfused A. polyphemus sensilla<br />

trichodea with the pheromone (E,Z)-6,11-hexadecadienyl acetate dissolved <strong>in</strong> R<strong>in</strong>ger<br />

solution with or without PBP (ApolPBP) and showed that the pheromone<br />

concentration required to elicit a response from the ORN was decreased 100-fold <strong>in</strong><br />

the presence <strong>of</strong> ApolPBP (van den Berg and Ziegelberger, 1991). As the pheromone<br />

solutions were applied to the ORN through a glass capillary, most <strong>of</strong> the pheromone<br />

that would have otherwise adhered to the capillary wall <strong>in</strong> the absence <strong>of</strong> the PBP<br />

would now be bound to the PBP hence rema<strong>in</strong> <strong>in</strong> the soluble phase and be conferred<br />

to the ORN. This suggests a role for the A. polyphemus PBP <strong>in</strong> solubility and perhaps<br />

transport <strong>of</strong> the pheromone through the sensillum lymph to the ORN.<br />

In vitro b<strong>in</strong>d<strong>in</strong>g assays have also been used to show a role for PBPs as solubilis<strong>in</strong>g<br />

agents for pheromones. Groβe-Wilde et al. (2006) generated stable cell l<strong>in</strong>es<br />

express<strong>in</strong>g B. mori OR1 (BmOR1) <strong>in</strong> Human Embryonic Kidney (HEK293) cells.<br />

Functional assays where the ligand was solubilised <strong>in</strong> the organic solvent dimethyl<br />

sulfoxide (DMSO) showed that BmOR1 express<strong>in</strong>g cells elicited significant response<br />

to bombykol and bombykal. When the DMSO was replaced by BmPBP as the<br />

solubilis<strong>in</strong>g agent, BmOR1 express<strong>in</strong>g cells were able to elicit a response to<br />

bombykol with similar <strong>in</strong>tensities (Große-Wilde et al., 2006). In another study the H.<br />

virescens pheromone receptor HvOR13, elicits a significant response to the ma<strong>in</strong> sex<br />

pheromone component Z11-hexadecenal when the pheromone is solubilised <strong>in</strong><br />

DMSO, and this response sensitivity is <strong>in</strong>creased by 10 4 -fold when the DMSO is<br />

replaced with HvPBP2, <strong>in</strong>dicat<strong>in</strong>g roles <strong>of</strong> the PBP as a solubilis<strong>in</strong>g agent. The PBP<br />

might also be act<strong>in</strong>g as a transporter, ferry<strong>in</strong>g and <strong>in</strong>creas<strong>in</strong>g the local concentration<br />

<strong>of</strong> Z11-hexadecenal <strong>in</strong> close proximity to the PR (Große-Wilde et al., 2007).


General Introduction 9<br />

OBPs have also been implicated <strong>in</strong> protect<strong>in</strong>g the bound ligand from degradation. In<br />

A. polyphemus, a pheromone degrad<strong>in</strong>g enzyme (PDE) has been isolated and shown<br />

to effectively degrade the pheromone <strong>in</strong> vitro (Vogt et al., 1985). However, this<br />

activity is reduced <strong>in</strong> the presence <strong>of</strong> PBP, an <strong>in</strong>dication that the PBP is protect<strong>in</strong>g the<br />

pheromone from the PDE (Vogt and Riddiford, 1986). This is also supported by <strong>in</strong><br />

vivo experiments <strong>in</strong> B. mori and A. polyphemus that show two different rates <strong>of</strong><br />

pheromone degradation, with 17% <strong>of</strong> the pheromone degraded <strong>in</strong>itially and the<br />

rema<strong>in</strong><strong>in</strong>g 83% degraded at a slower rate. One explanation for this might be the<br />

formation <strong>of</strong> PBP/pheromone complex as the pheromone enters the sensillum lymph<br />

and its consequent protection from PDE (Kasang, 1971; Kasang and Kaissl<strong>in</strong>g, 1972;<br />

Kasang, 1973; Kasang et al., 1988; Kasang et al., 1989a; Kasang et al., 1989b).<br />

In vivo b<strong>in</strong>d<strong>in</strong>g assays provide evidence that the pheromone/PBP complex <strong>in</strong>teracts<br />

with the receptor molecules. Syed et al. (2006) expressed BmOR1 <strong>in</strong> the empty-<br />

neuron system, a mutant fly stra<strong>in</strong> that is devoid <strong>of</strong> its native receptors <strong>in</strong> the ab3A<br />

neuron, and showed that receptor activity was achieved when the neuron was<br />

stimulated with bombykol, albeit with low sensitivity. This sensitivity was enhanced<br />

by the co-expression <strong>of</strong> the B. mori PBP, <strong>in</strong>dicat<strong>in</strong>g that the PBP is <strong>in</strong>volved <strong>in</strong> the<br />

pheromone/receptor <strong>in</strong>teraction. Selective response <strong>of</strong> receptor neurons to pheromone<br />

components have been observed <strong>in</strong> the presence <strong>of</strong> different PBPs. A. polyphemus<br />

sensilla trichodea express<strong>in</strong>g the ORN tuned to the pheromone component (E,Z)-6,11-<br />

hexadecadienyl acetate elicited a response upon stimulation with the pheromone<br />

solubilised <strong>in</strong> DMSO, ApolPBP1 and ApolPBP3 (Poph<strong>of</strong>, 2004). However, no<br />

response was elicited when the pheromone was solubilised <strong>in</strong> ApolPBP2, giv<strong>in</strong>g<br />

further evidence <strong>of</strong> pheromone/PBP complex <strong>in</strong>teraction with PR. Evidence for such<br />

pheromone/PBP complex <strong>in</strong>teractions have also been shown <strong>in</strong> vivo and <strong>in</strong> vitro <strong>in</strong> B.<br />

mori. When solubilised <strong>in</strong> DMSO, both bombykol and bombykal b<strong>in</strong>d BmOR1 but<br />

when DMSO is replaced with BmPBP, only the bombykol/BmPBP complex is able to<br />

activate BmOR1 (Große-Wilde et al., 2006).<br />

Interaction <strong>of</strong> the pheromone/PBP complex with receptor molecules has also been<br />

shown <strong>in</strong> flies. The Drosophila OBP LUSH has been suggested to <strong>in</strong>teract directly<br />

with ORs (Kim et al., 1998). Flies mutant for LUSH were previously shown to have<br />

odour defects <strong>in</strong> that they were unable to avoid high concentrations <strong>of</strong> alcohol, as


General Introduction 10<br />

opposed to wild type flies that avoid high alcohol levels. Further research with LUSH<br />

mutants showed that this OBP is <strong>in</strong>volved <strong>in</strong> activity <strong>of</strong> the pheromone-sensitive<br />

neurons. The activity <strong>of</strong> the male-specific lipid, 11-cis vaccenyl acetate (cVA) that<br />

causes aggregation <strong>in</strong> wild type flies is lost <strong>in</strong> mutants. The activity <strong>of</strong> this pheromone<br />

is restored with transgenic expression <strong>of</strong> the LUSH prote<strong>in</strong> suggest<strong>in</strong>g that the loss <strong>of</strong><br />

function <strong>in</strong> mutants is specifically due to absence <strong>of</strong> LUSH <strong>in</strong> the pheromone<br />

sensitive sensilla (Zhou et al., 2004; Xu et al., 2005). Further to this, Laughl<strong>in</strong> et al.<br />

(2008) showed through solv<strong>in</strong>g the structures <strong>of</strong> various bound and unbound states <strong>of</strong><br />

LUSH, as well as <strong>of</strong> several LUSH prote<strong>in</strong>s with po<strong>in</strong>t mutations that the cVA bound<br />

LUSH acts as an activator <strong>of</strong> pheromone sensitive neurons (Laughl<strong>in</strong> et al., 2008).<br />

When cVA is bound to LUSH, a conformational change occurs that disrupts a salt<br />

bridge <strong>in</strong> LUSH and causes shifts <strong>in</strong> its surface loop. This altered state <strong>of</strong> LUSH then<br />

activates the pheromone neuron, as depicted <strong>in</strong> Figure 1.2. The salt bridge is<br />

ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> the unbound state and when an alcohol such as butanol is bound to<br />

LUSH. This <strong>in</strong>dicates that the OBP/ligand complex <strong>in</strong>teracts with the PR (Laughl<strong>in</strong> et<br />

al., 2008).<br />

Figure 1.2: Odorant receptor activation as depicted by Laughl<strong>in</strong> et al. (2008). The<br />

formation <strong>of</strong> the cVA/LUSH complex results <strong>in</strong> a conformational change <strong>of</strong> LUSH<br />

mak<strong>in</strong>g it an active ligand that <strong>in</strong>teracts directly with the pheromone receptor. From<br />

Stowers and Logan (2008).<br />

PR


General Introduction 11<br />

A model has been proposed for ligand deactivation by modification <strong>of</strong> the PBP <strong>in</strong> the<br />

pheromone complex to a scavenger form (Kaissl<strong>in</strong>g, 2009). In B. mori, when the<br />

PBP/bombykol complex <strong>in</strong>teracts with the receptor molecule on the dendrite <strong>of</strong> the<br />

ORN, the low pH at the dendrite leads to a conformational change and formation <strong>of</strong> a<br />

C- term<strong>in</strong>al α-helix <strong>in</strong> the PBP, thereby releas<strong>in</strong>g the bombykol (Leal et al., 2005).<br />

Several mechanisms by which the odorant is conferred to the ORs by the<br />

OBP/odorant complex have been proposed. The stable OBP/odorant complex<br />

<strong>in</strong>teracts directly with the OR; or the OBP/odorant complex is unstable and at the<br />

dendritic membrane, the odorant b<strong>in</strong>ds to the OR selectively. It could also be that a<br />

dendritic membrane bound dock<strong>in</strong>g prote<strong>in</strong>, for example, sensory neuron membrane<br />

prote<strong>in</strong> 1, b<strong>in</strong>ds the OBP/odorant complex and this <strong>in</strong> turn releases the odorant from<br />

the complex and confers it to the OR (Vogt et al., 1985; Prestwich et al., 1995;<br />

Ste<strong>in</strong>brecht, 1996; Kaissl<strong>in</strong>g, 1998). A recent structural based study <strong>of</strong> A. polyphemus<br />

PBP proposes a pH <strong>in</strong>duced release <strong>of</strong> odorants from the complex (Zubkov et al.,<br />

2005). B<strong>in</strong>d<strong>in</strong>g <strong>of</strong> odorant to PBP occurs at neutral lymph pH and release <strong>of</strong> the<br />

odorant from the complex is achieved by the open<strong>in</strong>g <strong>of</strong> the ligand b<strong>in</strong>d<strong>in</strong>g cavity at<br />

the lower pH near the dendritic membrane.<br />

Other classes <strong>of</strong> OBPs <strong>in</strong>clude chemosensory prote<strong>in</strong>s (CSPs) and antennal b<strong>in</strong>d<strong>in</strong>g<br />

prote<strong>in</strong>s (ABPs). The ABPs are expressed specifically <strong>in</strong> the antennae (Ste<strong>in</strong>brecht et<br />

al., 1995; Zhang et al., 2001). The CSPs b<strong>in</strong>d odorants and pheromones and also have<br />

implications <strong>in</strong> taste reception. The first CSP was identified by subtractive<br />

hybridisation experiments <strong>in</strong> Drosophila melanogaster antennae (McKenna et al.,<br />

1994; Pikielny et al., 1994). Refer to Vogt et al. (2003) for a review.<br />

Another class <strong>of</strong> soluble prote<strong>in</strong>s present <strong>in</strong> the sensillum lymph are the odorant<br />

degrad<strong>in</strong>g enzymes (ODEs) (Vogt and Riddiford, 1981). Once the signal has been<br />

relayed to the neuron, the odorant needs to be cleared from the lymph so that<br />

sensitivity <strong>of</strong> the olfactory system is ma<strong>in</strong>ta<strong>in</strong>ed and new signals can be detected as<br />

cont<strong>in</strong>uous fir<strong>in</strong>g <strong>of</strong> the neurons with the same odorant can cause a loss <strong>of</strong> sensitivity<br />

to that odorant. Also, not all odorants that enter the sensillum are detected by the<br />

olfactory system and need to be cleared from the lymph. This may <strong>in</strong>clude odorants<br />

toxic to the cells so degradation <strong>of</strong> these odorants is important <strong>in</strong> the lymph. This role<br />

<strong>in</strong> <strong>in</strong>sects is suggested to be played by ODEs such as carboxylesterases, glutathione-


General Introduction 12<br />

S-transferases, and cytochrome P450. See Vogt et al. (1985) and Vogt et al. (2003)<br />

for reviews.<br />

1.5.2 Membrane prote<strong>in</strong>s<br />

Two classes <strong>of</strong> membrane prote<strong>in</strong>s have been postulated to be <strong>of</strong> importance to moth<br />

olfaction, the first one be<strong>in</strong>g sensory neuron membrane prote<strong>in</strong>s (SNMPs) and the<br />

other be<strong>in</strong>g odorant receptors. SNMPs were discovered <strong>in</strong> 1988 (Vogt et al., 1988;<br />

Rogers et al., 2001) <strong>in</strong> the antennal lymph <strong>of</strong> A. polyphemus. There are two classes <strong>of</strong><br />

SNMPs (SNMP1 and SNMP2) <strong>in</strong> moths, with 25 to 75% am<strong>in</strong>o acid identity between<br />

them (Vogt et al., 2003). They were thought to be candidates for the PR <strong>of</strong> moths due<br />

to their antennal specific expression, pheromone b<strong>in</strong>d<strong>in</strong>g ability and expression onset<br />

synchronised with olfactory function development <strong>in</strong> moths (Vogt et al., 1988; Rogers<br />

et al., 2001a). This hypothesis was later rejected when evidence emerged that SNMPs<br />

were expressed <strong>in</strong> similar levels <strong>in</strong> both male and female antennae, had only two<br />

transmembrane doma<strong>in</strong>s, were found <strong>in</strong> most olfactory sensilla types and had low<br />

diversity (Rogers et al., 1997; Rogers et al., 2001). SNMPs are similar to mammalian<br />

membrane bound CD36 prote<strong>in</strong>s. CD36 has implications <strong>in</strong> cell-to-cell <strong>in</strong>teraction, <strong>in</strong><br />

the transport <strong>of</strong> small hydrophobic molecules and b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>/lipid complexes<br />

(Vogt et al., 2003). A study carried out by Benton et al. (2007) showed that SNMP is<br />

essential <strong>in</strong> pheromone <strong>in</strong>duced response <strong>of</strong> PRs <strong>in</strong> D. melanogaster (Figure 1.3).<br />

SNMP is expressed <strong>in</strong> high concentrations <strong>in</strong> the trichoid sensilla <strong>of</strong> Drosophila but<br />

does not affect the development <strong>of</strong> trichoid OSNs and support cells. Several different<br />

pheromone b<strong>in</strong>d<strong>in</strong>g experiments showed that Drosophila SNMP is essential for cVA<br />

mediated response <strong>of</strong> OR67d. H. virescens PR, HvOR13 when expressed <strong>in</strong> OR67d<br />

neurons also required SNMP to elicit response to the H. virescens pheromone<br />

component (Z)-11-hexadecenal. This study <strong>in</strong>dicates a role <strong>of</strong> <strong>in</strong>sect SNMPs as a co-<br />

receptor for pheromone reception, due to their preference for b<strong>in</strong>d<strong>in</strong>g lipids (Benton et<br />

al., 2007). SNMP may function <strong>in</strong> olfactory perception <strong>of</strong> pheromones by <strong>in</strong>teract<strong>in</strong>g<br />

with different components <strong>of</strong> the olfactory system. It could be <strong>in</strong>volved <strong>in</strong><br />

destabilis<strong>in</strong>g the PBP/pheromone complex and act<strong>in</strong>g as a prelim<strong>in</strong>ary receptor for the<br />

pheromone, convey<strong>in</strong>g it to the PR. It could also couple to the signal transduction<br />

pathway or be <strong>in</strong>volved <strong>in</strong> signal term<strong>in</strong>ation (Rogers et al., 2001; Vogt et al., 2003).


General Introduction 13<br />

Figure 1.3: Model <strong>of</strong> <strong>in</strong>sect pheromone detection demonstrat<strong>in</strong>g a role <strong>of</strong> SNMP <strong>in</strong><br />

cVA mediated response <strong>of</strong> OR67d, as depicted by Benton et al. (2007). The SNMP<br />

may be act<strong>in</strong>g as a prelim<strong>in</strong>ary receptor, destabilis<strong>in</strong>g the OBP/pheromone complex<br />

and convey<strong>in</strong>g the pheromone to the PR for signal transduction to occur.<br />

Odorant receptors are seven transmembrane (7TM) doma<strong>in</strong> prote<strong>in</strong>s localised on the<br />

dendritic membranes <strong>of</strong> ORNs and detect odorants that enter the sensillum lymph<br />

(Benton, 2006). ORs function <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g odorants as they are transported to the<br />

dendritic membrane. This <strong>in</strong>teraction sets <strong>of</strong>f a signall<strong>in</strong>g cascade caus<strong>in</strong>g the ORN to<br />

depolarise and a signal is relayed via the axonal end <strong>of</strong> the ORN to the antennal lobe<br />

<strong>in</strong> the bra<strong>in</strong>. In contrast to the broadly tuned ORs, pheromone receptors <strong>in</strong> <strong>in</strong>sects, like<br />

moths, are thought to be highly tuned to sex pheromones (Nakagawa et al., 2005;<br />

Mitsuno et al., 2008; Forstner et al., 2009; Wang et al., 2010). This represents the<br />

holy grail <strong>of</strong> the <strong>in</strong>sect olfactory system <strong>in</strong> that the identification <strong>of</strong> PRs will give a<br />

better understand<strong>in</strong>g <strong>of</strong> the underly<strong>in</strong>g mechanisms <strong>of</strong> sex pheromone reception and<br />

pheromone controlled behaviours such as mat<strong>in</strong>g.<br />

1.5.2.1 Odorant receptor identification <strong>in</strong> moths<br />

The first putative ORs to be identified <strong>in</strong> moths were from the tobacco budworm H.<br />

virescens us<strong>in</strong>g the basic local alignment search tool (BLAST) <strong>of</strong> Heliothis genomic


General Introduction 14<br />

sequence with candidate OR sequences from Drosophila as queries (Altschul et al.,<br />

1990; Krieger et al., 2002). The genomic sequences obta<strong>in</strong>ed from the BLAST search<br />

were used as probes for obta<strong>in</strong><strong>in</strong>g the exon regions <strong>of</strong> the sequences from a H.<br />

virescens cDNA library. This process identified n<strong>in</strong>e OR genes <strong>in</strong> H. virescens and<br />

further screen<strong>in</strong>g <strong>of</strong> the genomic sequences as well as transcriptome sequenc<strong>in</strong>g have<br />

led to the identification <strong>of</strong> a total <strong>of</strong> 18 ORs <strong>in</strong> H. virescens so far (Krieger et al.,<br />

2004). Five <strong>of</strong> these ORs (HvOR11, HvOR13–HvOR16) have been identified as<br />

potential PRs <strong>of</strong> H. virescens based on the high levels <strong>of</strong> expression <strong>of</strong> these genes <strong>in</strong><br />

male than female antennae (Krieger et al., 2004; Vásquez et al., 2010). In situ<br />

hybridisations <strong>of</strong> HvOR11, HvOR13, HvOR14 and HvOR16 showed their expression<br />

was conf<strong>in</strong>ed to antennal cells below the sensilla trichodea (Große-Wilde et al., 2007;<br />

Baker, 2009; Krieger et al., 2009) and functional characterisation <strong>in</strong> Xenopus laevis<br />

oocytes has confirmed HvOR13, HvOR14 and HvOR16 to b<strong>in</strong>d (Z)-11-hexadecenal,<br />

(Z)-11-hexadecenyl acetate and (Z)-11-hexadecen-1-ol respectively (Wang et al.,<br />

2010). (Z)-9-tetradecenal forms 5% <strong>of</strong> the sex pheromone component <strong>of</strong> H. virescens<br />

and HvOR6, which is expressed <strong>in</strong> both male and female antennae, and was identified<br />

as the receptor for this pheromone component by expression and characterisation <strong>in</strong> X.<br />

laevis oocytes (Wang et al., 2010).<br />

The H. virescens ORs share very little homology to other known <strong>in</strong>sect ORs, except<br />

for HvOR2, which is highly conserved across <strong>in</strong>sects and has a role as a co-receptor <strong>in</strong><br />

olfactory signall<strong>in</strong>g (refer to section 1.5.2.2 for details <strong>of</strong> this receptor type).<br />

Digoxigen<strong>in</strong> (DIG) labelled probes designed to HvOR2 were used to isolate the first<br />

ORs from B. mori and Antheraea pernyi antennal cDNA libraries (Krieger et al.,<br />

2003). Screen<strong>in</strong>g <strong>of</strong> a B. mori male antennal cDNA library as well as the B. mori<br />

genome sequences <strong>in</strong> GenBank with other H. virescens OR sequences led to the<br />

identification <strong>of</strong> a total <strong>of</strong> six ORs (Sakurai et al., 2004; Krieger et al., 2005). In situ<br />

hybridisations has revealed BmOR1 and BmOR3 to be expressed <strong>in</strong> the sensilla<br />

trichodea, while BmOR2 has dispersed expression not conf<strong>in</strong>ed to the sensilla<br />

trichodea, and BmOR4 and BmOR5 are expressed <strong>in</strong> fewer cells. BmOR6 did not<br />

label any cells. Heterologous expression <strong>of</strong> BmOR1 and BmOR3 <strong>in</strong> X. laevis oocytes<br />

have identified their ligands as bombykol and bombykal, respectively (Nakagawa et<br />

al., 2005). The availability <strong>of</strong> the B. mori genome sequence has facilitated the<br />

identification <strong>of</strong> a total <strong>of</strong> 68 putative ORs (Sakurai et al., 2004; Wanner et al., 2007;


General Introduction 15<br />

Tanaka et al., 2009), and tissue expression analysis has identified four ORs<br />

(BmOR19, BmOR30, BmOR45 and BmOR47) with higher expression levels <strong>in</strong><br />

female than male antennae (Anderson et al., 2009). In situ hybridisation has shown<br />

BmOR19 to be co-localised with either BmOR45 or BmOR47. Heterologous<br />

expression <strong>of</strong> three <strong>of</strong> these ORs <strong>in</strong> Sf9 <strong>in</strong>sect cells has identified BmOR19 to be<br />

tuned to l<strong>in</strong>alool, and BmOR45 and BmOR47 to be tuned to benzoic acid, 2-<br />

phenylethanol and benzaldehyde (Anderson et al., 2009).<br />

Mitsuno et al. (2008) identified ten ORs from three different moth species; Plutella<br />

xylostella (PxOR1, PxOR2, PxOR3 and PxOR4), Mythimna separate (MsOR1,<br />

MsOR2 and MsOR3), and Diaphania <strong>in</strong>dica (DiOR1, DiOR2 and DiOR3), through<br />

degenerate PCR <strong>of</strong> male antennal cDNA <strong>of</strong> the moths with primers designed from the<br />

conserved regions <strong>of</strong> BmOR1 and other male-specific ORs. Three <strong>of</strong> the ORs were<br />

homologues <strong>of</strong> the highly conserved BmOR2. Five <strong>of</strong> the ORs (PxOR1, PxOR4,<br />

DiOR1, MsOR1 and MsOR3) were expressed exclusively <strong>in</strong> male antennae, while<br />

two (PxOR3, DiOR3) were expressed <strong>in</strong> both male and female antennae, although<br />

higher levels were observed <strong>in</strong> male antennae. PxOR1, MsOR1 and DiOR1 were<br />

exclusively expressed <strong>in</strong> male antennae and shown to b<strong>in</strong>d sex pheromone<br />

component(s) <strong>of</strong> the respective moths (P. xylostella b<strong>in</strong>ds (Z)-11-hexadecenal; M.<br />

separate b<strong>in</strong>ds (Z)-11-hexadecenyl acetate and D. <strong>in</strong>dica b<strong>in</strong>ds (E)-11-Hexadecenal)<br />

(Mitsuno et al., 2008).<br />

M. sexta ORs were identified by differential screen<strong>in</strong>g <strong>of</strong> a male antennal cDNA<br />

library (OR1), degenerate PCR (OR2, which is the homologue <strong>of</strong> BmOR2) and from a<br />

subtracted male antennal transcriptome (OR3) (Patch et al., 2009). Tissue expression<br />

analysis revealed MsexOR1 to be expressed <strong>in</strong> male antennae and MsexOR3 to be<br />

expressed highly <strong>in</strong> female antennae. Three further ORs, named MsexOR4, MsexOR5<br />

and MsexOR6 have been identified <strong>in</strong> M. sexta recently from a subtractive male<br />

antennal cDNA library (Große-Wilde et al., 2010). MsexOR4 has been shown to be<br />

expressed <strong>in</strong> male antennae while MsexOR5 and MsexOR6 have been detected <strong>in</strong><br />

female antennae. Functional analysis <strong>of</strong> MsexOR1 with pheromone components did<br />

not reveal any ligands for this receptor as yet and MsexOR4 has been postulated to<br />

b<strong>in</strong>d one <strong>of</strong> the sex pheromone components <strong>of</strong> M. sexta however, no functional data<br />

for odorant b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> these ORs are available as yet.


General Introduction 16<br />

Ten ORs have been identified by degenerate PCR <strong>of</strong> male moth antennae from eight<br />

different Ostr<strong>in</strong>ia species. The B.mori homologues <strong>of</strong> BmOR1 and BmOR2 have been<br />

identified from O. scapulalis and O. latipennis, while only the BmOR1 homologue<br />

has been identified <strong>in</strong> O. furnacalis, O. palustralis, O. zaguliaevi, O. zealis, O.<br />

ovalipennis and O. nubilalis. A further five ORs (BmOR2 homologue and four more<br />

candidate PRs named OnOr3–6) have been identified from a male antennae EST<br />

library <strong>of</strong> O. nubilalis, br<strong>in</strong>g<strong>in</strong>g the total number <strong>of</strong> Ostr<strong>in</strong>ia ORs identified to date to<br />

15 (Wanner et al., 2010). Functional analysis <strong>of</strong> O. scapulalis OR1 (OscaOR1) and O.<br />

latipennis OR1 (OlatOR1) <strong>in</strong> X. laevis oocytes revealed them to b<strong>in</strong>d (E)-11-<br />

tetradecenol, a sex pheromone component <strong>of</strong> O. latipennis (Miura et al., 2009). O.<br />

nubulalis OR6 b<strong>in</strong>ds Z11-tetradecenyl acetate, a sex pheromone component <strong>of</strong><br />

Ostr<strong>in</strong>ia species, while OnOr1, 3 and 5 b<strong>in</strong>d four sex pheromonal components (Z11-<br />

and E11- tetradecenyl acetate, and Z12- and E12- tetradecenyl acetate) and the<br />

behavioural antagonist Z9-tetradecenyl acetate <strong>of</strong> Ostr<strong>in</strong>ia species (Wanner et al.,<br />

2010).<br />

In the giant silkmoth, A. polyphemus, a candidate PR (ApolOR1) has been identified<br />

from the screen<strong>in</strong>g <strong>of</strong> a cDNA library (Forstner et al., 2009). ApolOR1 is male<br />

antennae specific <strong>in</strong> its expression and found only <strong>in</strong> sensilla trichodea. Functional<br />

characterisation showed that it bound the pheromone components <strong>of</strong> A. polyphemus,<br />

(E,Z)-6,11-hexadecadienyl acetate, (E,Z)-6,11-hexadecadienal and (E,Z)-4,9-<br />

tetradecadienyl acetate (Meng et al., 1989; Forstner et al., 2009). A summary <strong>of</strong> the<br />

moth PRs and the sex pheromone component each b<strong>in</strong>ds is given <strong>in</strong> Table 1.1.


General Introduction 17<br />

Table 1.1: Summary <strong>of</strong> the moth pheromone receptors identified and characterised to<br />

date, together with their sex pheromone b<strong>in</strong>d<strong>in</strong>g components. Where more than one<br />

sex pheromone component is given, the component marked with * is the pheromone<br />

component bound by the correspond<strong>in</strong>g PR. O. nubulalis OR1, 3 and 5 b<strong>in</strong>d all the<br />

four correspond<strong>in</strong>g pheromone components given.<br />

Moth and<br />

Reference<br />

study<br />

H. virescens<br />

(Vetter and<br />

Baker, 1983)<br />

B. mori<br />

(Kaissl<strong>in</strong>g and<br />

Kasang, 1978)<br />

A.polyphemus<br />

(Meng et al.,<br />

1989)<br />

P. xylostella<br />

(Mitsuno et<br />

al., 2008)<br />

D. <strong>in</strong>dica<br />

(Mitsuno et<br />

al., 2008)<br />

M. separate<br />

(Mitsuno et<br />

al., 2008)<br />

O. scapulalis<br />

(Miura et al.,<br />

2009)<br />

O. latipennis<br />

(Miura et al.,<br />

2009)<br />

O. nubulalis OR1<br />

OR3<br />

OR5<br />

PR Sex pheromone component<br />

bound by correspond<strong>in</strong>g PR<br />

Percentage <strong>of</strong><br />

pheromone blend<br />

component<br />

HvOR6 (Z)-9-tetradecenal 5<br />

HvOR13 (Z)-11-hexadecenal 73<br />

HvOR14 (Z)-11-hexadecenyl acetate Produced by H. subflexa<br />

HvOR16 (Z)-11-hexadecen-1-ol 1<br />

BmOR1 Bombykol 90<br />

BmOR3 Bombykal 10<br />

ApolOR1 *(E,Z)-6,11-hexadecadienyl<br />

acetate<br />

90<br />

(E,Z)-6,11-hexadecadienal 10<br />

(E,Z)-4,9-tetradecadienyl acetate Trace amounts<br />

PxOR1 *(Z)-11-hexadecenal 50<br />

(Z)-11-hexadecenyl acetate 50<br />

(Z)-11-hexadecen-1-ol Trace amounts<br />

DiOR1 *(E)-11-Hexadecenal 66.1<br />

(E,E)-10,12-hexadecadienal 32.6<br />

MsOR1 *(Z)-11-hexadecenyl acetate 55.8<br />

(Z)-11-hexadecen-1-ol 29.2<br />

OR1 (E)-11-tetradecenol Not the sex pheromone<br />

for this moth but is for<br />

O. latipennis<br />

OR1 (E)-11-tetradecenol Unknown<br />

Z11- and E11- tetradecenyl<br />

acetate, and Z12- and E12tetradecenyl<br />

acetate<br />

OR6 Z11-tetradecenyl acetate<br />

97:3 (Z11-, Z12- : E11-,<br />

E12)


General Introduction 18<br />

1.5.2.2 Insect odorant receptor activation<br />

The novel receptor type <strong>in</strong> <strong>in</strong>sects, OR83b is a 7TM doma<strong>in</strong> prote<strong>in</strong> that is found co-<br />

expressed with the OR <strong>in</strong> almost all ORNs. Even though OR diversity with<strong>in</strong> and<br />

between <strong>in</strong>sect species is high, this receptor type is highly conserved <strong>in</strong> <strong>in</strong>sects such<br />

as D. melanogaster, B. mori, A. pernyi, H. virescens, Tenebrio molitor, Apis mellifera,<br />

Calliphora erythrocephala, P. xylostella, M. separate, D. <strong>in</strong>dica, and E. postvittana<br />

(Vosshall et al., 1999; Hill et al., 2002; Krieger et al., 2002; Larsson et al., 2004;<br />

Melo et al., 2004; Mitsuno et al., 2008; Jordan et al., 2009). The OR/OR83b<br />

heteromeric complex is formed <strong>in</strong> the cell body and it is thought that the OR83b<br />

functions <strong>in</strong> transport <strong>of</strong> the bound OR to the dendritic membrane <strong>in</strong> the sensilla. The<br />

complex is ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> the sensilla; suggest<strong>in</strong>g a role <strong>of</strong> OR83b as a co-receptor <strong>in</strong><br />

olfactory signall<strong>in</strong>g. Support<strong>in</strong>g evidence for this could be provided by the high<br />

sequence identity <strong>of</strong> this receptor type across different <strong>in</strong>sect species (64%–88%).<br />

Orthologs <strong>of</strong> OR83b <strong>in</strong> other <strong>in</strong>sect species have a conserved co-expression function<br />

and may be an important factor <strong>in</strong> odorant detection by conventional ORs <strong>in</strong> <strong>in</strong>sects<br />

(Larsson et al., 2004; Neuhaus et al., 2004; Nakagawa et al., 2005).<br />

The first <strong>in</strong>sect ORs were identified <strong>in</strong> Drosophila through homology search <strong>of</strong> the<br />

draft Drosophila genome us<strong>in</strong>g novel computer programs that identify mammalian<br />

ORs called G prote<strong>in</strong>-coupled receptor (GPCR) statistically by their physicochemical<br />

properties (Clyne et al., 1999; Vosshall et al., 1999). These algorithms were used to<br />

identify open read<strong>in</strong>g frames <strong>of</strong> two or more transmembrane doma<strong>in</strong> prote<strong>in</strong>s. Both<br />

GPCRs and <strong>in</strong>sect ORs have 7TM doma<strong>in</strong>s, however, <strong>in</strong>sect ORs have been<br />

demonstrated to have an <strong>in</strong>tracellular N-term<strong>in</strong>us, a topology unique to <strong>in</strong>sect ORs, as<br />

shown <strong>in</strong> Figure 1.4 (Benton et al., 2006). Lund<strong>in</strong> et al. (2007) used glycosylation<br />

scann<strong>in</strong>g to show the orientation and confirm the number <strong>of</strong> TM doma<strong>in</strong>s <strong>in</strong><br />

Drosophila OR83b. Further evidence for <strong>in</strong>sect OR topology came from epitope<br />

tagg<strong>in</strong>g <strong>of</strong> the term<strong>in</strong>i and the predicted loop regions <strong>of</strong> Drosophila OR22a (Smart et<br />

al., 2008). This unique orientation <strong>of</strong> <strong>in</strong>sect ORs suggests that these receptors would<br />

use dist<strong>in</strong>ct <strong>in</strong>sect-specific signall<strong>in</strong>g pathways.


General Introduction 19<br />

Figure 1.4: Topologies <strong>of</strong> <strong>in</strong>sect and mammalian odorant receptors <strong>in</strong> the cell<br />

membrane show<strong>in</strong>g opposite orientation <strong>of</strong> the N- and C- term<strong>in</strong>i <strong>in</strong> <strong>in</strong>sect OR as<br />

opposed to mammalian OR, as depicted by Benton et al. (2006), Lund<strong>in</strong> et al. (2007)<br />

and Smart et al. (2008).<br />

Recent evidence suggests a novel role <strong>of</strong> OR83b as an ion channel, that when co-<br />

expressed with a non-OR83b OR, confers ligand sensitivity (response shown by light<br />

blue arrow on Figure 1.5). Sato et al. (2008) heterologously expressed a number <strong>of</strong><br />

different <strong>in</strong>sect receptors <strong>in</strong> HeLa cells, together with the co-receptor OR83b and<br />

measured calcium <strong>in</strong>flux and whole-cell currents to analyse early onset <strong>of</strong> OR<br />

activation. They showed that the calcium <strong>in</strong>flux response <strong>of</strong> the <strong>in</strong>sect OR express<strong>in</strong>g<br />

cells to odorant stimulation is ten-fold faster than for mammalian ORs and no G<br />

prote<strong>in</strong>s are <strong>in</strong>volved <strong>in</strong> the signall<strong>in</strong>g, therefore none were <strong>in</strong>hibited, support<strong>in</strong>g<br />

similar f<strong>in</strong>d<strong>in</strong>gs by Smart et al. (2008). The response was also <strong>in</strong>dependent <strong>of</strong> second<br />

messengers such a cGMP and cAMP. The ion selectivity which is a property <strong>of</strong> ion<br />

channels, was dependent on the OR subunit composition; evidence suggest<strong>in</strong>g that<br />

ORs themselves act as ion channels rather than be<strong>in</strong>g associated with separate ion<br />

channels (Sato et al., 2008; Smart et al., 2008).<br />

A second hypothesis proposed by Wicher et al. (2008) suggests an ion channel–<br />

GPCR model (Figure 1.5). The ion channel model, similar to Sato et al. (2008)<br />

suggests OR83b form<strong>in</strong>g a channel complex with conventional ORs when co-<br />

expressed <strong>in</strong> HEK293 cells to evoke odour responses upon ligand b<strong>in</strong>d<strong>in</strong>g. The GPCR<br />

model shows that the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> a ligand to its OR/OR83b complex leads to the


General Introduction 20<br />

open<strong>in</strong>g <strong>of</strong> a cAMP dependent channel, suggest<strong>in</strong>g the odour evoked response also<br />

activates G prote<strong>in</strong>s. The activation <strong>of</strong> the OR complex occurs rapidly while the G<br />

prote<strong>in</strong> activation is slower and longer last<strong>in</strong>g (Wicher et al., 2008). No such evidence<br />

for a metabotropic signall<strong>in</strong>g mechanism has been observed <strong>in</strong> any other studies and<br />

further experiments have to be conducted to rectify the differences <strong>in</strong> these three<br />

studies.<br />

Figure 1.5: The ion channel–GPCR model <strong>of</strong> odorant signall<strong>in</strong>g <strong>in</strong> Drosophila. The<br />

ion channel formed by OR22a/OR83b complex depicts a rapid, short lived odorant<br />

evoked response upon stimulation by the ligand ethyl butyrate (light blue arrow) as<br />

suggested by Sato et al. (2008), Smart et al. (2008) and Wicher et al. (2008). The<br />

GPCR model suggests a slow and prolonged response upon ligand b<strong>in</strong>d<strong>in</strong>g to the<br />

OR22a/OR83b complex, lead<strong>in</strong>g to an <strong>in</strong>crease <strong>in</strong> cAMP production that opens up the<br />

motif that controls ion permeability <strong>in</strong> OR83b, thus result<strong>in</strong>g <strong>in</strong> membrane<br />

depolarisation (green arrow) as suggested by Wicher et al. (2008). This figure has<br />

been adapted from Wicher et al. (2008).<br />

ORs are a highly divergent group <strong>in</strong> that there is very little sequence homology both<br />

between and with<strong>in</strong> species, for example, D. melanogaster ORs share only 17%–26%<br />

am<strong>in</strong>o acid sequence identity (Vosshall et al., 2000); and the moth-specific<br />

pheromone receptor clade has an overall am<strong>in</strong>o acid sequence identity <strong>of</strong> 10% (Jordan<br />

et al., 2009). They are broadly tuned as one OR can b<strong>in</strong>d many odorants with different<br />

specificities and different ORs can b<strong>in</strong>d the same odorants. For example, B. mori<br />

OR45 b<strong>in</strong>ds five odorants, three <strong>of</strong> which are also ligands for BmOR47 (Anderson et<br />

al., 2009). Other studies <strong>in</strong> <strong>in</strong>sects support this broad tun<strong>in</strong>g <strong>of</strong> ORs (Malnic et al.,<br />

1999; Touhara et al., 1999; Malnic et al., 2004; Hallem and Carlson, 2006; Anderson


General Introduction 21<br />

et al., 2009; Jordan et al., 2009). The results <strong>in</strong> these studies show that the expression<br />

levels <strong>of</strong> ORs and PRs not only vary between the sexes but different ORs are<br />

expressed at different levels with<strong>in</strong> a specific tissue. A comb<strong>in</strong>atorial system is used<br />

by tens to hundreds <strong>of</strong> ORs to clearly detect and discrim<strong>in</strong>ate hundreds to thousands<br />

<strong>of</strong> different odorants. This comb<strong>in</strong>atorial function<strong>in</strong>g <strong>of</strong> ORs can be attributed to the<br />

organisation <strong>of</strong> the olfactory mach<strong>in</strong>ery, as different ORNs express<strong>in</strong>g the same ORs<br />

converge to the same glomerulus <strong>in</strong> the antennal lobe <strong>of</strong> the bra<strong>in</strong> (Vosshall et al.,<br />

2000; Couto et al., 2005; Fishilevich and Vosshall, 2005).<br />

1.6 Insect pest control strategies<br />

The problem <strong>of</strong> <strong>in</strong>sect pests has been ongo<strong>in</strong>g for centuries. They cause damage to<br />

food and crops <strong>in</strong> the field and <strong>in</strong> storage, and therefore measures need to be taken to<br />

br<strong>in</strong>g <strong>in</strong>sect pests under control and m<strong>in</strong>imise the damage and hence loss to the plant-<br />

based <strong>in</strong>dustries. Current <strong>in</strong>sect pest control strategies <strong>in</strong>clude spray<strong>in</strong>g food crops<br />

with pesticides, transgenic plants, as well as olfaction-based methods. These methods<br />

<strong>in</strong>clude lures to monitor pests <strong>in</strong> fields, mass trapp<strong>in</strong>g <strong>of</strong> <strong>in</strong>sects and kill<strong>in</strong>g them (lure<br />

and kill approach) and the use <strong>of</strong> mat<strong>in</strong>g disruption and use <strong>of</strong> biological controls<br />

(Jones, 1998; Suckl<strong>in</strong>g and Karg, 1999; Suckl<strong>in</strong>g et al., 1999).<br />

Control strategies us<strong>in</strong>g pheromones and plant volatiles have achieved success to<br />

some extent. Mat<strong>in</strong>g disruption reduces the number <strong>of</strong> successful mat<strong>in</strong>gs and hence<br />

the number <strong>of</strong> larvae that cause damage to plants (Jones, 1998). This system employs<br />

the release <strong>of</strong> sex pheromones <strong>in</strong> the atmosphere that masks the release <strong>of</strong> pheromone<br />

from call<strong>in</strong>g female <strong>in</strong>sects. This pheromone plume <strong>in</strong> the atmosphere confuses the<br />

males which are unable to detect the trail <strong>of</strong> pheromone towards the female thus<br />

reduces the number <strong>of</strong> successful mat<strong>in</strong>g. The success <strong>of</strong> this strategy is dependent on<br />

a number <strong>of</strong> factors. Knowledge <strong>of</strong> the time and place <strong>of</strong> mat<strong>in</strong>g will ensure that the<br />

pheromones are released at the right time and the number <strong>of</strong> generations <strong>of</strong> the target<br />

<strong>in</strong>sect <strong>in</strong> a year will also be <strong>in</strong>dicative <strong>of</strong> the number <strong>of</strong> pheromone applications that is<br />

needed. The life expectancy <strong>of</strong> the target <strong>in</strong>sects is an important factor also <strong>in</strong> that the<br />

shorter time the adults have, the better the rates <strong>of</strong> mat<strong>in</strong>g disruptions that can be<br />

achieved. The chances <strong>of</strong> an adult male <strong>in</strong> locat<strong>in</strong>g a female is less if the adult is short


General Introduction 22<br />

lived but it <strong>in</strong>creases if the adult is long lived (as they have more time <strong>in</strong> locat<strong>in</strong>g the<br />

female amidst the pheromone plumes) (Campion et al., 1989). This control method is<br />

susta<strong>in</strong>able as pheromones are non-toxic, non pollutant, naturally occurr<strong>in</strong>g<br />

compounds and can be used without caus<strong>in</strong>g harm to human health or the<br />

environment. S<strong>in</strong>ce they are species specific, they do not affect other non-target or<br />

beneficial <strong>in</strong>sects. This is not to say that there are no limitations to this method.<br />

Because pheromones are <strong>in</strong>sect specific and occur <strong>in</strong> specific ratios <strong>of</strong> the major and<br />

m<strong>in</strong>or components <strong>of</strong> the pheromone, the target range is limited (Campion et al.,<br />

1989). Its effectiveness is limited to just one particular species and different blends<br />

have to be synthesized for different target <strong>in</strong>sects. The tim<strong>in</strong>g <strong>of</strong> application has to be<br />

precise for it to be effective. The <strong>in</strong>sect density <strong>in</strong> the controlled area has to be<br />

monitored. Low density <strong>in</strong> a large area is preferred as a lower population <strong>of</strong> mat<strong>in</strong>g<br />

adults have less chance <strong>of</strong> locat<strong>in</strong>g mates <strong>in</strong> larger areas. If the control area has<br />

uncontrolled regions nearby, for example home gardens, then mated females from<br />

these areas can migrate and lay eggs <strong>in</strong> the controlled region. The larvae from these<br />

eggs will still cause damage to plants <strong>in</strong> the controlled region. F<strong>in</strong>ally, the cost<br />

effectiveness <strong>of</strong> such a system is questionable as pheromones are expensive to<br />

synthesise and are unstable volatiles that dis<strong>in</strong>tegrate easily <strong>in</strong> the environment<br />

(Bakke and Lie, 1989; Campion et al., 1989; Wall, 1989).<br />

Mass trapp<strong>in</strong>g <strong>of</strong> <strong>in</strong>sects is achieved by us<strong>in</strong>g lures such as sex pheromones and plant<br />

volatiles <strong>of</strong> host plants to direct <strong>in</strong>sects to traps (Haynes et al., 1986). Once trapped,<br />

the <strong>in</strong>sects are killed, either by tox<strong>in</strong>s <strong>in</strong> the trap or by depriv<strong>in</strong>g them <strong>of</strong> oxygen. The<br />

traps may also be coated with sticky substances to which the <strong>in</strong>sects get stuck and die.<br />

Mass trapp<strong>in</strong>g is done to reduce the population <strong>of</strong> <strong>in</strong>sects so that less mat<strong>in</strong>g pairs are<br />

available and the number <strong>of</strong> <strong>of</strong>fspr<strong>in</strong>g <strong>in</strong> the new generation is reduced (Bakke and<br />

Lie, 1989). The effectiveness <strong>of</strong> mass trapp<strong>in</strong>g is reduced <strong>in</strong> highly populated areas<br />

when the traps become full quickly. When sex pheromones are used as the lure, only<br />

male moths are caught, and the males that are not caught can fertilise more than one<br />

female hence a high proportion <strong>of</strong> the males have to be caught to render this method<br />

effective. Another approach that has been used <strong>in</strong> the control <strong>of</strong> <strong>in</strong>sect pests is lure<br />

and kill. A lure, such as sex pheromone that is able to attract <strong>in</strong>sects is associated with<br />

an <strong>in</strong>secticide that acts to kill the <strong>in</strong>sects. The effectiveness <strong>of</strong> this system depends on


General Introduction 23<br />

the attractiveness <strong>of</strong> the lure and the ability <strong>of</strong> the <strong>in</strong>secticide to <strong>in</strong>fect and kill the<br />

<strong>in</strong>sect (Haynes et al., 1986).<br />

One approach to overcome the shortfalls <strong>of</strong> mass trapp<strong>in</strong>g and lure and kill is lure and<br />

<strong>in</strong>fect (Suckl<strong>in</strong>g and Karg, 1999). This has been used on <strong>in</strong>sects such as tobacco<br />

budworm (Jackson et al., 1992), codl<strong>in</strong>g moth (Hrdy et al., 1996) and diamondback<br />

moth (Pell et al., 1993). This approach uses lures such as pheromones comb<strong>in</strong>ed with<br />

<strong>in</strong>sect pathogens such as virus, bacteria or fungi to attract <strong>in</strong>sects to the lure source<br />

(O‟Callaghan and Jackson., 1993). Once <strong>in</strong>fected, the <strong>in</strong>sects act as vectors <strong>of</strong> the<br />

disease and spread it across the population. An advantage <strong>of</strong> us<strong>in</strong>g such an approach<br />

is the speed with which the disease spreads <strong>in</strong> the population and with the <strong>in</strong>sect<br />

themselves act<strong>in</strong>g as vectors for the pathogen, the cont<strong>in</strong>uous need for the lure source<br />

to be ma<strong>in</strong>ta<strong>in</strong>ed and the number <strong>of</strong> lure sources is reduced (Suckl<strong>in</strong>g and Karg,<br />

1999). The downfall <strong>of</strong> this system is that bacterial tox<strong>in</strong>s and viruses are pathogenic<br />

only when consumed hence the lure system has to be designed such that the pathogen<br />

is consumed by the <strong>in</strong>sects. Also, for the pathogen to spread <strong>in</strong> the population, the<br />

<strong>in</strong>fected males have to mate to pass on the pathogen to the females, which must then<br />

be able to spread it to the surface <strong>of</strong> the eggs they lay, which will eventually have to<br />

be consumed by the larvae at eclosion. The success <strong>of</strong> the lure and <strong>in</strong>fect approach<br />

us<strong>in</strong>g bacteria or virus as the pathogen hence largely depend on the efficiency by<br />

which these series <strong>of</strong> events take place. When fungi is used as the pathogen, the lure<br />

has to be presented <strong>in</strong> such a way that it is able to attract the adult <strong>in</strong>sects and keep<br />

them with<strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the lure long enough for the fungi to be picked up. The<br />

adult has to be able to spread the fungi to other <strong>in</strong>sects <strong>in</strong> the population before dy<strong>in</strong>g<br />

or upon death; the fungi will grow on the dead <strong>in</strong>sect and spread with<strong>in</strong> the<br />

population. However, the fungi might be susceptible to environmental factors such as<br />

ultra violet (UV) radiation and die, a limit<strong>in</strong>g step to the success <strong>of</strong> this system.<br />

Economic production <strong>of</strong> the pathogen, ma<strong>in</strong>tenance <strong>of</strong> the pathogen <strong>in</strong> the<br />

environment, development <strong>of</strong> an efficient delivery system and ability <strong>of</strong> the pathogen<br />

to spread across the population are factors that have to be taken <strong>in</strong>to consideration<br />

when design<strong>in</strong>g lure and <strong>in</strong>fect control strategies (Suckl<strong>in</strong>g and Karg, 1999).<br />

Other methods <strong>of</strong> reduc<strong>in</strong>g the damage caused to plants are spray<strong>in</strong>g the host plants<br />

with feed<strong>in</strong>g and ovipostion deterrents. Oviposition deterrents will prevent females


General Introduction 24<br />

from lay<strong>in</strong>g eggs on host plants and the feed<strong>in</strong>g deterrents will render the plant<br />

<strong>in</strong>edible. The best approach as yet is the <strong>in</strong>tegrated pest management (IPM) system<br />

which was developed to <strong>in</strong>crease the use <strong>of</strong> biological resources (biological control<br />

and mat<strong>in</strong>g disruption) and supplement its use with reduced amounts <strong>of</strong> chemicals<br />

such as pesticides (Wear<strong>in</strong>g, 1993; Jones, 1998).<br />

1.7 Light brown apple moth<br />

The light brown apple moth, Epiphyas postvittana, is a species <strong>of</strong> leafroller moth, a<br />

member <strong>of</strong> the order Lepidoptera and belongs to the family Tortricidae (Wear<strong>in</strong>g et<br />

al., 1991). E. postvittana is endemic to Australia but has also <strong>in</strong>vaded Hawaii,<br />

California and parts <strong>of</strong> Europe (Suckl<strong>in</strong>g and Brockerh<strong>of</strong>f, 2010). The first report <strong>of</strong><br />

this moth <strong>in</strong> New Zealand was <strong>in</strong> 1891 (Danthanarayana, 1975). E. postvittana has a<br />

wide host range, from pipfruit (apples and pears) to citrus, grapes, kiwifruit and even<br />

some vegetables. E. postvittana attack pipfruit <strong>in</strong> all the fruit grow<strong>in</strong>g regions <strong>of</strong> New<br />

Zealand and is found <strong>in</strong> very high numbers <strong>in</strong> Nelson (Wear<strong>in</strong>g et al., 1991). The<br />

moth does not have a w<strong>in</strong>ter dormancy stage and the number <strong>of</strong> generations <strong>in</strong> a year<br />

differs <strong>in</strong> different regions <strong>of</strong> the country, with the far north hav<strong>in</strong>g four generations,<br />

central parts <strong>of</strong> the country hav<strong>in</strong>g three and two generations <strong>in</strong> the southern region. If<br />

the optimum temperature for larval development, 20°C is provided, it will take<br />

approximately 32–40 days for the completion <strong>of</strong> all the larval developmental stages.<br />

The larvae feed on leaves and fruits by creat<strong>in</strong>g silken shelters on either a s<strong>in</strong>gle leaf<br />

by roll<strong>in</strong>g or between a leaf and fruit (Geier and Briese, 1980). Pupae turn from a<br />

green to brown colour with maturation and male moth emergence occurs before<br />

females. The female moth mates only once <strong>in</strong> her life, and the eggs are laid <strong>in</strong> several<br />

batches <strong>of</strong> 30 eggs each so one female moth can lay between 150–900 eggs. Most <strong>of</strong><br />

these eggs hatch <strong>in</strong>to larvae as predation <strong>of</strong> the eggs is very low. Even though some<br />

biological control agents were imported from Australia, the zero tolerance<br />

requirement <strong>of</strong> this moth on exported fruits has still not been atta<strong>in</strong>ed (Wear<strong>in</strong>g et al.,<br />

1991).<br />

The damage caus<strong>in</strong>g stage <strong>of</strong> E. postvittana is the larva, which feeds on the buds,<br />

foliage, shoots and fruits (Suckl<strong>in</strong>g, 1998; Markwick et al., 2002). The roll<strong>in</strong>g <strong>of</strong>


General Introduction 25<br />

leaves and webb<strong>in</strong>g <strong>of</strong> leaves to fruits to make protective shelters causes calluses on<br />

the fruits and leaves (Lo et al., 2000). The export market requires nil tolerance <strong>of</strong> both<br />

<strong>in</strong>sect and <strong>in</strong>sect damage on fruits. These str<strong>in</strong>gent measures put <strong>in</strong> place calls for the<br />

development <strong>of</strong> better and susta<strong>in</strong>able control measures for this pest. From the 1960s,<br />

chemical control was used <strong>in</strong> the form <strong>of</strong> pesticide sprays. The major chemicals used<br />

were az<strong>in</strong>phos-methyl and chloropyrifos sprayed on fruits. However, these chemicals<br />

are toxic to some natural enemies <strong>of</strong> the moth and <strong>in</strong> 1980s, populations resistant to<br />

the chemicals were reported. This called for alternative control measures. Bacillus<br />

thur<strong>in</strong>giensis (B.t) was looked <strong>in</strong>to but due to high UV radiation <strong>in</strong> New Zealand, this<br />

microbial <strong>in</strong>secticide is easily deactivated (Suckl<strong>in</strong>g et al., 1994; Suckl<strong>in</strong>g et al.,<br />

1999; Markwick et al., 2002). The naturally occurr<strong>in</strong>g baculovirus enemy <strong>of</strong> the moth<br />

that kills the larvae, nucleopolyhedrovirus was used as sprays to some success.<br />

However, the most susta<strong>in</strong>able control came from the use <strong>of</strong> mat<strong>in</strong>g disruption and<br />

pheromone trapp<strong>in</strong>g <strong>of</strong> the adult moths. Small scale mat<strong>in</strong>g disruption showed that the<br />

best results were achieved through <strong>in</strong>tegrated pest management (IPM), by us<strong>in</strong>g the<br />

20:1 ratio <strong>of</strong> the sex pheromone components together with a few applications <strong>of</strong><br />

<strong>in</strong>secticide (Suckl<strong>in</strong>g and Shaw, 1990). Scale up <strong>of</strong> this mat<strong>in</strong>g disruption showed<br />

similar results <strong>in</strong> larger fields (Suckl<strong>in</strong>g and Shaw, 1995). The efficiency <strong>of</strong> this<br />

system is however h<strong>in</strong>dered by factors such as migration <strong>of</strong> mated females from<br />

nearby untreated sources, or a high population density requir<strong>in</strong>g a second application<br />

(McLaren et al., 1998). The biggest issue fac<strong>in</strong>g mat<strong>in</strong>g disruption is the high<br />

concentration <strong>of</strong> pheromone required hence the high cost <strong>of</strong> this pest control method.<br />

The limited knowledge <strong>of</strong> actual pheromone mechanism h<strong>in</strong>ders the success <strong>of</strong> this<br />

system. Understand<strong>in</strong>g the underly<strong>in</strong>g mechanisms <strong>of</strong> receptor adaptation to different<br />

compounds, the actual behaviour that mat<strong>in</strong>g disruption targets and whether males<br />

respond to any rema<strong>in</strong><strong>in</strong>g pheromone plumes will enhance its success. Reviewed <strong>in</strong><br />

Suckl<strong>in</strong>g and Brockerh<strong>of</strong>f (2010).<br />

The sex pheromone <strong>of</strong> E. postvittana is released at night or dur<strong>in</strong>g the dark period,<br />

and has been identified to be a 20:1 ratio <strong>of</strong> (E)-11-tetradecenyl acetate (E11-14:OAc)<br />

and (E,E)-9,11-tetradecadienyl acetate (E9,E11-14:OAc), respectively (Bellas et al.,<br />

1983) shown <strong>in</strong> Figure 1.6. The components on their own do not produce any<br />

response <strong>in</strong> male moths but when present <strong>in</strong> the specific ratio, the blend produced<br />

response <strong>in</strong> both field trials and laboratory bioassay. The biosynthetic pathways


General Introduction 26<br />

<strong>in</strong>volved <strong>in</strong> the production <strong>of</strong> the sex pheromone blend have been shown. The<br />

precursor, myristic acid underwent E11 desaturation event, followed by reduction and<br />

acetylation to give the pheromone component E11-14:OAc. E11 desaturation <strong>of</strong> the<br />

precursor palmitic acid followed by cha<strong>in</strong> shorten<strong>in</strong>g yielded E9-14:OAc and the<br />

diene resulted from further E11 desaturation (Foster and Roel<strong>of</strong>s, 1990). The<br />

pheromone production <strong>in</strong> female moths is regulated by pheromone biosynthesis<br />

activat<strong>in</strong>g neuropeptide (PBAN). The presence <strong>of</strong> sex pheromone <strong>in</strong> the female is low<br />

at eclosion, reaches its highest levels two days after eclosion and decreases to one<br />

third that <strong>of</strong> the peak at day seven, a gradual decl<strong>in</strong>e with the age <strong>of</strong> the moth (Foster,<br />

1993). Mat<strong>in</strong>g results <strong>in</strong> lower<strong>in</strong>g release <strong>of</strong> PBAN from females and thus stops the<br />

female from produc<strong>in</strong>g any more pheromone (Foster and Greenwood, 1997).<br />

Figure 1.6: The sex pheromone blend <strong>of</strong> E. postvittana is a 20:1 ratio <strong>of</strong> (E)-11tetradecenyl<br />

acetate and (E,E)-9,11-tetradecadienyl acetate (Bellas et al., 1983).


General Introduction 27<br />

Figure 1.7: Scann<strong>in</strong>g electronmicrograph <strong>of</strong> adult male (left) and female (right)<br />

antennaes <strong>of</strong> E.postvittana. The pheromone sensitive long sensilla trichodea are<br />

present <strong>in</strong> higher numbers on the male antennae than the female (shown with yellow<br />

arrow) suggest<strong>in</strong>g a role <strong>in</strong> sex pheromone reception <strong>of</strong> male moths (Jordon, 2006).<br />

Scale bars = 0.1mm.<br />

Pheromone and plant volatile recognition experiments <strong>in</strong> E. postvittana have been<br />

done us<strong>in</strong>g EAG response record<strong>in</strong>gs <strong>of</strong> antennae and ORNs (Rumbo, 1981; Karg et<br />

al., 1992; Suckl<strong>in</strong>g et al., 1996). The long sensilla trichodea localised <strong>in</strong> two radial<br />

rows on each male antennal segments are sensitive to pheromone components<br />

(Rumbo, 1981). This sensilla type is not present <strong>in</strong> the female antennae, as shown by<br />

scann<strong>in</strong>g electron microscopy (Figure 1.7). Experiments showed that each sensilla<br />

(<strong>in</strong>clud<strong>in</strong>g the pheromone sensitive sensilla) had three ORN cells associated with it,<br />

however, only two <strong>of</strong> these three were shown to be active <strong>in</strong> the pheromone specific<br />

sensillum (Rumbo, 1983). To determ<strong>in</strong>e what specific plant odorants the male and<br />

female moths respond to, EAG experiments and oviposition assays were carried out<br />

(Suckl<strong>in</strong>g et al., 1996). The odorants were chosen based on their presence <strong>in</strong> fruit and<br />

host plants <strong>of</strong> the moth (Table 1.2). These experiments revealed that eugenol, geraniol<br />

and citral acted as oviposition deterrents while hexanal, l<strong>in</strong>alool, nonanol, and octanol<br />

were attractants for the moth.


General Introduction 28<br />

Table 1.2: Plant volatiles detected by E. postvittana <strong>in</strong> electroantennogram<br />

experiments from (Suckl<strong>in</strong>g et al., 1996).<br />

Plant Volatiles<br />

Citral S-(-)-β-P<strong>in</strong>ene<br />

Z-3-Hexanol E-2-Hexenal<br />

Nonanol Octanol<br />

Phytol Hexanal<br />

α-Farnesene E-Caryophyllene<br />

Nonanal Heptanol<br />

Eugenol Thymol<br />

(+)-Limonene Heptanal<br />

L<strong>in</strong>alool (-)-α-Copaene<br />

Z-3-Hexyl acetate Acetophenone<br />

n-Decyl aldehyde Carvacrol<br />

Tetradecanol (-)-Limonene<br />

Geraniol α-Terp<strong>in</strong>eol<br />

Farnesol Benzaldehyde<br />

1-Indanone Nerolidol<br />

A molecular based approach has been pursued <strong>in</strong>to the odorant cod<strong>in</strong>g mechanism<br />

employed by this moth. Newcomb et al. (2002) isolated four OBPs from male and<br />

female antennae <strong>of</strong> E. postvittana. Native polyacrylamide gel electrophoresis (PAGE)<br />

separated four OBPs and the full sequence was then obta<strong>in</strong>ed by rapid amplification<br />

<strong>of</strong> cDNA ends polymerase cha<strong>in</strong> reaction (RACE-PCR) us<strong>in</strong>g primers based on N-<br />

term<strong>in</strong>al am<strong>in</strong>o acid sequenc<strong>in</strong>g (Newcomb et al., 2002). Two <strong>of</strong> the OBPs had<br />

sequence similarity to other lepidopteran PBPs, while two were GOBPs. The two E.<br />

postvittana PBPs differed by six am<strong>in</strong>o acid substitutions and were found on the same<br />

gene, hence they were two alleles <strong>of</strong> the same gene, however due to differential<br />

mobility‟s on native prote<strong>in</strong> gel, these were named PBP fast and PBP slow. Both these<br />

PBPs bound to radio-labelled major pheromone component E11-14:OAc <strong>in</strong> gel based<br />

assay and hence confirmed them as PBPs (Newcomb et al., 2002). Upon further<br />

sequence analysis, one <strong>of</strong> the GOBPs has been classified as a second PBP, PBP2.


General Introduction 29<br />

Expression and two-dimensional gel electrophoresis (2D-GE) separation <strong>of</strong> antennal<br />

prote<strong>in</strong>s revealed other soluble prote<strong>in</strong>s <strong>in</strong> the antennae <strong>of</strong> both the male and female<br />

moths. 26 prote<strong>in</strong>s were revealed, 17 <strong>of</strong> these prote<strong>in</strong>s were expressed more <strong>in</strong> male<br />

antennae than <strong>in</strong> female (Jordon, 2006). The creation and screen<strong>in</strong>g <strong>of</strong> a male E.<br />

postvittana antennal cDNA EST library revealed several chemosensory prote<strong>in</strong>s and<br />

takeout prote<strong>in</strong>s (Jordan et al., 2008; Hamiaux et al., 2009).<br />

Three putative olfactory receptors were identified from this EST database through<br />

phylogenetic analysis (Jordan et al., 2009). Full length sequences <strong>of</strong> these receptors<br />

was obta<strong>in</strong>ed via 5‟RACE and the am<strong>in</strong>o acid sequence identity <strong>of</strong> these receptors<br />

with other known <strong>in</strong>sect ORs was <strong>in</strong>vestigated. The receptor named EpOR1 has seven<br />

predicted transmenbrane doma<strong>in</strong>s, a cod<strong>in</strong>g region <strong>of</strong> 1245 nucleotides and a<br />

predicted prote<strong>in</strong> <strong>of</strong> 415 am<strong>in</strong>o acids. Homology search <strong>of</strong> this OR with other <strong>in</strong>sects<br />

showed that it had 36% sequence identity to B. mori OR1, a pheromone receptor<br />

(Figure 1.8). However, tissue expression analysis showed similar levels <strong>of</strong> expression<br />

<strong>in</strong> male and female antennae, <strong>in</strong>dicat<strong>in</strong>g that EpOR1 might not be a pheromone<br />

receptor; and functional analysis us<strong>in</strong>g calcium imag<strong>in</strong>g <strong>of</strong> this receptor <strong>in</strong> Spodoptera<br />

frugiperda 9 (Sf9) cells did not show any response to the E. postvittana sex<br />

pheromone. The N– and C– term<strong>in</strong>i <strong>of</strong> EpOR1 were c-Myc-epitope-tagged to<br />

determ<strong>in</strong>e the membrane topology (Jordan et al., 2009). The N-term<strong>in</strong>us was only<br />

accessible <strong>in</strong> the presence <strong>of</strong> sapon<strong>in</strong> while the C-term<strong>in</strong>us was accessible regardless<br />

<strong>of</strong> sapon<strong>in</strong> presence, <strong>in</strong>dicat<strong>in</strong>g an <strong>in</strong>tracellular N-term<strong>in</strong>us and an extracellular C-<br />

term<strong>in</strong>us, consistent with the membrane orientation observed for Drosophila Or83b<br />

and Or22a. The second putative receptor, EpOR2 has a cod<strong>in</strong>g region <strong>of</strong> 1422<br />

nucleotides and the predicted prote<strong>in</strong> is 474 am<strong>in</strong>o acids. This putative receptor has a<br />

long 3‟ untranslated region (UTR), as compared with EpOR1. Homology search<br />

revealed high sequence identity to the Drosophila co-receptor Or83b and 84%<br />

identity with BmOR2. Tissue expression analysis <strong>in</strong>dicated EpOR2 to be expressed <strong>in</strong><br />

both male and female antennae, with a higher level <strong>of</strong> expression <strong>in</strong> male antennae.<br />

Consistent with be<strong>in</strong>g the co-receptor <strong>in</strong> E. postvittana, EpOR2 is expressed about<br />

13–57 times more highly than EpOR1 and the third receptor, EpOR3. The third<br />

putative receptor EpOR3 has a cod<strong>in</strong>g region <strong>of</strong> 1230 nucleotides and a predicted<br />

prote<strong>in</strong> <strong>of</strong> 410 am<strong>in</strong>o acids. This receptor has 65% sequence identity with BmOR49J<br />

from B. mori (Anderson et al., 2009; Jordan et al., 2009) [<strong>in</strong> Anderson et al. (2009),


General Introduction 30<br />

this receptor is referred to as BmOR49 but <strong>in</strong> Tanaka et al. (2009), another OR has<br />

been annotated as BmOR49 so <strong>in</strong> this thesis, the BmOR49 <strong>of</strong> Anderson et al. (2009)<br />

will be referred to as BmOR49J], and tissue expression analysis showed it to be<br />

expressed <strong>in</strong> both male and female antennae at similar levels. Functional<br />

characterisation <strong>of</strong> EpOR3 by heterologous expression <strong>in</strong> Sf9 cells decoded its ligands<br />

to be members <strong>of</strong> terpenes, alcohols and esters at high concentrations, while at lower<br />

concentrations, the receptor bound citral, geraniol, geranial and geranyl acetate. Dose<br />

response curves showed EpOR3 has the highest aff<strong>in</strong>ity for the oviposition deterrent<br />

citral, with an EC50 value <strong>of</strong> 1.1 x 10 -13 M (Jordan et al., 2009), implicat<strong>in</strong>g a role <strong>of</strong><br />

this receptor as an OR and not PR. Three ORs have so far been identified from E.<br />

postvittana; EpOR2 is the Drosophila OR83b homologue, EpOR3 is the receptor for<br />

general plant odorants while EpOR1 rema<strong>in</strong>s to be functionally characterised. Further<br />

analysis <strong>of</strong> these receptors and the identification <strong>of</strong> new receptors will lead to a better<br />

understand<strong>in</strong>g <strong>of</strong> the olfactory mechanisms <strong>of</strong> this moth. This knowledge will aid <strong>in</strong><br />

the development <strong>of</strong> pest control strategies target<strong>in</strong>g the olfactory system <strong>of</strong> the moth.<br />

Two GOBP members have also been identified; however this class <strong>of</strong> soluble prote<strong>in</strong>s<br />

have very little functional data available, therefore an attempt will be made to deduce<br />

the role(s), if any, <strong>of</strong> this class <strong>of</strong> prote<strong>in</strong> <strong>in</strong> odour perception <strong>in</strong> <strong>in</strong> vitro assay<br />

systems.


General Introduction 31<br />

BmOr2<br />

HvOr2<br />

MsepOR2<br />

DiOR2<br />

PxOR2<br />

EpOR2<br />

0.1<br />

100<br />

*<br />

80<br />

HvOr15<br />

HvOr14<br />

MsepOR1<br />

HvOr6<br />

HvOr16<br />

HvOr11<br />

BmOr3<br />

PxOR1<br />

BmOr7<br />

BmOr5<br />

BmOr9<br />

BmOr4<br />

MsepOR3<br />

HvOr13<br />

DiOR3<br />

BmOr1<br />

DiOR1<br />

EpOR1<br />

PxOR3<br />

PxOR4<br />

Male-biased<br />

sex pheromone<br />

receptor clade<br />

31<br />

77<br />

99<br />

BmOr6<br />

100 BmOr22<br />

BmOr21<br />

HvOr19<br />

41 62<br />

99<br />

100<br />

98 HvOr21<br />

BmOr19<br />

BmOr17<br />

BmOr8<br />

BmOr20<br />

100<br />

43<br />

BmOr38<br />

BmOr35<br />

BmOr37<br />

96 HvOr12<br />

45<br />

41<br />

BmOr24<br />

69<br />

BmOr25<br />

99<br />

100<br />

88 89 HvOr7<br />

BmOr11<br />

HvOr9<br />

BmOr23<br />

BmOr42<br />

54<br />

100 HvOr18<br />

40<br />

100<br />

100<br />

HvOr20<br />

BmOr28<br />

57<br />

95<br />

BmOr14<br />

100 BmOr34<br />

100<br />

BmOr33<br />

BmOr30<br />

BmOr36<br />

34<br />

HvOr3<br />

100 BmOR22J<br />

100<br />

BmOr15<br />

99<br />

BmOr12<br />

30 73<br />

100<br />

100<br />

HvOr8<br />

BmOr13<br />

EpOR3 *<br />

EpOR3<br />

BmOR49 BmOR49J<br />

100<br />

100 BmOr48<br />

BmOr46 Female-biased<br />

100 BmOr47<br />

BmOr45 receptor clade<br />

100 BmOR19J<br />

94<br />

97<br />

HvOr10<br />

BmOr10<br />

40<br />

100<br />

BmOr16<br />

89 BmOR23J<br />

BmOr26<br />

HvOr17<br />

98<br />

BmOr29<br />

BmOr27<br />

Or83b clade<br />

98<br />

99<br />

48<br />

100<br />

60<br />

25 100<br />

Figure 1.8: Phylogenetic tree constructed from lepidopteran odorant receptors with<br />

the three E. postvittana ORs <strong>in</strong>dicated by *. Adapted from Jordan et al. (2009) where<br />

the tree was orig<strong>in</strong>ally constructed us<strong>in</strong>g the FITCH method from Jones-Thorton<br />

distances and the given bootstrap values calculated us<strong>in</strong>g 1000 replicates.<br />

50<br />

62<br />

96<br />

33<br />

100<br />

25<br />

34<br />

91<br />

100<br />

94<br />

100<br />

100<br />

100<br />

*


General Introduction 32<br />

1.8 Aims<br />

The implications <strong>of</strong> E. postvittana on the agricultural <strong>in</strong>dustries <strong>of</strong> New Zealand,<br />

Australia and California have <strong>in</strong>creased the <strong>in</strong>terest <strong>in</strong> study<strong>in</strong>g the mechanisms <strong>of</strong><br />

olfaction <strong>of</strong> this pest. In moths, pheromone receptors have been shown to be the<br />

receptors for species-specific sex pheromones and thus may play a crucial role <strong>in</strong><br />

male moth successfully perceiv<strong>in</strong>g sex pheromones released by call<strong>in</strong>g females.<br />

Odorant receptors on the other hand b<strong>in</strong>d a range <strong>of</strong> plant volatiles and might be<br />

<strong>in</strong>volved <strong>in</strong> host plant location <strong>in</strong> moths. With the crucial role that these prote<strong>in</strong>s play<br />

<strong>in</strong> the reproduction and survival <strong>of</strong> moths, a start<strong>in</strong>g po<strong>in</strong>t <strong>in</strong> combat<strong>in</strong>g these<br />

agricultural pests would be to identify and deorphan their ORs and PRs. The aims <strong>of</strong><br />

this study are to deorphan the b<strong>in</strong>d<strong>in</strong>g partners <strong>of</strong> EpOR1, which is phylogenetically<br />

related to PRs <strong>of</strong> other moth species; to <strong>in</strong>vestigate the role(s) <strong>of</strong> GOBP2 from E.<br />

postvittana <strong>in</strong> odorant recognition and to identify novel ORs and the PR(s) from E.<br />

postvittana.<br />

In chapter 2, EpOR1 is functionally characterised to identify its b<strong>in</strong>d<strong>in</strong>g partners;<br />

whether it b<strong>in</strong>ds the sex pheromone <strong>of</strong> E. postvittana predicted by its phylogenetic<br />

relatedness to PRs from other moth species or whether it b<strong>in</strong>ds semiochemicals <strong>of</strong><br />

general importance to the moth due to similar levels <strong>of</strong> expression <strong>in</strong> both male and<br />

female antennae. The b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> EpOR1 to its ligands and the relative sensitivity to<br />

different odorants is assessed by estimat<strong>in</strong>g EC50 values by obta<strong>in</strong><strong>in</strong>g dose response<br />

curves <strong>of</strong> ligand b<strong>in</strong>d<strong>in</strong>g.<br />

In chapter 3, the role <strong>of</strong> EpGOBP2 <strong>in</strong> ligand b<strong>in</strong>d<strong>in</strong>g is <strong>in</strong>vestigated. Current evidence<br />

suggests up to five roles <strong>of</strong> PBPs <strong>in</strong> odorant perception <strong>of</strong> moths, however, few if any<br />

tests have been conducted for the roles for GOBPs. The ability <strong>of</strong> recomb<strong>in</strong>ant<br />

EpGOBP2 to b<strong>in</strong>d the ligands <strong>of</strong> EpOR1 is tested, and common ligands for these two<br />

prote<strong>in</strong>s are established. The Sf9 cell assay system used for characteris<strong>in</strong>g EpOR1 is<br />

reconstituted with EpGOBP2 to test roles <strong>in</strong> odorant perception, <strong>in</strong>clud<strong>in</strong>g odorant<br />

solubilisation, transport and selectivity.<br />

In chapter 4 an attempt to <strong>in</strong>crease the receptor repertoire <strong>of</strong> E. postvittana is made. A<br />

total <strong>of</strong> 68 ORs have been identified from B. mori so far and this is used as the model


General Introduction 33<br />

for lepidopterans. Four different methods are compared to achieve this goal <strong>in</strong>clud<strong>in</strong>g<br />

microarray analysis, degenerate PCR, deep transcriptomics and low coverage whole<br />

genome sequenc<strong>in</strong>g. Questions on how the E. postvittana receptor repertoire<br />

correlates with that <strong>of</strong> B. mori receptor repertoire are addressed and tissue expression<br />

analysis <strong>of</strong> the ORs are conducted to identify any receptors that show sex-bias <strong>in</strong> their<br />

expression.<br />

F<strong>in</strong>ally the general discussion, synthesis<strong>in</strong>g the data obta<strong>in</strong>ed <strong>in</strong> this research with<br />

recent advances <strong>in</strong> lepidopteran olfaction is presented <strong>in</strong> chapter 5. A summary <strong>of</strong> the<br />

E. postvittana odorant receptors identified to date is presented, together with<br />

recommendations for future experiments that will confirm their tissue expression and<br />

deduce their roles <strong>in</strong> E. postvittana olfaction.


2.1 Introduction<br />

2<br />

Functional<br />

Characterisation <strong>of</strong><br />

Epiphyas postvittana<br />

Odorant Receptor 1<br />

<strong>Olfaction</strong>, or the sense <strong>of</strong> smell, plays a major role <strong>in</strong> the survival and reproduction <strong>of</strong><br />

moths whereby chemical signals present <strong>in</strong> their environment are used to<br />

communicate both with<strong>in</strong> and between species. Species-specific sex pheromones are<br />

released <strong>in</strong> the environment by females ready to mate and this is perceived by male<br />

moths over distances <strong>of</strong> up to 100 metres. Moths are also able to detect and<br />

discrim<strong>in</strong>ate a vast range <strong>of</strong> plant volatile compounds that guide them to food sources,<br />

<strong>in</strong>form them <strong>of</strong> plants already under herbivore attack and guide females to suitable<br />

oviposition sites (Honda, 1995; Bruce et al., 2005).<br />

ORs and PRs have been implicated as major players <strong>in</strong> odour detection for several<br />

moth species (Nakagawa et al., 2005; Große-Wilde et al., 2007; Mitsuno et al., 2008;<br />

Anderson et al., 2009; Forstner et al., 2009; Jordan et al., 2009; Miura et al., 2009).<br />

Genes encod<strong>in</strong>g moth ORs and PRs have been identified <strong>in</strong> a number <strong>of</strong> different


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 35<br />

species us<strong>in</strong>g degenerate PCR, transcriptome sequenc<strong>in</strong>g and whole genome<br />

sequenc<strong>in</strong>g techniques; E. postvittana (Jordan et al., 2009), A. polyphemus (Forstner<br />

et al., 2009), B. mori (Sakurai et al., 2004; Krieger et al., 2005; Nakagawa et al.,<br />

2005; Wanner et al., 2007; Tanaka et al., 2009), P. xylostella, M. separata, D. <strong>in</strong>dica<br />

(Mitsuno et al., 2008), H. virescens (Krieger et al., 2002; Krieger et al., 2004),<br />

Spodoptera exigua (Xiu et al., 2004) A. pernyi (Krieger et al., 2003), M. sexta (Patch<br />

et al., 2009; Große-Wilde et al., 2010) and eight different Ostr<strong>in</strong>ia species (Miura et<br />

al., 2009; Wanner et al., 2010). The majority <strong>of</strong> ORs from moths have been identified<br />

<strong>in</strong> B. mori and H. virescens.<br />

Typically <strong>in</strong> studies, the usual approach to characterise novel OR genes are to group<br />

them <strong>in</strong>to clades based on their phylogenetic relationships. The expression levels <strong>of</strong><br />

the ORs can be then tested <strong>in</strong> different tissues <strong>in</strong> male and female moths, those that<br />

show sex biased expression patterns are classified as putative pheromone receptors.<br />

De-orphan<strong>in</strong>g the receptors us<strong>in</strong>g available assay systems is carried out to confirm<br />

they are pheromone receptors and identify which components <strong>of</strong> sex pheromone<br />

blends they b<strong>in</strong>d. Pheromone receptors have so far been identified and functionally<br />

characterised for a number <strong>of</strong> moth species <strong>in</strong>clud<strong>in</strong>g B. mori, H. virescens, P.<br />

xylostella, M. separata, D. <strong>in</strong>dica, A. polyphemus, O. scapulalis, O. latipennis, and O.<br />

nubulalis, see section 1.5.2.1 (Nakagawa et al., 2005; Mitsuno et al., 2008; Forstner et<br />

al., 2009; Miura et al., 2009; Wang et al., 2010).<br />

Attempts to decode the olfactory systems <strong>of</strong> moths have seen the use <strong>of</strong> EAG to show<br />

the recognition <strong>of</strong> a vast range <strong>of</strong> important plant semiochemicals <strong>in</strong>clud<strong>in</strong>g esters,<br />

terpenes, alcohols, aliphatics and aldehydes by different species <strong>of</strong> moths. These may<br />

aid the <strong>in</strong>sects <strong>in</strong> locat<strong>in</strong>g oviposition sites and food sources (Suckl<strong>in</strong>g et al., 1996;<br />

Suckl<strong>in</strong>g et al., 1996; Fraser et al., 2003; Das et al., 2007). This method gives a broad<br />

picture <strong>of</strong> the wide range <strong>of</strong> volatiles recognised by the moth, however, <strong>in</strong>formation<br />

on the receptive range <strong>of</strong> <strong>in</strong>dividual ORs, which is at the forefront <strong>of</strong> <strong>in</strong>vestigat<strong>in</strong>g<br />

<strong>in</strong>to the molecular mechanisms <strong>of</strong> <strong>in</strong>sect olfaction is lack<strong>in</strong>g. Several characterisation<br />

assay systems have been successfully developed and used to confer functionality to<br />

<strong>in</strong>dividual ORs over the years. These <strong>in</strong>clude either the <strong>in</strong> vivo empty-neuron system<br />

developed <strong>in</strong> Drosophila (Stortkuhl and Kettler, 2001; Dobritsa et al., 2003) or <strong>in</strong><br />

vitro assays <strong>in</strong> heterologous cell-based expression systems (Wetzel et al., 2001;


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 36<br />

Neuhaus et al., 2004; Sakurai et al., 2004; Nakagawa et al., 2005; Kiely et al., 2007;<br />

Sato et al., 2008; Anderson et al., 2009; Jordan et al., 2009).<br />

The empty-neuron system, a mutant fly stra<strong>in</strong> that does not express the OR22a and<br />

OR22b receptors was orig<strong>in</strong>ally developed for characteris<strong>in</strong>g Drosophila ORs<br />

(Dobritsa et al., 2003).This approach has been successfully applied to the expression<br />

and characterisation <strong>of</strong> ORs from other <strong>in</strong>sects such as mosquitoes and moths (Syed et<br />

al., 2006; Kurtovic et al., 2007; Lu et al., 2007; Syed et al., 2010). Syed et al. (2006)<br />

successfully co-expressed functional B. mori OR1 with B. mori PBP <strong>in</strong> the<br />

Drosophila empty-neuron system and showed that it bound the B. mori sex<br />

pheromone, bombykol, albeit with low signal onset that lasted for an unusually long<br />

time, imply<strong>in</strong>g the empty-neuron express<strong>in</strong>g the BmOR1 is devoid <strong>of</strong> pheromone<br />

degrad<strong>in</strong>g enzymes that would otherwise degrade bombykol (Syed et al., 2006).<br />

Functional studies <strong>of</strong> moth ORs have ma<strong>in</strong>ly been carried out us<strong>in</strong>g heterologous<br />

expression systems. Several assays for characteris<strong>in</strong>g moth ORs have been carried<br />

out <strong>in</strong> X. laevis oocytes by two-electrode voltage clamp<strong>in</strong>g (Sakurai et al., 2004;<br />

Mitsuno et al., 2008; Miura et al., 2009). Briefly, the OR <strong>of</strong> <strong>in</strong>terest is co-expressed<br />

with the species-specific Drosophila OR83b homologue <strong>in</strong> X. laevis oocytes and the<br />

cells are left to grow for 2–3 days at 20°C. Two-electrode voltage clamps are used for<br />

record<strong>in</strong>g whole-cell currents which measures response upon stimulation with<br />

odorants as a variation from the hold<strong>in</strong>g voltage.<br />

Sakurai et al. (2004) expressed the B. mori pheromone receptor, BmOR1 together<br />

with BmGαq <strong>in</strong> X. laevis oocytes and showed that 10 to 15% <strong>of</strong> the transfected cells<br />

were responsive to the B. mori sex pheromone bombykol with an EC50 value <strong>of</strong> 3.4 x<br />

10 -5 M (Sakurai et al., 2004). In a similar experiment, the B. mori homologue <strong>of</strong><br />

OR83b, BmOR2 was co-expressed with BmOR1 and BmGαq <strong>in</strong> X. laevis oocytes<br />

(Nakagawa et al., 2005). Stimulation <strong>of</strong> the transfected cells with bombykol rendered<br />

~95% <strong>of</strong> the transfected cells responsive, with an EC50 value <strong>of</strong> 1.5 x 10 -6 M and the<br />

lowest threshold concentration for response <strong>of</strong> 1 x 10 -7 M, <strong>in</strong>dicat<strong>in</strong>g a role <strong>of</strong> OR83b<br />

<strong>in</strong> enhanc<strong>in</strong>g the sensitivity <strong>of</strong> ORs.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 37<br />

Miura et al. (2009) co-expressed O. scapulalis OR1 (OscaOR1) together with its<br />

species-specific OR83b homologue <strong>in</strong> X. laevis oocytes and showed that it responded<br />

highly to E11-14:OH and a small response was shown to Z11-16:OAc (Miura et al.,<br />

2009). These compounds have been identified as pheromone components <strong>of</strong> other<br />

Ostr<strong>in</strong>ia species, while the sex pheromone components <strong>of</strong> O. scapulalis (E11- and<br />

Z11- 14:OAc) did not confer a response from OscaOR1, <strong>in</strong>dicat<strong>in</strong>g that even though<br />

OscaOR1 is homologous to the pheromone receptors <strong>of</strong> Ostr<strong>in</strong>ia species, it is not<br />

specific to O. scapulalis sex pheromone. Wanner et al. (2010) expressed Ostr<strong>in</strong>ia<br />

nubilalis OR1, and ORs 3–6 <strong>in</strong> X. laevis oocytes and showed that they bound all the<br />

components <strong>of</strong> the Ostr<strong>in</strong>ia sex pheromone, as shown <strong>in</strong> Table 1.1 (Wanner et al.,<br />

2010).<br />

Mitsuno et al. (2008) expressed the pheromone receptor together with their species<br />

specific OR83b homologue from three different moth species (P. xylostella, M.<br />

separata and D. <strong>in</strong>dica) <strong>in</strong> X. laevis oocytes and showed via electrophysiological<br />

record<strong>in</strong>gs that they all bound their respective sex pheromone components with an<br />

EC50 <strong>in</strong> the micromolar range and the lowest b<strong>in</strong>d<strong>in</strong>g threshold <strong>of</strong> 10 -7 M, values<br />

which are comparable with B. mori PR assays conducted by Nakagawa et al. (2005).<br />

Modified versions <strong>of</strong> HEK293 cells, conta<strong>in</strong><strong>in</strong>g a mouse Gα15 gene have also been<br />

used for successfully express<strong>in</strong>g and characteris<strong>in</strong>g moth ORs <strong>in</strong> fluorometric assays.<br />

Stable cell l<strong>in</strong>es <strong>of</strong> H. virescens HvOR13, HvOR14 and HvOR16 have been generated<br />

and characterised to b<strong>in</strong>d H. virescens sex pheromone components <strong>in</strong> HEK cells with<br />

high aff<strong>in</strong>ity. HvOR13 can recognise Z11-16:Al <strong>in</strong> DMSO with an EC50 <strong>of</strong> 1.2 x 10 -9<br />

M (Große-Wilde et al., 2007). These results are comparable to an earlier study <strong>in</strong><br />

which B. mori OR1 and OR3 were expressed and characterised <strong>in</strong> HEK293 cells.<br />

BmOR1 bound the sex pheromone bombykol with an EC50 <strong>of</strong> 10 -10 M (Große-Wilde<br />

et al., 2006). Together these results show that expression <strong>of</strong> moth ORs <strong>in</strong> HEK293<br />

cells yield functional ORs, that are highly responsive and sensitive to their ligands.<br />

Kiely et al. (2006) developed an <strong>in</strong>sect cell system for assay<strong>in</strong>g <strong>in</strong>sect ORs.<br />

Drosophila OR22a was transiently expressed <strong>in</strong> Sf9 cells, and after a 48 hour<br />

<strong>in</strong>cubation period, the cells were loaded with the calcium sensitive dye Fluo-4 and<br />

change <strong>in</strong> fluorescence <strong>of</strong> cells <strong>in</strong> response to stimulation by odorants was measured


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 38<br />

<strong>in</strong> a Leitz Fluovert FS <strong>in</strong>verted microscope (Kiely et al., 2007). OR22a was shown to<br />

respond to ethyl butyrate and other odorants previously characterised by <strong>in</strong> vivo<br />

assays (Dobritsa et al., 2003; Hallem et al., 2004; Hallem and Carlson, 2006). Us<strong>in</strong>g<br />

<strong>in</strong>sect cell l<strong>in</strong>es would provide the best conditions for assay<strong>in</strong>g <strong>in</strong>sect ORs as the<br />

heterologously expressed ORs are able to efficiently couple with the Sf9 cell signal<br />

transduction pathway. Evidence for this can be seen <strong>in</strong> the threshold concentrations <strong>of</strong><br />

ethyl butyrate that is required to elicit a response <strong>in</strong> OR22a. When OR22a is<br />

expressed <strong>in</strong> HEK293 cells, a high concentration <strong>of</strong> 10 -3 M <strong>of</strong> ethyl butyrate is<br />

required to elicit a response (Neuhaus et al., 2004). This threshold concentration is<br />

decreased by 10 8 -folds to 10 -12 M when OR22a is expressed <strong>in</strong> Sf9 cells (Kiely et al.,<br />

2007). This provides confidence that an <strong>in</strong>sect cell assay system will be able to<br />

provide a better environment for assay<strong>in</strong>g moth receptors and functional<br />

characterisation assays for moth ORs us<strong>in</strong>g this system has shown to confer high<br />

sensitivity to the ORs for their respective ligands. Four female biased B. mori ORs<br />

were expressed and tested with a panel <strong>of</strong> odorants <strong>of</strong> importance to the moth <strong>in</strong> the<br />

Sf9 system. Two <strong>of</strong> these, BmOR45 and BmOR47 recognised the compounds 2-<br />

phenyl ethanol and benzoic acid respectively to concentrations as low as 10 -14 M,<br />

implicat<strong>in</strong>g the high sensitivity <strong>of</strong> the system, while BmOR30 did not respond to any<br />

<strong>of</strong> the tested compounds and BmOR19 responded to l<strong>in</strong>alool with an EC50 <strong>of</strong> 4.69 x<br />

10 -9 M (Anderson et al., 2009). E. postvittana OR3 has also been de-orphaned us<strong>in</strong>g<br />

the Sf9 system; it has been shown to recognise the oviposition deterrent <strong>of</strong> the moth,<br />

the terpene citral to concentrations as low as 10 -15 M (Jordan et al., 2009). Unlike the<br />

X. laevis oocyte and HEK293 cell assay systems mentioned above, the Sf9 cell assay<br />

system does not require the co-expression <strong>of</strong> exogenous factors like OR83b and Gα15<br />

prote<strong>in</strong>. This is likely due to the presence <strong>of</strong> an endogenous homologue <strong>of</strong> Drosophila<br />

OR83b <strong>in</strong> Sf9 cells as detected by RT-PCR (Smart et al., 2008).<br />

Of the three ORs identified from E. postvittana, multiple sequence alignment and<br />

phylogenetic analysis <strong>of</strong> EpOR1 with five other moth species placed EpOR1 <strong>in</strong> a<br />

receptor clade that has 19 other OR members albeit with low homology (an overall<br />

receptor am<strong>in</strong>o acid identity <strong>of</strong> 10% <strong>in</strong> the clade, with 32–40% identity with<br />

<strong>in</strong>dividual receptor sequences <strong>in</strong> the clade), as shown <strong>in</strong> Figure 1.8 (Jordan et al.,<br />

2009). Some <strong>of</strong> these member ORs have been characterised for their expression<br />

patterns <strong>in</strong> male and female moth antennae and many have been shown to be male


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 39<br />

biased, be<strong>in</strong>g highly expressed <strong>in</strong> male antennae with low or no expression <strong>in</strong> female<br />

antennae. Functional odorant b<strong>in</strong>d<strong>in</strong>g data is available for some <strong>of</strong> these ORs,<br />

show<strong>in</strong>g tun<strong>in</strong>g <strong>of</strong> the ORs to the female sex pheromone components <strong>of</strong> the respective<br />

species. The phylogenetic relatedness <strong>of</strong> EpOR1 to PRs from other moths make it a<br />

candidate PR <strong>of</strong> E. postvittana. Quantitative real time polymerase cha<strong>in</strong> reaction<br />

(qRT-PCR) <strong>of</strong> EpOR1 was carried out on different male and female moth tissues and<br />

the results showed that the expression was restricted to the antennae. Interest<strong>in</strong>gly, no<br />

sex-specific bias <strong>in</strong> expression levels <strong>of</strong> EpOR1 was observed, an <strong>in</strong>dication that<br />

EpOR1 was perhaps like HvOR6, BmOR9, PxOR3 and DiOR3, members <strong>of</strong> the clade<br />

that do not have sex specific expression and may not respond to sex pheromone<br />

components.<br />

2.1.1 Aims<br />

The aim <strong>of</strong> this research chapter is to deorphan EpOR1, to determ<strong>in</strong>e if EpOR1 is a<br />

PR or an OR. If it is an OR then what compounds does it recognise and how does it<br />

b<strong>in</strong>d the different compounds that it recognises? Is EpOR1 broadly or narrowly tuned,<br />

recognis<strong>in</strong>g a few or many compounds? How sensitive is EpOR1 to its ligands, is it<br />

able to recognise low concentrations <strong>of</strong> odorants? To address these questions, EpOR1<br />

is expressed and functionally characterised <strong>in</strong> the Sf9 cell assay system us<strong>in</strong>g calcium<br />

imag<strong>in</strong>g as it does not require the employment <strong>of</strong> exogenous factors for functional<br />

expression <strong>of</strong> the receptor. EpOR1 is tested with the major pheromone component <strong>of</strong><br />

the sex pheromone blend <strong>of</strong> E. postvittana, as well as with a range <strong>of</strong> plant<br />

semiochemicals <strong>of</strong> importance to the moth, as shown <strong>in</strong> behavioural studies and <strong>in</strong><br />

whole antennae EAG experiments (Suckl<strong>in</strong>g et al., 1996).<br />

2.2 Materials and methods<br />

2.2.1 Materials<br />

Stock odorants were made at a concentration <strong>of</strong> 10 -2 M <strong>in</strong> DMSO and stored at -20°C<br />

until required. The odorants used are as follows: -p<strong>in</strong>ene (99%, racemic, BDH),<br />

citral (95%, Sigma, mixture <strong>of</strong> neral (36%) and geranial (64%)), geraniol (97%,


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 40<br />

BDH), geranial (96.1%, synthesised by the method <strong>of</strong> Dess and Mart<strong>in</strong> (1983),<br />

geranyl acetate (98%, Sigma), nerol (97%, Sigma), l<strong>in</strong>alool (97%, racemic, Sigma),<br />

(+)-limonene (97%, Sigma), 1,8 c<strong>in</strong>eole (90%, Fluka, Ronkonkoma, NY), myrcene<br />

(Sigma), -terp<strong>in</strong>eol (98%, Sigma), -farnesene (Sigma), caryophyllene (Koch-light<br />

labs, Cambridge, MA), citronellol (95%, racemic, Sigma), humulene (ABD), octanol<br />

(99.5%, Fluka), nonanol (98%, Fluka), hexanol (98%, Fluka), hexanal (98%, Sigma),<br />

hexyl acetate (99%, Sigma), methyl salicylate (99%, Sigma), eugenol (BDH), ethyl<br />

butyrate (Hopk<strong>in</strong> and Williams, Chadwell Health, Essex, UK), ethyl hexanoate (99%,<br />

Sigma), pentyl acetate (99%, Sigma), dodecyl acetate (97%, Sigma), squalene (99%,<br />

Sigma), pentanol (99.8%, Fluka), octanol (99.5%, Fluka), ethyl octanoate (99%,<br />

Sigma), propyl propanoate (99.7%, Fluka), octyl acetate (99%, Sigma), propyl acetate<br />

(99.7%, Fluka) and (E)-11-tetradecenyl acetate (97%, Sigma). Nerol, pentyl acetate,<br />

octanol, geranyl acetate and methyl salicylate were stored at RT, while all rema<strong>in</strong><strong>in</strong>g<br />

compounds were stored at 4°C.<br />

The assay sal<strong>in</strong>e buffer (pH 7.2) conta<strong>in</strong>ed 21 mM KCl, 12 mM NaCl, 18 mM MgCl2,<br />

3 mM CaCl2.H2O, 170 mM d-glucose, 1 mM probenecid (Sigma–Aldrich) and 10<br />

mM PIPES-dipotassium salt. The assay sal<strong>in</strong>e buffer was sterilised by filtration<br />

through a 0.22µM SteriCup–GP filter unit (Millipore) and stored at room temperature.<br />

2.2.2 Insect cell culture<br />

S. frugiperda Sf9 cells (Invitrogen) were ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> shaker cultures <strong>in</strong> 50 mL<br />

flasks at 28°C <strong>in</strong> the dark. The cells were grown to a density <strong>of</strong> 1 x 10 7 cells/mL <strong>in</strong> Sf<br />

900 II serum free media (Invitrogen) after which they were seeded <strong>in</strong> new flasks at 1 x<br />

10 5 cells/mL.<br />

2.2.3 Transfection <strong>of</strong> cells<br />

Viable cells were seeded at a density <strong>of</strong> 1 x 10 6 cells/mL <strong>in</strong> 12 well tissue culture<br />

Nunclone plates with 400 µL Sf 900 II media. 0.5 µg pIB-EpOR1 (N-myc-tagged<br />

EpOR1 <strong>in</strong> pIB-V5/His vector, constructed by Melissa Jordan) or the empty vector<br />

control pIB-V5/His was <strong>in</strong>cubated with 12 µL <strong>of</strong> the transfection reagent escort IV<br />

(Sigma) and 100 µL <strong>of</strong> Sf 900 II media for each well at room temperature for 15


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 41<br />

m<strong>in</strong>utes. This mixture was added to the well with the Sf9 cells and <strong>in</strong>cubated at 27°C<br />

<strong>in</strong> the dark for 7 hours. The cells were then washed and topped up with fresh media<br />

and left to grow for a further 48 hours.<br />

2.2.4 Detection <strong>of</strong> mRNA encod<strong>in</strong>g pIB-EpOR1 by RT-PCR<br />

A s<strong>in</strong>gle well <strong>of</strong> each <strong>of</strong> the transfected cells (with pIB-V5/His or pIB-EpOR1) was<br />

harvested 48 hours post transfection by remov<strong>in</strong>g the media and wash<strong>in</strong>g the cells<br />

twice with 2 mL phosphate buffered sal<strong>in</strong>e (PBS). The cells were resuspended <strong>in</strong> 1<br />

mL PBS, and pelleted by centrifugation at 5000g. The cells were stored at -80°C till<br />

required. Total RNA was extracted us<strong>in</strong>g TRIzol reagent (Invitrogen) follow<strong>in</strong>g<br />

manufacturer‟s protocol, with the RNA resuspended <strong>in</strong> 10 µL diethylpyrocarbonate-<br />

treated (DEPC-treated) water. Five microlitres <strong>of</strong> the RNA was used for synthesiz<strong>in</strong>g<br />

cDNA us<strong>in</strong>g iScript cDNA synthesis kit (Bio-Rad) follow<strong>in</strong>g the manufacturer‟s<br />

protocol. A one <strong>in</strong> ten dilution <strong>of</strong> the cDNA was made and 2 µL <strong>of</strong> this was used <strong>in</strong> a<br />

PCR reaction together with 1x PCR buffer <strong>in</strong>clud<strong>in</strong>g 1.5 mM magnesium<br />

(Invitrogen), 0.2 mM <strong>of</strong> each dNTP (Invitrogen), 2 units <strong>of</strong> Taq DNA polymerase<br />

(Invitrogen) and 0.5 µM <strong>of</strong> each primer. The f<strong>in</strong>al volume was made up to 20 µL with<br />

sterile water. The forward and reverse primers used were EpOR1 gene specific<br />

primers, EpOR1 forward– 5‟ATGGATGTATTCAATTTAAAATACATGCGA3‟ and<br />

EpOR1 reverse– 5‟TCACTGATTTGCAAATGTTCTCAGCATCAG3‟, which<br />

amplify the full length cod<strong>in</strong>g region <strong>of</strong> the EpOR1 cDNA (1248bp). The PCR<br />

cycl<strong>in</strong>g conditions were as follows: <strong>in</strong>itial denaturation at 94°C for two m<strong>in</strong>utes,<br />

followed by 35 cycles <strong>of</strong> 94°C for 20 seconds, 55°C for 30 seconds and 72°C for one<br />

m<strong>in</strong>ute, and a f<strong>in</strong>al extension at 72°C for ten m<strong>in</strong>utes. PCR products were analysed on<br />

1% agarose gel sta<strong>in</strong>ed with 1x SYBR safe DNA gel sta<strong>in</strong> (Invitrogen) and images<br />

captured on an ImageQuant 300 system (GE Healthcare Life Sciences).<br />

2.2.5 Membrane Fraction Isolation and Western Blot<br />

The method <strong>of</strong> isolation <strong>of</strong> membrane fractions was adapted from Bordier (1980). To<br />

pre-condense Triton X-114, 5 mL <strong>of</strong> Triton X-114 was mixed with 250 mL PBS and<br />

cooled to 0°C. The mixture was then heated at 37°C until the phases separated. The<br />

top aqueous phase was discarded and 250 mL <strong>of</strong> PBS was added to the bottom phase


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 42<br />

and left on ice for 30 m<strong>in</strong>utes. The mixture was heated at 37°C until phase separation<br />

occurred and the aqueous phase was removed completely. The rema<strong>in</strong><strong>in</strong>g lower phase<br />

after removal <strong>of</strong> the aqueous phase was the pre-condensed Triton X-114 and this was<br />

diluted to 2% (<strong>in</strong> PBS) for use <strong>in</strong> membrane extraction.<br />

For extraction <strong>of</strong> the membrane prote<strong>in</strong>s, the transfected Sf9 cells were washed twice<br />

with ice cold PBS. The cells were then dislodged <strong>in</strong> 400 µL PBS, spun down at 3000g<br />

for three m<strong>in</strong>utes and the pellet resuspended <strong>in</strong> 200 µL Lysis buffer (2% pre-<br />

condensed Triton X-114 made up <strong>in</strong> PBS). The cells were sonicated (2x 30 second<br />

passes), <strong>in</strong>cubated on ice for at least one hour and then spun at 13000g at 4°C for ten<br />

m<strong>in</strong>utes to remove <strong>in</strong>soluble debris. The supernatant was then transferred to a new<br />

tube, <strong>in</strong>cubated at 37°C for ten m<strong>in</strong>utes, followed by centrifugation at 13000g at 25ºC<br />

for ten m<strong>in</strong>utes to separate phases. The aqueous phase was discarded and the bottom<br />

layer was mixed with 1 mL <strong>of</strong> fresh cold PBS and <strong>in</strong>cubated at 0°C for ten m<strong>in</strong>utes.<br />

The mixture was heated at 37°C for ten m<strong>in</strong>utes to separate phases and centrifuged at<br />

13000g at 25ºC for ten m<strong>in</strong>utes. The aqueous phase was aga<strong>in</strong> discarded and the<br />

wash<strong>in</strong>g steps repeated twice. The prote<strong>in</strong> was resuspended <strong>in</strong> 25 µL <strong>of</strong> the follow<strong>in</strong>g<br />

mix for runn<strong>in</strong>g on a western blot: 6.25 µL <strong>of</strong> NuPage LDS Buffer, 2.5 µL <strong>of</strong> NuPage<br />

sample reduc<strong>in</strong>g agent, 16.25 µL <strong>of</strong> distil water and 6 M urea. It was then heated at<br />

37°C for 20 m<strong>in</strong>utes and spun at 13000g for 30 seconds to remove any particulates.<br />

Ten microlitres was loaded on a NuPage 4–12% Bis-Tris gel (Invitrogen) and run at<br />

200 volts for 35 m<strong>in</strong>utes. The iBlot Dry Blott<strong>in</strong>g System (Invitrogen) was used to<br />

transfer prote<strong>in</strong> from the gel to a nitrocellulose transfer membrane follow<strong>in</strong>g<br />

manufacturer‟s <strong>in</strong>structions. 20 mL <strong>of</strong> 1x milk powder <strong>in</strong> 1x PBS with 0.05% Tween<br />

20 was used for block<strong>in</strong>g the membrane for two hours. The block<strong>in</strong>g buffer was<br />

washed <strong>of</strong>f with 1x PBS 0.05% Tween 20 and the membrane <strong>in</strong>cubated with a one <strong>in</strong><br />

1000 dilution <strong>in</strong> 1x PBS with 0.05% Tween 20 with the primary antibody, anti-myc<br />

mouse antibody (Roche) for one and half hours. The antibody solution was washed<br />

<strong>of</strong>f and the membrane washed three times with 1x PBS conta<strong>in</strong><strong>in</strong>g 0.05% Tween 20.<br />

The membrane was then <strong>in</strong>cubated with secondary antibody, goat-anti mouse IgGiAP<br />

(Stressgen, diluted one <strong>in</strong> 2000) for one hour. The secondary antibody solution was<br />

washed <strong>of</strong>f and 1-Step NBT-BCIP (Nitro blue tetrazolium chloride 5-Bromo-4-<br />

chloro-3-<strong>in</strong>dolyl phosphate, toluid<strong>in</strong>e salt, Pierce) substrate was added to visualise<br />

bands on the membrane.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 43<br />

2.2.6 EpOR1 functional assay<br />

Functional assay <strong>of</strong> pIB-EpOR1 transfected Sf9 cells was carried out follow<strong>in</strong>g the<br />

method <strong>of</strong> Kiely et al. (2007). Briefly, 48 hours after transfection with EpOR1, the<br />

cells were washed with assay buffer. This was done by tilt<strong>in</strong>g the plate, remov<strong>in</strong>g the<br />

media with a pipette and add<strong>in</strong>g 500 µL <strong>of</strong> assay buffer conta<strong>in</strong><strong>in</strong>g 2 μM Fluo-4<br />

acetoxymethyl ester (Invitrogen) and 0.01% pluronic acid F–127 (Sigma-Aldrich) to<br />

each well. The plate was <strong>in</strong>cubated <strong>in</strong> the dark at 28°C for 20 m<strong>in</strong>utes after which it<br />

was washed with assay buffer to remove traces <strong>of</strong> any rema<strong>in</strong><strong>in</strong>g dye. 400 µL <strong>of</strong> fresh<br />

assay buffer was added to the wells and a second <strong>in</strong>cubation at 28°C was done for 20<br />

m<strong>in</strong>utes. Calcium imag<strong>in</strong>g <strong>of</strong> the Fluo-4 loaded cells was carried out on a Leitz<br />

Fluovert FS <strong>in</strong>verted fluorescence microscope fitted with a MicroMax CCD camera<br />

system (model RTE/CCD-1300-Y/HS; Pr<strong>in</strong>ceton Instruments). An I3 filter set<br />

(excitation filter 450-490 nm, dichroic mirror 510 nm, long pass emission filter 520<br />

nm) was used to image the Fluo-4 labelled cells. MetaFluor v5.0 (Universal Imag<strong>in</strong>g<br />

Corporation) was used to operate the camera and control the image acquisition.<br />

Transfected cells were tested with each <strong>of</strong> the odorants listed <strong>in</strong> section 2.2.1.<br />

Odorants were made up to 0.1 M <strong>in</strong> DMSO and then further diluted <strong>in</strong> assay buffer.<br />

Dose response data were collected for compounds that elicited consistent responses at<br />

concentrations below 10 -5 M. Three wells on each 12 well Nunclone plate were tested<br />

with the same concentration, and all the plates tested with a particular compound on a<br />

given day were transfected us<strong>in</strong>g the same stock transfection mix.<br />

To carry out calcium imag<strong>in</strong>g <strong>of</strong> the cells, a field <strong>of</strong> view with evenly spread healthy<br />

cells was selected <strong>in</strong> each well with the 10x objective lens. Six <strong>in</strong>itial images were<br />

taken at ten second <strong>in</strong>tervals, then 50 µL <strong>of</strong> the assay buffer was added (to control for<br />

cells that auto-fluoresce upon the addition <strong>of</strong> a solvent) and another six images<br />

acquired. 50 µL <strong>of</strong> the test compound was then added, six further images were<br />

acquired and f<strong>in</strong>ally 50 µL <strong>of</strong> the ionophore, ionomyc<strong>in</strong> (Sigma) was added to a f<strong>in</strong>al<br />

concentration <strong>of</strong> 10 µM to estimate the maximum possible fluorescence <strong>of</strong> the cells.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 44<br />

2.2.7 Data analysis<br />

The data collected from the calcium imag<strong>in</strong>g was analysed us<strong>in</strong>g MetaFluor Analyst<br />

v5.1 (Universal Imag<strong>in</strong>g Corporation). Circles were drawn with the region <strong>of</strong> <strong>in</strong>terest<br />

(ROI) tool around the perimeter <strong>of</strong> the cells <strong>in</strong> the field <strong>of</strong> view and also three<br />

background regions free <strong>of</strong> cells. The fluorescence data for the selected ROIs over the<br />

image acquisition period was obta<strong>in</strong>ed with the „calculate plot‟ function <strong>of</strong> the<br />

s<strong>of</strong>tware. The data from the three background ROIs was averaged and the ROI data<br />

from the cells was subtracted from this background, to m<strong>in</strong>imise differences <strong>in</strong> cell<br />

image backgrounds.<br />

The change <strong>in</strong> fluorescence (∆F) <strong>of</strong> the cells that elicit a response upon odorant<br />

stimulation was calculated as follows:<br />

Where,<br />

FBlank is the average fluorescence <strong>in</strong>tensity <strong>of</strong> the cell, from the first image to the<br />

twelfth image before addition <strong>of</strong> the ligand;<br />

FSubstrate is the highest fluorescence <strong>in</strong>tensity <strong>of</strong> the cell after the addition <strong>of</strong> the ligand,<br />

from the thirteenth to the eighteenth image, and<br />

FMax is the maximum fluorescence <strong>of</strong> the cell achieved after the addition <strong>of</strong> the<br />

ionophore, from image n<strong>in</strong>eteen to twenty-four.<br />

The average ∆F <strong>of</strong> all the cells respond<strong>in</strong>g to a particular ligand concentration was<br />

take as the mean ∆F for that concentration (with a m<strong>in</strong>imum threshold <strong>of</strong> at least 3<br />

respond<strong>in</strong>g cells) when plott<strong>in</strong>g the dose response curve us<strong>in</strong>g Orig<strong>in</strong>Pro v7.5<br />

(Orig<strong>in</strong>Lab) s<strong>of</strong>tware. Cells that respond to stimulation with assay buffer only were<br />

discarded from the analysis. EC50 (concentration at which 50% <strong>of</strong> the maximal effect<br />

or activation <strong>of</strong> the ligand is observed) and Hill‟s slope (slope <strong>of</strong> the dose response<br />

curve that gives the level <strong>of</strong> cooperativity <strong>of</strong> the ligands) (Motulsky and<br />

Christopoulos, 2003) were calculated us<strong>in</strong>g Graphpad Prism s<strong>of</strong>tware. A compound<br />

was considered unable to activate EpOR1 if no response was observed <strong>in</strong> cells upon<br />

stimulation with the odorant <strong>in</strong> triplicates at concentration <strong>of</strong> 10 -5 M.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 45<br />

2.3 Results<br />

To determ<strong>in</strong>e if the transfection <strong>of</strong> pIB-EpOR1 <strong>in</strong>to Sf9 cells had been successful,<br />

reverse transcription polymerase cha<strong>in</strong> reaction (RT-PCR) was performed on mRNA<br />

extracted from the transfected cells us<strong>in</strong>g primers specific for the full length cod<strong>in</strong>g<br />

region <strong>of</strong> the EpOR1 gene (Figure 2.1A). As a control, cells transfected with the<br />

pIB/V5 empty vector were also tested (Figure 2.1B). The empty vector transfected<br />

cells did not show the presence <strong>of</strong> EpOR1 transcripts, while the EpOR1-transfected<br />

cells showed the presence <strong>of</strong> PCR products <strong>of</strong> the expected size, 1245bp.<br />

Figure 2.1: Reverse transcription PCR analysis <strong>of</strong> transfected pIB-EpOR1 Sf9 cells.<br />

A. RT-PCR with myc-EpOR1 specific primers <strong>of</strong> pIB-EpOR1 transfected Sf9 cells 48<br />

hours post transfection. B. RT-PCR <strong>of</strong> Sf9 cells transfected with the pIB/V5, an<br />

empty vector control, with myc-EpOR1 specific primers.<br />

To confirm the expression <strong>of</strong> myc-EpOR1 prote<strong>in</strong> <strong>in</strong> transfected Sf9 cells, a western<br />

blot <strong>of</strong> the membrane fraction from these cells was probed with an anti-myc antibody.<br />

Three bands <strong>of</strong> approximately 25 kDa, 50 kDa and 60 kDa were observed on the<br />

western blot (Figure 2.2). The band „b‟ observed at approximately 50 kDa<br />

corresponds to the theoretical molecular weight <strong>of</strong> myc-tagged EpOR1. The higher<br />

band „a‟ could be aggregation <strong>of</strong> the prote<strong>in</strong> that did not separate out on the gel or a


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 46<br />

glycosylated form <strong>of</strong> the receptor prote<strong>in</strong> while the lower band observed at „c‟ could<br />

be degraded prote<strong>in</strong> or background noise, as seen <strong>in</strong> the vector only lane 3.<br />

Figure 2.2: Western blot <strong>of</strong> myc-tagged EpOR1 <strong>in</strong> Sf9 cells transfected with pIB-<br />

EpOR1. Lane 1 is the magic marker western ladder (Invitrogen), lane 2 is the pIB-<br />

EpOR1 prote<strong>in</strong> membrane fraction and lane 3 is the pIB-V5/His prote<strong>in</strong> membrane<br />

fraction from cells transfected with vector only. Three bands <strong>of</strong> 60 (a), 50 (b) and 25<br />

(c) kDa were observed <strong>in</strong> lane 2, the 50 kDa band correspond<strong>in</strong>g to the molecular<br />

weight <strong>of</strong> myc-EpOR1.<br />

2.3.1 EpOR1 functional characterisation <strong>in</strong> Sf9 cells<br />

In a typical calcium assay experiment, no change <strong>in</strong> fluorescence <strong>in</strong>tensity <strong>of</strong> the Sf9<br />

cells is observed upon addition <strong>of</strong> the control sal<strong>in</strong>e solution. An <strong>in</strong>crease <strong>in</strong> the<br />

<strong>in</strong>tracellular calcium levels, which is measured as an <strong>in</strong>crease <strong>in</strong> fluorescence <strong>of</strong> the<br />

pIB-EpOR1 express<strong>in</strong>g cells is observed upon the addition <strong>of</strong> the ligand, and the<br />

maximum fluorescence <strong>in</strong>tensity <strong>of</strong> the cells is achieved upon addition <strong>of</strong> the<br />

ionophore, ionomyc<strong>in</strong> as depicted <strong>in</strong> Figure 2.3. Sf9 cells transfected with the empty


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 47<br />

pIB-V5/His vector were also tested <strong>in</strong> a dose-dependent manner with the methyl<br />

salicylate and the data obta<strong>in</strong>ed is given <strong>in</strong> Appendix A.<br />

Figure 2.3: The change <strong>in</strong> response <strong>of</strong> a s<strong>in</strong>gle Sf9 cell transfected with pIB-EpOR1<br />

over the timecourse <strong>of</strong> a calcium assay experiment. Arrow on the images <strong>in</strong>dicates the<br />

change <strong>in</strong> fluorescence <strong>in</strong>tensity <strong>of</strong> the respond<strong>in</strong>g cell plotted on the graph over the<br />

course <strong>of</strong> the assay: six <strong>in</strong>itial images are acquired followed by the addition <strong>of</strong> the<br />

control assay sal<strong>in</strong>e buffer represented <strong>in</strong> frames 6–12, then 10 -5 M citral is added and<br />

six more images are acquired (frames 12–18). Frames 18–24 are acquired after the<br />

addition <strong>of</strong> the ionophore, ionomyc<strong>in</strong>.<br />

Of the 33 compounds tested (Table 2.1), ten putative ligands were identified for<br />

EpOR1, based on the ability <strong>of</strong> the compounds to elicit a calcium response at a<br />

concentration <strong>of</strong> 10 -5 M <strong>in</strong> Sf9 cells transfected with pIB-EpOR1. These putative<br />

ligands belong to a wide range <strong>of</strong> chemical groups rang<strong>in</strong>g from monoterpenes<br />

through to alcohols, esters and benzoates. The average normalised change <strong>in</strong><br />

fluorescence <strong>of</strong> the cells upon <strong>in</strong>teraction with each ligand at 10 -5 M was calculated<br />

and is given <strong>in</strong> Table 2.1.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 48<br />

Table 2.1: Change <strong>in</strong> fluorescence at 10 -5 M <strong>of</strong> the compounds tested with EpOR1 <strong>in</strong><br />

the Sf9 cell assay. Each odorant was tested <strong>in</strong> triplicate from a s<strong>in</strong>gle transfection mix.<br />

Compound<br />

Group<br />

Compound Name ∆F a N c<br />

Monoterpene Limonene NR b 0<br />

Myrcene NR 0<br />

-p<strong>in</strong>ene NR 0<br />

L<strong>in</strong>alool NR 0<br />

Citronellol NR 0<br />

-terp<strong>in</strong>eol NR 0<br />

1,8 c<strong>in</strong>eole 0.43 ± 0.25 6<br />

Nerol 0.41 ± 0.14 8<br />

Geraniol 0.31 ± 0.03 8<br />

Citral 0.32 ± 0.02 7<br />

Geranial 0.36 ± 0.02 6<br />

Geranyl acetate 0.32 ± 0.02 6<br />

Sesquiterpene -farnesene 0.41 ± 0.12 7<br />

Caryophyllene NR 0<br />

Humulene NR 0<br />

Triterpene Squalene NR 0<br />

Alcohol Hexanol NR 0<br />

Nonanol NR 0<br />

Pentanol NR 0<br />

Octanol 0.25 ± 0.21 6<br />

Aldehyde Hexanal NR 0<br />

Ester Ethyl butyrate NR 0<br />

Ethyl octanoate NR 0<br />

Ethyl hexanoate NR 0<br />

Propyl propanoate NR 0<br />

Pentyl acetate 0.37 ± 0.31 8<br />

Hexyl acetate NR 0<br />

Dodecyl acetate NR 0<br />

Octyl acetate NR 0<br />

Propyl acetate NR 0<br />

(E)-11-tetradecenyl acetate NR 0<br />

Benzoate Eugenol NR 0<br />

Methyl salicylate 0.31 ± 0.03 6<br />

a. Average change <strong>in</strong> fluorescence ± SE <strong>of</strong> at least three cells.<br />

b. NR = no response as determ<strong>in</strong>ed by no change <strong>in</strong> fluorescence <strong>of</strong> the cells<br />

upon stimulation with the odorant <strong>in</strong> triplicates at 10 -5 M.<br />

c. N = number <strong>of</strong> respond<strong>in</strong>g cells. A m<strong>in</strong>imum <strong>of</strong> 300 cells were looked at to<br />

confirm that a compound did not elicit a calcium response <strong>in</strong> the cells.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 49<br />

To exam<strong>in</strong>e the sensitivity <strong>of</strong> EpOR1 for the ten respond<strong>in</strong>g ligands, calcium assays<br />

were conducted over a range <strong>of</strong> ligand concentrations to enable the construction <strong>of</strong><br />

dose response curves. Ligands were tested at ten-fold decreas<strong>in</strong>g concentrations<br />

start<strong>in</strong>g at 10 -5 M. Only five <strong>of</strong> the ten compounds (methyl salicylate, geranyl acetate,<br />

citral, geraniol and geranial) elicited a calcium response at concentrations lower than<br />

10 -5 M and the compounds differed <strong>in</strong> their levels <strong>of</strong> sensitivity as shown <strong>in</strong> Figure<br />

2.4.<br />

A relative measure <strong>of</strong> the difference between the b<strong>in</strong>d<strong>in</strong>g aff<strong>in</strong>ities <strong>of</strong> EpOR1 to the<br />

different ligands can be represented by compar<strong>in</strong>g estimated EC50 values, the<br />

concentration at which 50% <strong>of</strong> the maximal effect or activation <strong>of</strong> the ligand is<br />

observed (Motulsky and Christopoulos, 2003). The level <strong>of</strong> cooperativity <strong>of</strong> the<br />

ligands is given by the slope <strong>of</strong> the dose response curve and is calculated as the Hill<br />

slope. The lowest concentration <strong>of</strong> ligand to which a response <strong>of</strong> the cells was<br />

observed varied between the five best ligands and is given <strong>in</strong> Table 2.2, together with<br />

the EC50 and Hill slopes from the dose response curves.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 50<br />

A<br />

C D<br />

E<br />

Figure 2.4: Dose response curves <strong>of</strong> the change <strong>in</strong> fluorescence exhibited by Sf9 cells<br />

transfected with pIB-EpOR1 to (A) methyl salicylate, (B) geranyl acetate, (C) citral<br />

(D) geraniol and (E) and geranial. The change <strong>in</strong> fluorescence <strong>of</strong> the pIB-EpOR1<br />

transfected cells elicited by a range <strong>of</strong> ligand concentrations was calculated and the<br />

data represented by a sigmoidal curve for each ligand. Error bars represent the<br />

standard error <strong>of</strong> six to n<strong>in</strong>e respond<strong>in</strong>g cells.<br />

B


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 51<br />

Table 2.2: EC50 (± standard error) and Hill slope estimates <strong>of</strong> the dose response<br />

curves <strong>of</strong> five best ligands for EpOR1. The recognition threshold is the lowest<br />

concentration to which EpOR1 transfected Sf9 cells elicited a measurable response to<br />

the test compound.<br />

Compound EC50 (M) Hill slope Recognition threshold (M)<br />

Methyl salicylate<br />

Geraniol<br />

Citral<br />

Geranial<br />

Geranyl acetate<br />

2.4 Discussion<br />

(1.8 ± 0.95) x 10 -12<br />

(5.8 ± 0.35) x 10 -11<br />

(1.3 ± 2.5) x 10 -9<br />

(5.3 ± 2.5) x 10 -8<br />

(2.8 ± 0.4) x 10 -8<br />

0.35<br />

3.2<br />

0.33<br />

0.61<br />

0.83<br />

10 -15<br />

10 -12<br />

10 -10<br />

10 -11<br />

10 -10<br />

Us<strong>in</strong>g an <strong>in</strong>sect cell assay system, E. postvittana OR1 was functionally characterised<br />

to be a receptor for general plant odorants, <strong>in</strong>clud<strong>in</strong>g citral, geraniol, geranial, methyl<br />

salicylate, geranyl acetate, pentyl acetate, octanol, α-farnesene, nerol and 1,4-c<strong>in</strong>eole.<br />

The presence <strong>of</strong> EpOR1 at the mRNA and prote<strong>in</strong> levels <strong>in</strong> transfected Sf9 cells<br />

(detected by RT-PCR and western blot respectively) suggests that EpOR1 is be<strong>in</strong>g<br />

transcribed and translated <strong>in</strong> the Sf9 cells giv<strong>in</strong>g confidence that a functional assay is<br />

possible. Western blot <strong>of</strong> membrane fraction extracted from Sf9 cells transfected with<br />

pIB-EpOR1 showed three bands, one at approximately 25 kDa, the second at 50 kDa<br />

and the third at 60 kDa. The 50 kDa band corresponds closely to the theoretical<br />

molecular weight <strong>of</strong> myc-tagged EpOR1 at 49.6 kDa, while the higher 60 kDa band<br />

could be a glycosylated form <strong>of</strong> the prote<strong>in</strong>, as has been postulated by Henn<strong>in</strong>gsen et<br />

al. (2002) for multiple bands on immunoblots <strong>of</strong> human histam<strong>in</strong>e H2 receptor. The<br />

lowest band could be degraded prote<strong>in</strong> or nonspecific b<strong>in</strong>d<strong>in</strong>g as the pIB-V5/His<br />

transfected Sf9 cell membrane prote<strong>in</strong> fraction (lane 3) also shows non-specific<br />

b<strong>in</strong>d<strong>in</strong>g around 25 kDa.<br />

When the pIB-EpOR1 transfected Sf9 cells were stimulated with 10 -5 M citral, as<br />

shown <strong>in</strong> Figure 2.3, not all the cells showed an <strong>in</strong>crease <strong>in</strong> fluorescence. This is due


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 52<br />

to the low transfection rate <strong>of</strong> the ORs <strong>in</strong> Sf9 cells, a major drawback <strong>of</strong> this assay<br />

system (Kiely, 2008). The expression <strong>of</strong> the OR by the Sf9 cells is random, with only<br />

about 5–10% <strong>of</strong> the cells be<strong>in</strong>g transfected and the percentage <strong>of</strong> these transfected<br />

cells <strong>in</strong> the chosen field <strong>of</strong> view can vary anywhere between 0 to 100%. Therefore, a<br />

large number <strong>of</strong> assays have to be conducted to obta<strong>in</strong> mean<strong>in</strong>gful data as well as<br />

assign a test odorant as a responder or a non-responder. In Figure 2.3, not all the cells<br />

show the same fluorescent levels upon the addition <strong>of</strong> the ionophore, ionomyc<strong>in</strong>. This<br />

is due to the differential uptake <strong>of</strong> Fluo-4 by the Sf9 cells, therefore each cell <strong>in</strong> the<br />

field <strong>of</strong> view is selected manually and analysed us<strong>in</strong>g a semi-automated excel<br />

algorithm (Kiely, 2008).<br />

E. postvittana OR1 is activated by ten <strong>of</strong> the 33 compounds it was tested with. These<br />

compounds were chosen either due to their ability to elicit an EAG response <strong>in</strong> whole<br />

antennae <strong>of</strong> E. postvittana (Suckl<strong>in</strong>g et al., 1996), or their occurrence as plant<br />

semiochemicals. The major sex pheromone component <strong>of</strong> E. postvittana, E-11-<br />

tetradecenyl acetate, was also tested but no response was obta<strong>in</strong>ed, suggest<strong>in</strong>g a role<br />

<strong>of</strong> this receptor as a general OR and not a PR. This conclusion is consistent with the<br />

qRT-PCR data obta<strong>in</strong>ed <strong>in</strong> Jordan et al. (2009) which showed no sex bias <strong>of</strong> this<br />

receptor <strong>in</strong> male and female antennal tissue. To date, all moth PRs that have been<br />

characterised show male biased expression (Krieger et al., 2004; Sakurai et al., 2004;<br />

Mitsuno et al., 2008; Forstner et al., 2009; Patch et al., 2009). These two results are,<br />

however, <strong>in</strong> contrast with phylogenetic data where EpOR1 belongs to a cluster <strong>of</strong><br />

mostly sexually dimorphic ORs from five other moths, most <strong>of</strong> which are male biased<br />

PRs (Figure 1.8) (Krieger et al., 2004; Sakurai et al., 2004; Mitsuno et al., 2008;<br />

Jordan et al., 2009). The am<strong>in</strong>o acid identity with<strong>in</strong> this clade is only about 13%, with<br />

EpOR1 hav<strong>in</strong>g an average <strong>of</strong> 35% sequence identity with <strong>in</strong>dividual OR sequences.<br />

EpOR1 has the highest am<strong>in</strong>o acid sequence identity (40%) with D. <strong>in</strong>dica OR1<br />

(DiOR1), which is a member <strong>of</strong> Lepidoptera:Crambidae. Two-electrode voltage<br />

clamp record<strong>in</strong>gs <strong>of</strong> X. laevis oocytes transiently express<strong>in</strong>g DiOR1 show this<br />

receptor to be most sensitive to the major sex pheromone component for this moth,<br />

E11-16: aldehyde (Mitsuno et al., 2008). Even though EpOR1 and DiOR1 are closely<br />

related phylogenetically, the ligands they b<strong>in</strong>d are very different imply<strong>in</strong>g that<br />

phylogenetic relatedness does not <strong>in</strong>dicate functional conservation <strong>in</strong> evolutionary<br />

diverse moths.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 53<br />

EpOR1 is highly sensitive, recognis<strong>in</strong>g odorants at very low concentrations. For<br />

example, EpOR1 can recognise methyl salicylate at a concentration as low as 10 -15 M,<br />

with an EC50 <strong>of</strong> 1.8 x 10 -12 M when transiently expressed <strong>in</strong> Sf9 cells (Figure 2.4A).<br />

This sensitivity <strong>of</strong> <strong>in</strong>sect ORs has been shown <strong>in</strong> other studies <strong>of</strong> ORs expressed <strong>in</strong><br />

Sf9 cell l<strong>in</strong>es. Jordan et al. (2009) has shown that another OR from E. postvittana,<br />

EpOR3 b<strong>in</strong>ds citral at concentrations as low as 10 -15 M when expressed <strong>in</strong> Sf9 cells.<br />

Two other studies showed high sensitivity <strong>of</strong> ORs to their respective ligands when<br />

expressed <strong>in</strong> Sf9 cells. Kiely et al. (2007) showed D. melanogaster OR22a can<br />

respond to ethyl butyrate at the low concentration <strong>of</strong> 10 -13 M with an EC50 <strong>of</strong> 1.58 x<br />

10 -11 M, while Anderson et al. (2009) showed the B. mori female specific OR,<br />

BmOR47 can respond to benzoic acid at 10 -14 M with an EC50 <strong>of</strong> 1.42 x 10 -11 M. This<br />

sensitivity <strong>of</strong> <strong>in</strong>sect ORs is ma<strong>in</strong>ta<strong>in</strong>ed even when the transiently express<strong>in</strong>g cell l<strong>in</strong>e<br />

is <strong>of</strong> non-<strong>in</strong>sect orig<strong>in</strong>. B. mori OR1 expressed <strong>in</strong> HEK293 cells b<strong>in</strong>d bombykol to<br />

concentrations <strong>of</strong> 10 -12 M with EC50 <strong>of</strong> 10 -10 M (Große-Wilde et al., 2006) while the<br />

pheromone receptor from H. virescens HvOR13 b<strong>in</strong>d to the major sex pheromone<br />

component <strong>of</strong> this moth <strong>in</strong> the presence <strong>of</strong> pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> at a<br />

concentration <strong>of</strong> 10 -14 M with an EC50 <strong>of</strong> 2 x 10 -13 M (Große-Wilde et al., 2007).<br />

However, while highly sensitive ORs with different k<strong>in</strong>etics for various odorants are<br />

seen <strong>in</strong> the Sf9 cell assay systems <strong>in</strong> both moths and flies, the response pr<strong>of</strong>ile might<br />

be very different <strong>in</strong> a biological context. Inferences on how the concentration <strong>of</strong><br />

certa<strong>in</strong> odorants affect the behaviour <strong>of</strong> the moth <strong>in</strong> the environment cannot be fully<br />

justified based on this <strong>in</strong> vitro data and behavioural studies need to be conducted to<br />

reconcile these differences.<br />

The dose response <strong>of</strong> EpOR1 to different odorants can either be „gently slop<strong>in</strong>g‟<br />

where EpOR1 response <strong>in</strong>creases gradually with the concentration <strong>of</strong> the compound,<br />

(for example, the dose response curves <strong>of</strong> methyl salicylate, citral, geranyl acetate and<br />

geranial show this effect; these gradual change can be attributed as a concentration<br />

effect; as the concentration <strong>of</strong> the ligand <strong>in</strong>creases, the fluorescence <strong>of</strong> the cells<br />

<strong>in</strong>creases similarly); or „abrupt‟ where the odorant acts like a „switch‟, turn<strong>in</strong>g the OR<br />

„on/<strong>of</strong>f‟ at a particular concentration (for example, there is a sharp <strong>in</strong>crease <strong>in</strong> the<br />

b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> geraniol to EpOR1 from 10 -11 M to 10 -10 M <strong>in</strong> Figure 2.4D. It appears that a<br />

concentration <strong>of</strong> 10 -10 M acts as a switch that turns on the receptor). This difference <strong>in</strong>


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 54<br />

odorant b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> EpOR1 can be seen <strong>in</strong> the Hill slopes <strong>of</strong> the dose response curves.<br />

The Hill slope for geraniol is 3.2, suggest<strong>in</strong>g a steep <strong>in</strong>crease <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g with<br />

<strong>in</strong>creas<strong>in</strong>g concentrations <strong>of</strong> the odorant, while the Hill slopes for the other<br />

compounds are all below 1, suggest<strong>in</strong>g shallower curves with a lower level <strong>of</strong><br />

cooperativity <strong>of</strong> the ligand <strong>in</strong>to the receptor b<strong>in</strong>d<strong>in</strong>g pocket.<br />

EpOR1 is broadly tuned, recognis<strong>in</strong>g a range <strong>of</strong> compounds with different chemical<br />

properties. Ligand responses at a concentration <strong>of</strong> 10 -5 M show the broad response<br />

repertoire <strong>of</strong> EpOR1 (Table 2.1), from a seven carbon ester to a 15 carbon<br />

sesquiterpene; from cyclic to acyclic compounds. This is comparable with results<br />

obta<strong>in</strong>ed by Jordan et al. (2009) for EpOR3, which b<strong>in</strong>ds fifteen different compounds,<br />

from a six carbon ester to a 15-carbon sesquiterpene. The broad tun<strong>in</strong>g <strong>of</strong> EpOR1 at<br />

10 -5 M is reduced to only five compounds when EpOR1 express<strong>in</strong>g Sf9 cells were<br />

challenged with decreas<strong>in</strong>g concentrations <strong>of</strong> the ten compounds listed <strong>in</strong> Table 2.1 <strong>in</strong><br />

a dose-dependent manner. These compounds, shown <strong>in</strong> Table 2.2 are 8-12 carbon<br />

aromatics, with the smaller compounds (methyl salicylate, has eight carbons as<br />

compared with geranyl acetate that has 12 carbons) elicit<strong>in</strong>g higher sensitivity.<br />

Perhaps the chemistry <strong>of</strong> the compound affects the b<strong>in</strong>d<strong>in</strong>g specificity <strong>of</strong> the ORs, as<br />

the results obta<strong>in</strong>ed by Hallem and Carlson, (2006) shows the preference for small<br />

aromatic r<strong>in</strong>g compounds by Drosophila OR49a. The activation threshold <strong>of</strong> EpOR3<br />

by citral from Jordan et al. (2009) is 10 -15 M. This concentration is 10 5 -fold lower than<br />

the threshold <strong>of</strong> EpOR1 for citral. The recognition <strong>of</strong> the same compound by the two<br />

E. postvittana ORs suggests that the moth employs these two receptors <strong>in</strong> a<br />

comb<strong>in</strong>atorial manner that may help to expand the dynamic range <strong>of</strong> its olfactory<br />

system. This comb<strong>in</strong>atorial approach has also been observed <strong>in</strong> Drosophila, where a<br />

range <strong>of</strong> ORs are able to detect the same odorants with different sensitivities, for<br />

example, the compounds isobutyl acetate is detected by seven different ORs, and 2-<br />

pentanone is detected by ten different ORs (Hallem and Carlson, 2006). This broad<br />

tun<strong>in</strong>g <strong>of</strong> ORs may be a mechanism for the <strong>in</strong>sect to be able to detect more<br />

compounds <strong>in</strong> various comb<strong>in</strong>ations, for example, to f<strong>in</strong>d suitable oviposition sites, so<br />

as to lay eggs <strong>in</strong> close proximity <strong>of</strong> food source for the larvae, away from <strong>in</strong>fested<br />

plants, with the limited number <strong>of</strong> ORs that they have [from the genome sequence <strong>of</strong><br />

B. mori 68 ORs have been annotated so far, (Wanner et al., 2007; Tanaka et al.,<br />

2009)]. This broad tun<strong>in</strong>g <strong>of</strong> EpOR1 to its ligands as determ<strong>in</strong>ed <strong>in</strong> the Sf9 cell assay


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 55<br />

might be very different to how the moth perceives these same ligands <strong>in</strong> the<br />

environment and does not reflect that the moth will behave <strong>in</strong> a similar manner <strong>in</strong> the<br />

environment. The comb<strong>in</strong>ation <strong>of</strong> the hundreds <strong>of</strong> odorants present <strong>in</strong> the environment<br />

together with the comb<strong>in</strong>atorial mechanism <strong>of</strong> OR function<strong>in</strong>g might give vary<strong>in</strong>g<br />

results to what is seen here.<br />

EpOR1 also recognises odorants that are relevant to the biology <strong>of</strong> E. postvittana.<br />

Geraniol for example, is a ten carbon monoterpernoid alcohol that has a rose-like<br />

odour (Gochnauer et al., 1979). Geraniol is a deterrent that affects the mean number<br />

<strong>of</strong> eggs laid by female E. postvittana and the site <strong>of</strong> oviposition (Suckl<strong>in</strong>g et al.,<br />

1996). Citral is a ten carbon monoterpene which is a mixture <strong>of</strong> the cis and trans<br />

isomers. Citral has previously been shown to act as an oviposition deterrent <strong>in</strong> E.<br />

postvittana, that reduces the number <strong>of</strong> females lay<strong>in</strong>g eggs and also the number <strong>of</strong><br />

eggs laid per female (Suckl<strong>in</strong>g et al., 1996). Methyl salicylate is an eight carbon<br />

organic ester that has a sweet woody odour. Plants produce this compound when they<br />

are under herbivore attack (James and Price, 2004). Its release results <strong>in</strong> the<br />

recruitment <strong>of</strong> beneficial <strong>in</strong>sects that will kill the herbivorous <strong>in</strong>sects. It is also used as<br />

an alarm pheromone by plants that are under pathogenic attack to warn other plants <strong>of</strong><br />

the danger. The ability <strong>of</strong> EpOR1 to recognise this compound might be an adaptation<br />

<strong>of</strong> E. postvittana towards screen<strong>in</strong>g out plants that are already <strong>in</strong>fested with other<br />

<strong>in</strong>sects, thus avoid<strong>in</strong>g these plants when look<strong>in</strong>g for suitable oviposition sites and<br />

choos<strong>in</strong>g healthy non-<strong>in</strong>fested plants where the competition for resources is less.<br />

Geranyl acetate is a 12 carbon monoterpene and has a pleasant floral or fruity aroma<br />

(Gildemeister and H<strong>of</strong>fman, 1966). This odorant has been shown to enhance the flight<br />

<strong>of</strong> grape berry moth towards its sex pheromone (Roel<strong>of</strong>s et al., 1971); hence it might<br />

act as an attractant <strong>in</strong> E. postvittana also, direct<strong>in</strong>g the moth towards either food<br />

sources or when comb<strong>in</strong>ed with the sex pheromone, enhances the attraction <strong>of</strong> male<br />

moths towards the female moths.<br />

The same odorant receptor has similar b<strong>in</strong>d<strong>in</strong>g specificities for attractants and<br />

deterrents. One <strong>of</strong> the ligands for EpOR1, citral, hav<strong>in</strong>g an EC50 <strong>of</strong> 1.3 x 10 -9 M, has<br />

previously been shown to act as an oviposition deterrent for E. postvittana <strong>in</strong><br />

behavioural experiments (Suckl<strong>in</strong>g et al., 1996). At the same time, EpOR1 can b<strong>in</strong>d<br />

geranyl acetate (with an EC50 <strong>of</strong> 2.8 x 10 -8 M), which may act as an attractant for E.


Functional characterisation <strong>of</strong> Epiphyas postvittana odorant receptor 1 56<br />

postvittana as discussed <strong>in</strong> the previous paragraph. There is only a ten-fold difference<br />

<strong>in</strong> the b<strong>in</strong>d<strong>in</strong>g aff<strong>in</strong>ity <strong>of</strong> these two compounds to the same OR that may elicit<br />

opposite behaviour <strong>in</strong> the moth. Moreover, EpOR1 has the same recognition threshold<br />

for both these compounds, recognis<strong>in</strong>g the compounds to a low concentration <strong>of</strong> 10 -10<br />

M.<br />

Functional characterisation <strong>of</strong> EpOR1 suggests that this is a general odorant receptor<br />

tuned to recognis<strong>in</strong>g important plant volatiles. The level <strong>of</strong> tun<strong>in</strong>g differs for each<br />

odorant, with the sensitivity for methyl salicylate be<strong>in</strong>g the highest. Perhaps EpOR1 is<br />

the ma<strong>in</strong> receptor that E. postvittana uses for the detection <strong>of</strong> methyl salicylate,<br />

thereby identify<strong>in</strong>g plants that are under attack. It will be <strong>in</strong>terest<strong>in</strong>g to see if any <strong>of</strong><br />

the other ORs <strong>in</strong> E. postvittana recognise methyl salicylate thus us<strong>in</strong>g a comb<strong>in</strong>atorial<br />

mechanism <strong>of</strong> b<strong>in</strong>d<strong>in</strong>g for this ligand. The high aff<strong>in</strong>ity <strong>of</strong> <strong>in</strong>sect ORs for plant<br />

volatiles makes them potential targets for <strong>in</strong>sect repellent development. F<strong>in</strong>ally, the<br />

identification and decod<strong>in</strong>g <strong>of</strong> more E. postvittana ORs will enable the decod<strong>in</strong>g <strong>of</strong> an<br />

odour–ligand map for E. postvittana and aid <strong>in</strong> deal<strong>in</strong>g with this important<br />

agricultural pest.


3.1 Introduction<br />

3<br />

The roles <strong>of</strong> Epiphyas<br />

postvittana GOBP2 <strong>in</strong><br />

odour detection by<br />

EpOR1<br />

The abundance <strong>of</strong> odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>in</strong> the sensillum lymph poses numerous<br />

questions about their function. Although OBPs were discovered <strong>in</strong> <strong>in</strong>sects nearly two<br />

decades before the odorant receptors, there cont<strong>in</strong>ues to be much debate over their<br />

roles(s). Several roles for OBPs have been postulated based on experimental data<br />

obta<strong>in</strong>ed from studies <strong>of</strong> PBPs. These prote<strong>in</strong>s form a family with 32–92% sequence<br />

identity <strong>in</strong> moths (Abraham et al., 2005). They are present at high concentrations <strong>in</strong><br />

the range <strong>of</strong> 10 µM <strong>in</strong> the sensillum lymph (Kle<strong>in</strong>, 1987) and have a role <strong>in</strong><br />

pheromone b<strong>in</strong>d<strong>in</strong>g (Vogt and Riddiford, 1981; Vogt et al., 1989; Maibeche-Coisne et<br />

al., 1997; Newcomb et al., 2002). OBPs are thought to form complexes with ligands<br />

and have been suggested to: aid the solubilisation and transport <strong>of</strong> ligands; protect<br />

ligands from enzymatic degradation <strong>in</strong> the sensillum lymph; enable <strong>in</strong>teraction <strong>of</strong><br />

ligands with dendritic bound ORs, and facilitate deactivation <strong>of</strong> the ligand (see section<br />

1.5.1 for a detailed overview on the roles <strong>of</strong> OBPs).


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 58<br />

Different methods have been used to exam<strong>in</strong>e the roles <strong>of</strong> PBPs both <strong>in</strong> vivo and <strong>in</strong><br />

vitro. Electrophysiological record<strong>in</strong>gs <strong>of</strong> s<strong>in</strong>gle antennal branches <strong>in</strong> A. polyphemus<br />

upon stimulation with pheromone have shown that ApolPBP decreases the<br />

concentration <strong>of</strong> pheromone required to elicit ORN response by 100-fold, hence<br />

<strong>in</strong>dicat<strong>in</strong>g a role <strong>of</strong> ApolPBP as a solubiliser <strong>of</strong> the pheromone (van den Berg and<br />

Ziegelberger, 1991). In <strong>in</strong> vitro characterisation assays us<strong>in</strong>g calcium imag<strong>in</strong>g <strong>of</strong><br />

HEK293 cells express<strong>in</strong>g BmOR1 and HvOR13, organic solvents were successfully<br />

replaced by the respective species‟ PBP as the solubiliser <strong>of</strong> the pheromone (Große-<br />

Wilde et al., 2006; Große-Wilde et al., 2007). A. polyphemus PBP has also been<br />

shown by <strong>in</strong> vitro b<strong>in</strong>d<strong>in</strong>g assays to act as a protector <strong>of</strong> the pheromone, as measured<br />

by reduced activity <strong>of</strong> PDE on radio-labelled pheromone <strong>in</strong> the presence <strong>of</strong> PBP (Vogt<br />

et al., 1985). Several assays have provided evidence for <strong>in</strong>teraction <strong>of</strong> the OBP/ligand<br />

complex with the dendritic bound ORs. The ab3A neuron <strong>of</strong> Drosophila has been<br />

used for express<strong>in</strong>g BmOR1 and sensitivity <strong>of</strong> the OR to its ligand bombykol was<br />

enhanced upon co-expression <strong>of</strong> BmPBP <strong>in</strong> the empty neuron (Syed et al., 2006).<br />

Similarly, when BmOR1 was expressed <strong>in</strong> HEK293 cells, replacement <strong>of</strong> the organic<br />

solubiliser DMSO with BmPBP selectively showed activity to BmPBP-bombykol<br />

complex and the DMSO solubilised activity <strong>of</strong> bombykal was lost, as measured by<br />

calcium imag<strong>in</strong>g <strong>of</strong> the BmOR1 express<strong>in</strong>g HEK293 cells (Große-Wilde et al., 2006).<br />

In vitro fluorescence measurements <strong>of</strong> endogenous tryptophan residues <strong>of</strong> BmPBP <strong>in</strong><br />

complex with bombykol at physiological pH <strong>of</strong> 7.0 and dendritic pH <strong>of</strong> 4.7 showed a<br />

decrease <strong>in</strong> fluorescence <strong>in</strong>dicat<strong>in</strong>g a dissociation <strong>of</strong> the OBP/ligand complex at low<br />

pH (Leal et al., 2005). A similar study has been done <strong>in</strong> A. polyphemus to show<br />

conformational changes <strong>of</strong> ApolPBP at different pH values (Mohanty et al., 2003;<br />

Mohanty et al., 2004).<br />

Other characterisation assays have also been used to deorphan <strong>in</strong>sect OBPs. These<br />

<strong>in</strong>clude the cold b<strong>in</strong>d<strong>in</strong>g assay, a two-phase b<strong>in</strong>d<strong>in</strong>g assay and a volatiles odorant<br />

b<strong>in</strong>d<strong>in</strong>g assay (Briand et al., 2001; Leal et al., 2005; Zhou et al., 2009). The cold<br />

b<strong>in</strong>d<strong>in</strong>g assay allows for test<strong>in</strong>g b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> ligands to OBPs at different test conditions<br />

such as pH and temperature and is able to separate free ligand from prote<strong>in</strong> bound<br />

ligand us<strong>in</strong>g a centrifugal filter device (Leal et al., 2005; Zhou et al., 2009). In this<br />

assay, the test odorants are added to the prote<strong>in</strong> solution and the mixture left to<br />

<strong>in</strong>cubate. The mixture is then centrifuged to remove unbound odorants <strong>in</strong> the filtrate


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 59<br />

while the prote<strong>in</strong> bound ligands are ma<strong>in</strong>ta<strong>in</strong>ed <strong>in</strong> the retentate. The ligand is released<br />

from the prote<strong>in</strong> by alter<strong>in</strong>g the pH <strong>of</strong> the mixture and eluted <strong>in</strong> a layer <strong>of</strong> hexane,<br />

which is then analysed for the presence <strong>of</strong> the ligands that were bound to the prote<strong>in</strong><br />

by gas chromatography (GC). Recently, Zhou et al. (2009) proposed the two-phase<br />

b<strong>in</strong>d<strong>in</strong>g assay, due to the potential release <strong>of</strong> the ligand dur<strong>in</strong>g the wash<strong>in</strong>g step <strong>in</strong><br />

cold b<strong>in</strong>d<strong>in</strong>g assay. In this assay, the ligands are dissolved <strong>in</strong> a layer <strong>of</strong> hexane that<br />

covers the prote<strong>in</strong> solution, and the depletion <strong>of</strong> the odorants <strong>in</strong> the hexane layer is<br />

measured as the uptake <strong>of</strong> odorants by the prote<strong>in</strong>s <strong>in</strong> the bottom layer.<br />

The identification <strong>of</strong> ligands for prote<strong>in</strong>s, especially if the test odorants are<br />

hydrophobic volatiles becomes difficult due to the tendency for such odorants to<br />

adhere to surfaces such as test vial walls and if mixed directly to the solution, may<br />

form micelles. The technique developed <strong>in</strong> the volatiles odorant b<strong>in</strong>d<strong>in</strong>g assay<br />

(VOBA) circumvents these problems by <strong>in</strong>troduc<strong>in</strong>g the volatile slowly <strong>in</strong>to the<br />

prote<strong>in</strong> solution via diffusion through a gas phase <strong>in</strong>termediate (Briand et al., 2001).<br />

This technique was successfully developed and used by Briand et al. (2000) to<br />

characterise an OBP isolated from rat nasal mucus. It has also been shown to facilitate<br />

the uptake <strong>of</strong> a range <strong>of</strong> compounds by the honeybee Apis mellifera OBP (Briand et<br />

al., 2001). The VOBA technique <strong>in</strong>troduces the volatile slowly, molecule by molecule<br />

so the distribution <strong>of</strong> odorants <strong>in</strong> the prote<strong>in</strong> solution is uniform. An excess <strong>of</strong> the<br />

odorant is used and the reaction is allowed to reach equilibrium so any adherence <strong>of</strong><br />

the odorant to the walls <strong>of</strong> the test vial becomes negligible.<br />

In E. postvittana, three PBPs and one GOBP were identified from native prote<strong>in</strong> gel<br />

analysis and one PBP and another GOBP each identified from EST sequences<br />

(Newcomb et al., 2002; Jordan et al., 2008). All the E. postvittana PBPs and one<br />

GOBP have been shown on western blot and by qRT-PCR analysis to be expressed <strong>in</strong><br />

both male and female antennae, albeit with vary<strong>in</strong>g levels. Two <strong>of</strong> the PBPs were<br />

shown to have higher expression <strong>in</strong> male than female antennae. Further sequence<br />

analysis revealed them to be allelic variants <strong>of</strong> the same PBP, consist<strong>in</strong>g <strong>of</strong> six am<strong>in</strong>o<br />

acid substitutions and differ<strong>in</strong>g <strong>in</strong> their electrophoretic properties, thus they were<br />

named PBP1-fast and PBP1-slow. Both the PBP1 prote<strong>in</strong>s b<strong>in</strong>d the major sex<br />

pheromone component <strong>of</strong> E. postvittana, (E)-11-tetradecenyl acetate <strong>in</strong> gel assays.<br />

The third PBP, called PBP2 was shown to have higher expression <strong>in</strong> female than male


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 60<br />

antennae hence postulated to have a female oriented role, perhaps <strong>in</strong> the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong><br />

some as yet unknown male specific pheromone. The fourth PBP, called PBP3 showed<br />

higher expression <strong>in</strong> male than female antennae, and has been postulated to be<br />

<strong>in</strong>volved <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> the pheromone component(s) <strong>of</strong> E. postvittana.<br />

The two GOBPs are members <strong>of</strong> the GOBP1 and GOBP2 family <strong>of</strong> GOBPs <strong>in</strong> moths<br />

and GOBP2 has a similar level <strong>of</strong> expression <strong>in</strong> male and female antennae (Newcomb<br />

et al., 2002). No data on sensillum localisation <strong>of</strong> EpGOBPs is available yet, however,<br />

the expression <strong>of</strong> A. polyphemus, B. mori and H. armigera GOBPs have been shown<br />

to be restricted to sensilla basiconica, hence a role <strong>of</strong> GOBPs <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g plant volatiles<br />

<strong>of</strong> importance to moths is assumed (Ste<strong>in</strong>brecht et al., 1995; Wang et al., 2003). Very<br />

limited functional data is as yet available for any moth GOBP, with the only study<br />

reveal<strong>in</strong>g that M. sexta GOBP2 b<strong>in</strong>ds the plant volatiles (Z)-3-hexen-1-ol, geraniol,<br />

geranyl acetate and limonene (Feng and Prestwich, 1997) and a recent study revealed<br />

B. mori GOBP2 b<strong>in</strong>ds bombykol and is able to discrim<strong>in</strong>ate it from the antagonist,<br />

bombykal (Zhou et al., 2009). Both M. sexta and B. mori GOBP1 have so far failed to<br />

b<strong>in</strong>d any <strong>of</strong> the ligands they have been tested with (Vogt et al., 2002; Zhou et al.,<br />

2009) lead<strong>in</strong>g to the conclusion that the prote<strong>in</strong>s tested might be <strong>in</strong>active allelic<br />

variants.<br />

3.1.1 Aim<br />

The aim <strong>of</strong> this research chapter is to test hypotheses <strong>of</strong> possible roles for GOBPs.<br />

GOBPs could be non-specific solubilis<strong>in</strong>g agents for a range <strong>of</strong> plant volatiles, or they<br />

could be ligand specific, like that observed for moth PBPs. GOBPs could form a<br />

complex with the ligand, and <strong>in</strong>teraction <strong>of</strong> this complex confers <strong>in</strong>creased sensitivity<br />

and selectivity onto the OR, as shown for H. virescens and B. mori PBPs <strong>in</strong><br />

heterologous cell assay systems. With this knowledge <strong>of</strong> roles <strong>of</strong> PBPs <strong>in</strong><br />

heterologous systems, a similar approach <strong>of</strong> us<strong>in</strong>g a reconstituted heterologous assay<br />

system can be used to test the hypothesised roles <strong>of</strong> GOBPs with E. postvittana<br />

GOBP2 (EpGOBP2) as a model: can EpGOBP2 act as a solubilis<strong>in</strong>g agent for plant<br />

volatiles? GOBP2 homologues <strong>in</strong> M. sexta and B. mori have been shown to b<strong>in</strong>d plant<br />

volatiles and sex pheromone components hence we assume EpGOBP2 to have a<br />

similar ligand b<strong>in</strong>d<strong>in</strong>g range. As part <strong>of</strong> this PhD research, E. postvittana OR1 was


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 61<br />

characterised <strong>in</strong> an Sf9 cell assay system and shown to b<strong>in</strong>d a range <strong>of</strong> plant volatiles.<br />

Some <strong>of</strong> the plant volatiles identified as ligands <strong>of</strong> EpOR1 have also been shown to be<br />

ligands for M. sexta GOBP2 (geraniol and geranyl acetate); hence lead<strong>in</strong>g us to<br />

hypothesis that EpGOBP2 might share ligands with EpOR1. The odorants used for<br />

test<strong>in</strong>g EpGOBP2 hence is limited to the b<strong>in</strong>d<strong>in</strong>g repertoire <strong>of</strong> EpOR1, as determ<strong>in</strong>ed<br />

<strong>in</strong> chapter two. The Sf9 cell assay system heterologously express<strong>in</strong>g EpOR1 forms a<br />

good start<strong>in</strong>g po<strong>in</strong>t for test<strong>in</strong>g the role(s) <strong>of</strong> EpGOBP2.<br />

3.2 Materials and methods<br />

3.2.1 Recomb<strong>in</strong>ant Expression <strong>of</strong> EpGOBP2<br />

A pET30a (Novagen) plasmid clone <strong>of</strong> EpGOBP2 (pET30a-EpGOBP2) was obta<strong>in</strong>ed<br />

from Duncan Stanley at Plant & Food Research. This clone encodes a prote<strong>in</strong><br />

conta<strong>in</strong><strong>in</strong>g an N-term<strong>in</strong>al His-tag and TEV protease recognition site upstream <strong>of</strong> the<br />

EpGOBP2 sequence (411 bp, GenBank accession no. AF411460), which is miss<strong>in</strong>g<br />

its N-term<strong>in</strong>al 20 am<strong>in</strong>o acid signal peptide (Newcomb et al., 2002).<br />

The pET30a-EpGOBP2 plasmid was transformed <strong>in</strong>to Rosetta-Gami2 cells (Novagen)<br />

and positive clones selected on kanamyc<strong>in</strong> and chloramphenicol luria broth (LB)<br />

plates. pET30a-EpGOBP2 was over-expressed <strong>in</strong> buffered terrific broth (TB) with<br />

isopropyl-beta-D-1 thiogalactopyranoside (IPTG) <strong>in</strong>duction, as stated <strong>in</strong> Hamiaux et<br />

al. (2009). Refer to appendix A for recipe <strong>of</strong> TB. Briefly, a colony <strong>of</strong> the transformed<br />

cells was used to <strong>in</strong>oculate a starter culture for prote<strong>in</strong> expression consist<strong>in</strong>g <strong>of</strong> 10 mL<br />

LB media, 15 µg/mL kanamyc<strong>in</strong>, 34 µg/mL chloramphenicol and 0.01% glucose.<br />

This was <strong>in</strong>cubated at 37°C overnight with shak<strong>in</strong>g at 220 rpm after which 5 mL <strong>of</strong><br />

the culture was used to <strong>in</strong>oculate 500 mL <strong>of</strong> buffered TB with 15 µg/mL kanamyc<strong>in</strong><br />

and 34 µg/mL chloramphenicol. The culture was left to grow at 37°C, 220 rpm until<br />

an optical density (OD600) <strong>of</strong> 0.6 to 0.8 was atta<strong>in</strong>ed. IPTG was then added to the<br />

culture to a f<strong>in</strong>al concentration <strong>of</strong> 0.5 mM, and the culture let to grow overnight at<br />

20°C shak<strong>in</strong>g at 220 rpm. The cells were harvested by centrifugation at 13,000g at<br />

4°C for 30 m<strong>in</strong>utes, with the result<strong>in</strong>g pellet stored at -20°C until purification.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 62<br />

3.2.1.1 Purification <strong>of</strong> recomb<strong>in</strong>ant EpGOBP2<br />

The cell pellet was resuspended <strong>in</strong> 20 mL HisTrap b<strong>in</strong>d<strong>in</strong>g buffer (consist<strong>in</strong>g <strong>of</strong> 20<br />

mM Tris pH 8.0, 500 mM NaCl and 20 mM imidazole) conta<strong>in</strong><strong>in</strong>g EDTA free<br />

protease <strong>in</strong>hibitors (Roche). This solution was emulsified at 15000 psi by pass<strong>in</strong>g the<br />

samples through an emulsiflex-CS high-pressure homogeniser (Avestic Inc.) three<br />

times to get maximum cell lysis. The sample was then centrifuged at 13,000g at 4°C<br />

for 30 m<strong>in</strong>utes, the result<strong>in</strong>g pellet discarded and the soluble fraction filtered through<br />

a 0.22 µm filter (Millipore) to remove any rema<strong>in</strong><strong>in</strong>g cell debris and kept on ice until<br />

purification.<br />

Purification <strong>of</strong> the His6-EpGOBP2 was carried out on a 5 mL HiTrap chelat<strong>in</strong>g HP<br />

column (GE Healthcare) by aff<strong>in</strong>ity chromatography on ÄKTAprime fast prote<strong>in</strong><br />

liquid chromatography (GE Healthcare). The prote<strong>in</strong> solution was loaded onto the<br />

column equilibrated <strong>in</strong> HisTrap b<strong>in</strong>d<strong>in</strong>g buffer conta<strong>in</strong><strong>in</strong>g 20 mM imidazole. Bound<br />

His6-EpGOBP2 was eluted from the column by a 20 to 250 mM imidazole l<strong>in</strong>ear<br />

elution gradient. The fractions conta<strong>in</strong><strong>in</strong>g His6-EpGOBP2 were identified by runn<strong>in</strong>g<br />

on 4-12% Bis-Tris SDS PAGE gel (Invitrogen) and visualised us<strong>in</strong>g colloidal<br />

commassie G-250 blue sta<strong>in</strong> (Neuh<strong>of</strong>f et al., 1988). Fractions show<strong>in</strong>g the presence <strong>of</strong><br />

His6-EpGOBP2 were pooled and dialysed overnight <strong>in</strong>to ion exchange b<strong>in</strong>d<strong>in</strong>g buffer<br />

consist<strong>in</strong>g <strong>of</strong> 20mM Tris HCl, pH 8 and 25 mM NaCl and loaded onto a Q sepharose<br />

HP column (GE Healthcare). Bound prote<strong>in</strong> was eluted from the column by a 25 to<br />

500 mM NaCl l<strong>in</strong>ear elution gradient. The fractions conta<strong>in</strong><strong>in</strong>g His6-EpGOBP2 were<br />

identified by SDS-PAGE analysis, pooled and then dialysed aga<strong>in</strong>st 50 mM Tris-HCl<br />

pH 8.0, 0.5 mM EDTA overnight. After dialysis, the His6-tag was cleaved <strong>of</strong>f with<br />

AcTEV Protease (Invitrogen) follow<strong>in</strong>g manufacturer‟s <strong>in</strong>structions. After cleavage,<br />

the prote<strong>in</strong> sample was dialysed <strong>in</strong>to HisTrap b<strong>in</strong>d<strong>in</strong>g buffer for purification <strong>of</strong> the tag<br />

free prote<strong>in</strong> on the HiTrap chelat<strong>in</strong>g HP column as stated above, this time<br />

collect<strong>in</strong>g the prote<strong>in</strong> <strong>in</strong> the flow through.<br />

3.2.1.2 Delipidation <strong>of</strong> recomb<strong>in</strong>ant EpGOBP2<br />

Recomb<strong>in</strong>ant OBPs have a tendency to b<strong>in</strong>d exogenous ligands from the expression<br />

cells hence these artefact ligands have to be removed from the prote<strong>in</strong>s before any


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 63<br />

functional analysis is carried out (Oldham et al., 2000; Hamiaux et al., 2009). The<br />

tag-cleaved prote<strong>in</strong> was delipidated accord<strong>in</strong>g to Oldham et al. (2001). Briefly, the<br />

prote<strong>in</strong> was concentrated to 50 µL us<strong>in</strong>g a Vivasp<strong>in</strong> ultrafiltration concentrator<br />

(capacity 500 µL, MWCO 5 kDa) at 10,000 rpm, 4°C (Vivaproducts, MA). 500 µL <strong>of</strong><br />

25 mM ammonium acetate pH 4.5 was then added to the retentate and centrifuged<br />

aga<strong>in</strong> to 50 µL. This was repeated a second time and then 350 µL <strong>of</strong> 25 mM<br />

ammonium acetate pH 4.5 was added aga<strong>in</strong> to the retentate and placed on ice. 500 µL<br />

<strong>of</strong> the Lipidex 1000 methanol suspension (Perk<strong>in</strong> Elmer) was centrifuged <strong>in</strong> a MC-<br />

Ultrafree centrifuge filter cup (capacity 500 µL, pore size 0.22 µm, Millipore,<br />

Bedford, MA) at 10,000 rpm for one m<strong>in</strong>ute to remove the liquid phase. The flow<br />

through was discarded and another 500 µL <strong>of</strong> Lipidex was added to the filter cup,<br />

centrifuged and the flow through discarded aga<strong>in</strong>. The Lipidex <strong>in</strong> the filter cup was<br />

mixed with 500 µL <strong>of</strong> 25 mM ammonium acetate pH 4.5, centrifuged at 10,000 rpm,<br />

for one m<strong>in</strong>ute and the flow through discarded. This was repeated aga<strong>in</strong> and then the<br />

EpGOBP2 sample was mixed thoroughly with the Lipidex, and the result<strong>in</strong>g<br />

suspension <strong>in</strong>cubated <strong>in</strong> the filter cup at 37°C for one hour with gentle shak<strong>in</strong>g. The<br />

suspension was then centrifuged at 10,000 rpm, 4°C; the flow through collected and<br />

kept on ice. The Lipidex was washed with 500 µL <strong>of</strong> 25 mM ammonium acetate pH<br />

4.5 (to remove any rema<strong>in</strong><strong>in</strong>g traces <strong>of</strong> EpGOBP2), centrifuged and flow through<br />

collected aga<strong>in</strong>. Both the flow through samples were comb<strong>in</strong>ed and concentrated<br />

us<strong>in</strong>g the Vivasp<strong>in</strong> ultrafiltration concentrator to 50 µL. The concentrated, delipidated<br />

EpGOBP2 was diluted <strong>in</strong> 2.5 mM ammonium acetate at neutral pH to a concentration<br />

<strong>of</strong> 100 µM and dialysed aga<strong>in</strong>st 50 mM Tris-HCl buffer pH 7.5 for conduct<strong>in</strong>g<br />

b<strong>in</strong>d<strong>in</strong>g assays.<br />

3.2.2 Volatile odorant b<strong>in</strong>d<strong>in</strong>g assay<br />

The volatile odorant b<strong>in</strong>d<strong>in</strong>g assay (VOBA) (Briand et al., 2000) was used for test<strong>in</strong>g<br />

the ligand preferences <strong>of</strong> EpGOBP2. Fifty microlitres <strong>of</strong> 5 µM prote<strong>in</strong> <strong>in</strong> 50 mM Tris-<br />

HCl, pH 7.5 or 50 µL <strong>of</strong> 50 mM Tris-HCl buffer pH 7.5 (control) were setup <strong>in</strong><br />

triplicates <strong>in</strong> 200 µL PCR tubes. The tubes were placed with lids open <strong>in</strong> Conway<br />

microdiffusion dish (Gallenkamp). A drop <strong>of</strong> the absolute compound to be tested was<br />

added to the dish and the system was sealed with a glass lid. The system was left to<br />

equilibrate for approximately 12 hours at room temperature, as depicted <strong>in</strong> Figure 3.1.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 64<br />

Figure 3.1: Schematic representation <strong>of</strong> the VOBA sett<strong>in</strong>g to show the distribution <strong>of</strong><br />

the odorant at equilibrium <strong>in</strong> a sealed chamber.<br />

The tubes were removed after the 12 hour <strong>in</strong>cubation. To release the bound<br />

compounds, 5 µL <strong>of</strong> 1 mg/mL prote<strong>in</strong>ase K (Roche) was added to the tube content,<br />

which was then covered with a layer <strong>of</strong> hexane (to capture odorants released from<br />

digested prote<strong>in</strong>). This was done for all the tubes, <strong>in</strong>clud<strong>in</strong>g the buffer only control. 5<br />

µM <strong>of</strong> an <strong>in</strong>ternal standard, methyl tetradecanoate (14OMe) was added to all the<br />

tubes. The closed tubes were then <strong>in</strong>cubated at room temperature for two hours. The<br />

hexane layer was removed from tubes <strong>in</strong>to clean 2 mL glass vials and analyzed by GC<br />

us<strong>in</strong>g a HP 6890 Series GC System (Hewlett Packard) equipped with an on-column<br />

<strong>in</strong>jector and FID detector (300°C). The analytical column used was a HP-INNOWax<br />

Polyethylene Glycol (nom<strong>in</strong>al length = 30.0 m, nom<strong>in</strong>al diameter = 250.00 µm and<br />

nom<strong>in</strong>al film thickness = 0.25 µm; Agilent Technologies). The oven temperature<br />

gradient was applied from 120°C to 260°C at 5°C.m<strong>in</strong> -1 . The carrier gas was helium at<br />

0.7mL.m<strong>in</strong> -1 . Calibration was done by odorants diluted <strong>in</strong> chlor<strong>of</strong>orm.<br />

The concentration <strong>of</strong> the compounds was determ<strong>in</strong>ed by measur<strong>in</strong>g surface <strong>of</strong> peaks<br />

with respect to the calibration curve. The concentration <strong>of</strong> the bound ligand and the<br />

b<strong>in</strong>d<strong>in</strong>g aff<strong>in</strong>ity (dissociation constant, ) <strong>of</strong> the ligands with the prote<strong>in</strong> were<br />

calculated us<strong>in</strong>g the method <strong>of</strong> Lazar (2001) as follows:<br />

Prote<strong>in</strong><br />

Test ligand<br />

Buffer


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 65<br />

From the results obta<strong>in</strong>ed <strong>in</strong> the GC,<br />

[L] = free ligand (total ligand present <strong>in</strong> buffer only)<br />

[P] = prote<strong>in</strong> concentration used for VOBA (5 µM)<br />

[PL] = concentration <strong>of</strong> bound ligand (Total ligand present <strong>in</strong> prote<strong>in</strong> sample – [L])<br />

= dissociation constant (the smaller the , the tighter the b<strong>in</strong>d<strong>in</strong>g)<br />

3.2.3 Reconstituted cell assay<br />

The reconstituted cell assay was carried out as described <strong>in</strong> Groβe–Wilde et al. (2006)<br />

with a few modifications. Briefly, odorants were made up <strong>in</strong> the concentration range<br />

<strong>of</strong> 10 -5 M to 10 -14 M as described <strong>in</strong> section 2.2.6 <strong>in</strong> either assay sal<strong>in</strong>e alone or assay<br />

sal<strong>in</strong>e conta<strong>in</strong><strong>in</strong>g 10 -6 M EpGOBP2 solution. The Sf9 cell transfection, assay and data<br />

analysis were performed as stated <strong>in</strong> sections 2.2.6 and 2.2.7.<br />

3.3 Results<br />

3.3.1 Purification <strong>of</strong> recomb<strong>in</strong>ant GOBP2<br />

Recomb<strong>in</strong>ant EpGOBP2 was purified <strong>in</strong> order to carry out functional ligand b<strong>in</strong>d<strong>in</strong>g<br />

assays to deduce its ligands and also to test its role(s) <strong>in</strong> odorant b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> EpOR1 <strong>in</strong><br />

the Sf9 cell assay system. His6-EpGOBP2 eluted as a s<strong>in</strong>gle peak from 60 to 75%<br />

imidazole. The six fractions collected dur<strong>in</strong>g this elution peak were run on 4-12%<br />

SDS-PAGE gel (Figure 3.2). No EpGOBP2 was present <strong>in</strong> the flow through<br />

<strong>in</strong>dicat<strong>in</strong>g all the EpGOBP2 loaded onto the column was bound to it. However, not all<br />

<strong>of</strong> the EpGOBP2 separated <strong>in</strong>to the soluble phase as there is still a thick band <strong>in</strong> the<br />

<strong>in</strong>soluble fraction at 22.6 kDa.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 66<br />

Figure 3.2: Purification <strong>of</strong> His6-EpGOBP2 by HiTrap chelat<strong>in</strong>g HP column. A<br />

dom<strong>in</strong>ant EpGOBP2 band is observed at 22.6 kDa <strong>in</strong> lanes 4–9. Lane 1 is the<br />

<strong>in</strong>soluble fraction conta<strong>in</strong><strong>in</strong>g the cell debris, lane 2 is the crude prote<strong>in</strong> sample that<br />

was loaded onto the column and lane 3 is the flow through.<br />

His6-tagged EpGOBP2 was present <strong>in</strong> all six fractions over the elution peak, shown<br />

by a dom<strong>in</strong>ant band at 22.6 kDa. However, other bands were present <strong>in</strong> the fractions<br />

also, possibly E.coli prote<strong>in</strong>s that <strong>in</strong>teracted with the nickel column. To further purify<br />

EpGOBP2, the six fractions from the HiTrap chelat<strong>in</strong>g column were comb<strong>in</strong>ed and<br />

dialysed <strong>in</strong>to ion exchange b<strong>in</strong>d<strong>in</strong>g buffer for purification on Q-sepharose HP column<br />

(Amersham Biosciences). The prote<strong>in</strong> was eluted us<strong>in</strong>g a gradient <strong>of</strong> 500 mM NaCl.<br />

The seven elution peak fractions obta<strong>in</strong>ed were run on 4–12% a SDS-PAGE gel as<br />

shown <strong>in</strong> Figure 3.3.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 67<br />

Figure 3.3: Purification <strong>of</strong> His6-EpGOBP2 from the pooled HiTrap fractions by Qsepharose<br />

anion exchange chromatography. Lane 1 is the EpGOBP2 dialysed <strong>in</strong>to ion<br />

exchange b<strong>in</strong>d<strong>in</strong>g buffer and lanes 2–8 are the elution peak fractions from the Qsepharose<br />

HP column conta<strong>in</strong><strong>in</strong>g the His6-EpGOBP2.<br />

The fractions obta<strong>in</strong>ed after the ion exchange were pooled and the His6-tag was<br />

cleaved <strong>of</strong>f by AcTEV digestion. After tag removal, the prote<strong>in</strong> was recovered by<br />

runn<strong>in</strong>g the sample on HiTrap chelat<strong>in</strong>g HP column, this time collect<strong>in</strong>g the prote<strong>in</strong><br />

<strong>in</strong> the flow through (Figure 3.4I). The prote<strong>in</strong> sample was f<strong>in</strong>ally put through a<br />

delipidation procedure to remove any exogenous ligands that bound to the prote<strong>in</strong><br />

from the E. coli cells (Figure 3.4II).


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 68<br />

Figure 3.4: Purification <strong>of</strong> AcTEV cleaved EpGOBP2 by HiTrap chelat<strong>in</strong>g column<br />

chromatography. (I) is the EpGOBP2 after treatment with AcTEV protease. Lane 1 is<br />

the prote<strong>in</strong> sample that was loaded onto the column, lane 2 is the prote<strong>in</strong> bound to the<br />

nickel column and lane 3 is the flow through from the Nickel column. Band A is the<br />

His6-EpGOB2, band B is EpGOBP without the His6 tag and band C is the His6-tag.<br />

(II) is the prote<strong>in</strong> without the tag after go<strong>in</strong>g through the delipidation process.<br />

3.3.2 Volatile Odorant B<strong>in</strong>d<strong>in</strong>g Assay<br />

Ligands that bound to EpGOBP2 were identified us<strong>in</strong>g the volatile odorant b<strong>in</strong>d<strong>in</strong>g<br />

assay (Briand et al., 2000). The b<strong>in</strong>d<strong>in</strong>g aff<strong>in</strong>ity <strong>of</strong> EpGOBP2 to a panel <strong>of</strong> ten<br />

odorants that EpOR1 recognises was determ<strong>in</strong>ed. EpGOBP2 was found to b<strong>in</strong>d seven<br />

out <strong>of</strong> the ten compounds that were tested <strong>in</strong> the micromolar range (Figure 3.5).


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 69<br />

Figure 3.5: Overnight VOBA <strong>of</strong> EpGOBP2 with the ligands <strong>of</strong> EpOR1 <strong>in</strong>dicat<strong>in</strong>g<br />

that EpGOBP2 b<strong>in</strong>ds seven <strong>of</strong> these compounds <strong>in</strong> the micromolar range. The error<br />

bars represent standard errors calculated from triplicate repeats. * = no detectable<br />

b<strong>in</strong>d<strong>in</strong>g.<br />

Figure 3.6: Dissociation constants ( ) for ten odorants aga<strong>in</strong>st EpGOBP2. are<br />

the average from 3 replicates with error bars represent<strong>in</strong>g standard errors. * = no<br />

detectable b<strong>in</strong>d<strong>in</strong>g.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 70<br />

For the seven ligands identified <strong>in</strong> the scope <strong>of</strong> this study, octanol had the highest<br />

b<strong>in</strong>d<strong>in</strong>g capacity for EpGOBP2, with 99 µM <strong>of</strong> octanol b<strong>in</strong>d<strong>in</strong>g to 5 µM <strong>of</strong> EpGOBP2<br />

while only 9 µM <strong>of</strong> pentyl acetate bound to the same concentration <strong>of</strong> EpGOBP2<br />

(Figure 3.5).<br />

Ten compounds were identified as ligands for EpOR1 <strong>in</strong> Chapter 2, however, at lower<br />

concentrations, EpOR1 was able to elicit a response to only five <strong>of</strong> these (geranyl<br />

acetate, citral, methyl salicylate, geraniol and geranial) hence dose response curves<br />

were able to be obta<strong>in</strong>ed for only these five compounds. From these five ligands <strong>of</strong><br />

EpOR1, geranyl acetate and methyl salicylate are able to b<strong>in</strong>d to EpGOBP2, hence<br />

both these compounds were used for test<strong>in</strong>g the response <strong>of</strong> EpOR1 express<strong>in</strong>g Sf9<br />

cells by reconstitut<strong>in</strong>g the Sf9 cell assay system with EpGOBP2.<br />

3.3.3 Reconstituted EpOR1 receptor activation assays<br />

In order to carry out reconstituted EpOR1 receptor activation assays <strong>in</strong> the Sf9 assay<br />

system with EpGOBP2, the common ligands <strong>of</strong> both EpGOBP2 and EpOR1 for<br />

which dose response curves have been determ<strong>in</strong>ed <strong>in</strong> Chapter 2 are used. These<br />

ligands are methyl salicylate and geranyl acetate. The dose response <strong>of</strong> EpOR1<br />

expressed <strong>in</strong> the Sf9 cell assay to methyl salicylate and geranyl acetate was measured<br />

under the follow<strong>in</strong>g conditions: <strong>in</strong> the absence <strong>of</strong> both EpGOBP2 and DMSO (Figures<br />

3.8A and 3.9A, no solubilis<strong>in</strong>g agent present hence the odorant does not elicit a<br />

response <strong>in</strong> the EpOR1 express<strong>in</strong>g cells); <strong>in</strong> the presence <strong>of</strong> EpGOBP2 only (Figures<br />

3.8C and 3.9C, a dose response <strong>of</strong> the EpOR1 express<strong>in</strong>g cells is observed to both<br />

geranyl acetate and methyl salicylate <strong>in</strong>dicat<strong>in</strong>g that the GOBP2 is able to solubilise<br />

these two odorants); or DMSO only (Figures 3.8B and 3.9B, DMSO acts as a<br />

solubilis<strong>in</strong>g agent for methyl salicylate and geranyl acetate as determ<strong>in</strong>ed <strong>in</strong> Chapter<br />

2); and <strong>in</strong> the presence <strong>of</strong> both DMSO and EpGOBP2 (Figures 3.8D and 3.9D, to<br />

determ<strong>in</strong>e whether the changes <strong>in</strong> the EC50 values <strong>of</strong> the various dose response curves<br />

are due to the presence <strong>of</strong> EpGOBP2 or DMSO). These sets <strong>of</strong> experiements showed<br />

that EpGOBP2 can act as a solubilis<strong>in</strong>g agent for methyl salicylate and geranyl<br />

acetate. To show that the dose response observed for both methyl salicylate and<br />

geranyl acetate <strong>in</strong> the presence <strong>of</strong> EpGOBP2 is not due to a concentration effect <strong>of</strong> the


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 71<br />

EpGOBP2, a no-odorant dose response control assay <strong>of</strong> the EpOR1 express<strong>in</strong>g Sf9<br />

cells was carried out. This showed that <strong>in</strong> the absence <strong>of</strong> a ligand, EpGOBP2 does not<br />

elicit a response <strong>in</strong> EpOR1 express<strong>in</strong>g Sf9 cells, as depicted <strong>in</strong> Figure 3.7. The<br />

concentrations at which half the maximal responses (EC50) were obta<strong>in</strong>ed for the<br />

different test concentrations is given <strong>in</strong> Table 3.1.<br />

Figure 3.7: Dose response <strong>of</strong> EpOR1 express<strong>in</strong>g Sf9 cells to EpGOBP2 <strong>in</strong> the<br />

concentration range <strong>of</strong> 10 -5 M to 10 -13 M. Error bars represent the standard error <strong>of</strong> six<br />

to eight cells.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 72<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

A<br />

-12 -10 -8 -6 -4<br />

[Geranyl Acetate] (Log M)<br />

-12 -10 -8 -6 -4<br />

[Geranyl Acetate] (Log M)<br />

C<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

B<br />

-12 -10 -8 -6 -4<br />

[Geranyl Acetate] (Log M)<br />

D<br />

-12 -10 -8 -6 -4<br />

[Geranyl Acetate] (Log M)<br />

Figure 3.8: Dose response curves <strong>of</strong> EpOR1 to geranyl acetate <strong>in</strong> the concentration<br />

range <strong>of</strong> 10 -5 M to 10 -12 M <strong>in</strong> the absence <strong>of</strong> both DMSO and EpGOBP2 (A), with<br />

DMSO (B), or with EpGOBP2 (C), or with geranyl acetate solubilised <strong>in</strong> DMSO and<br />

the cells <strong>in</strong>cubated with EpGOBP2 (D). Error bars represent the standard error <strong>of</strong> six<br />

to eight respond<strong>in</strong>g cells.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 73<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

A<br />

-16 -14 -12 -10 -8 -6 -4<br />

[Methyl Salicylate] (Log M)<br />

C<br />

-16 -14 -12 -10 -8 -6 -4<br />

[Methyl Salicylate] (Log M)<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

∆F<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

B<br />

-16 -14 -12 -10 -8 -6 -4<br />

[Methyl Salicylate] (Log M)<br />

D<br />

-16 -14 -12 -10 -8 -6 -4<br />

[Methyl Salicylate] (Log M)<br />

Figure 3.9: Dose response curves <strong>of</strong> EpOR1 to methyl salicylate <strong>in</strong> the concentration<br />

range <strong>of</strong> 10 -5 M to 10 -16 M <strong>in</strong> the absence <strong>of</strong> both DMSO and EpGOBP2 (A), with<br />

DMSO (B), or with EpGOBP2 (C) or with methyl salicylate solubilised <strong>in</strong> DMSO and<br />

the cells <strong>in</strong>cubated with EpGOBP2 (D). Error bars represent the standard error <strong>of</strong> six<br />

to eight respond<strong>in</strong>g cells.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 74<br />

Table 3.1: A comparison <strong>of</strong> the EC50 values from dose response curves <strong>of</strong> EpOR1<br />

obta<strong>in</strong>ed <strong>in</strong> Sf9 cell assays with DMSO and/or EpGOBP2 used as a solubilisation<br />

agent for the ligands methyl salicylate and geranyl acetate.<br />

Ligand components EC50[Methyl Salicylate] (M) EC50[Geranyl Acetate] (M)<br />

Sal<strong>in</strong>e only NR a NR<br />

DMSO only (1.8 ± 0.9) -12 (2.7 ± 0.4) -8<br />

EpGOBP2 only (1.19 ± 0.82) -13 (1.9 ± 0.66) -10<br />

DMSO + EpGOBP2 (1.33 ± 0.96) -13 (1.97 ± 2.5) -10<br />

a. NR = no response<br />

3.4 Discussion<br />

<strong>Insects</strong> can recognise a range <strong>of</strong> chemicals, most <strong>of</strong> which are hydrophobic <strong>in</strong> nature.<br />

The actual mechanism <strong>of</strong> odour recognition by <strong>in</strong>sects is yet to be deduced; however<br />

theoretical models suggest solubilisation and target<strong>in</strong>g <strong>of</strong> the hydrophobic odours to<br />

the receptor neurons via soluble prote<strong>in</strong> <strong>in</strong>termediates present <strong>in</strong> the sensillum lymph<br />

(Kaissl<strong>in</strong>g, 2009). We can test the roles <strong>of</strong> these soluble prote<strong>in</strong>s <strong>in</strong> an <strong>in</strong> vitro sett<strong>in</strong>g<br />

by express<strong>in</strong>g the receptor <strong>in</strong> cell l<strong>in</strong>es such as the Sf9 <strong>in</strong>sect cell l<strong>in</strong>es used <strong>in</strong> this<br />

study, and look<strong>in</strong>g at the effect on receptor activation upon addition <strong>of</strong> these prote<strong>in</strong>s.<br />

The ability <strong>of</strong> EpGOBP2 to solubilise plant volatiles <strong>of</strong> importance to E. postvittana<br />

was tested <strong>in</strong> a VOBA sett<strong>in</strong>g. The ten compounds tested were chosen based on their<br />

ability to activate EpOR1 expressed <strong>in</strong> the Sf9 cell assay system. EpGOBP2 was able<br />

to solubilise seven <strong>of</strong> these compounds; however the b<strong>in</strong>d<strong>in</strong>g repertoire <strong>of</strong> EpGOBP2<br />

may be more extensive if the range <strong>of</strong> tested compounds was to be expanded. For the<br />

scope <strong>of</strong> this study, only those compounds that elicited response to EpOR1 were<br />

chosen for test<strong>in</strong>g their b<strong>in</strong>d<strong>in</strong>g to EpGOBP2.<br />

The chemistries <strong>of</strong> these seven compounds differ considerably. Pentyl acetate is an<br />

ester that has a straight cha<strong>in</strong> structure, while geranyl acetate and nerol are l<strong>in</strong>ear<br />

monoterpenes, eucalyptol is a cyclic monoterpene and α-farnesene is a sesquiterpene.<br />

Methyl salicylate is a benzoic r<strong>in</strong>g phenol and octanol is a l<strong>in</strong>ear alcohol (O'Neil,<br />

2006). The amount <strong>of</strong> each <strong>of</strong> these ligands bound by EpGOBP2 differs accord<strong>in</strong>g to<br />

their chemistries, with EpGOBP2 hav<strong>in</strong>g highest aff<strong>in</strong>ity for alcohol followed by the


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 75<br />

phenol, then the terpenes and lastly the ester (Figure 3.5). However, the difference <strong>in</strong><br />

the degree <strong>of</strong> b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> these ligands is small, with only a ten-fold difference<br />

between the highest octanol, and the lowest, pentyl acetate. Nevertheless, all these<br />

compounds are plant volatiles and may play a role <strong>in</strong> E. postvittana olfaction (as<br />

discussed <strong>in</strong> chapter 2).<br />

The moderate b<strong>in</strong>d<strong>in</strong>g constants <strong>of</strong> EpGOBP2 for the range <strong>of</strong> plant volatiles tested<br />

reflect low aff<strong>in</strong>ity which supports the conclusion that GOBPs function to b<strong>in</strong>d a<br />

range <strong>of</strong> odorants. These moderate b<strong>in</strong>d<strong>in</strong>g constants <strong>of</strong> EpGOBP2 for a range <strong>of</strong><br />

volatiles are <strong>in</strong>dicative <strong>of</strong> a role <strong>in</strong> b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> a range <strong>of</strong> odorants by OBPs. Previous<br />

studies <strong>of</strong> GOBPs <strong>in</strong> other lepidopterans have shown their ability to recognise plant<br />

volatiles. M. sexta GOBP2 b<strong>in</strong>ds plant odorants such as (Z)-3-hexen-1-ol, geraniol,<br />

geranyl acetate and limonene (Feng and Prestwich, 1997) with moderate aff<strong>in</strong>ities (10<br />

to 40 µM <strong>of</strong> these odorants were required <strong>in</strong> displacement assays to displace 50 % <strong>of</strong><br />

the bound pheromone analog), comparable to those obta<strong>in</strong>ed for EpGOBP2 <strong>in</strong> this<br />

study. In the honeybee Apis mellifera, the OBP called ASP2 b<strong>in</strong>ds its ligands, mostly<br />

plant scents, with high b<strong>in</strong>d<strong>in</strong>g aff<strong>in</strong>ities <strong>in</strong> the micromolar range (Briand et al., 2001).<br />

This shows that perhaps the Hymenoptera and Lepidoptera OBPs differ <strong>in</strong> their<br />

mechanisms <strong>of</strong> b<strong>in</strong>d<strong>in</strong>g their ligands. S<strong>in</strong>ce EpGOBP2 b<strong>in</strong>ds a range <strong>of</strong> compounds<br />

with different shapes and chemical structures, it can be deduced that the b<strong>in</strong>d<strong>in</strong>g<br />

pocket <strong>of</strong> EpGOBP2 must be spacious and flexible enough to harbor this varied range<br />

<strong>of</strong> compounds. The range <strong>of</strong> compounds tested <strong>in</strong> this study is limited; hence there<br />

might be other stronger b<strong>in</strong>d<strong>in</strong>g ligand(s) that have not been tested yet. Increas<strong>in</strong>g the<br />

test compound range will perhaps give <strong>in</strong>sights <strong>in</strong>to an even wider range <strong>of</strong> ligands for<br />

EpGOBP2, add<strong>in</strong>g to the observation that while the PBP recognise species specific<br />

pheromones, the GOBPs are broadly tuned to recognise a range <strong>of</strong> compounds that<br />

<strong>in</strong>cludes (and might not be limited to) plant volatiles.<br />

When EpGOBP2 was pre-<strong>in</strong>cubated with geranyl acetate or methyl salicylate, and<br />

added to Sf9 cells express<strong>in</strong>g EpOR1, a receptor activation response was observed<br />

(Figures 3.8 and 3.9), <strong>in</strong>dicat<strong>in</strong>g that EpGOBP2 was able to solubilise the odorants<br />

enabl<strong>in</strong>g it to <strong>in</strong>teract with the membrane bound EpOR1. This is the first documented<br />

role <strong>of</strong> a lepidopteran GOBP <strong>in</strong> odorant b<strong>in</strong>d<strong>in</strong>g <strong>in</strong> an <strong>in</strong> vitro OR characterisation<br />

assay. Similar reconstitution assay studies have been done with lepidopteran PBPs,


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 76<br />

namely BmPBP1 and HvPBP2 (Große-Wilde et al., 2006; Große-Wilde et al., 2007).<br />

In these studies, the PBPs were shown to be able to replace the DMSO as a<br />

solubilis<strong>in</strong>g agent for the sex pheromone components tested, and obta<strong>in</strong> functional<br />

data for pheromone receptors when tested <strong>in</strong> a heterologous expression system.<br />

Despite the different types <strong>of</strong> odorants GOBP2 b<strong>in</strong>ds when compared to these<br />

narrowly tuned PBPs, the EpGOBP2 was able to replace DMSO <strong>in</strong> EpOR1<br />

characterisation assays, act<strong>in</strong>g as a solubilis<strong>in</strong>g agent for methyl salicylate and geranyl<br />

acetate.<br />

When DMSO was replaced by EpGOBP2 <strong>in</strong> the Sf9 cell assay, the EC50 value<br />

decreased (Table 3.1), <strong>in</strong>dicat<strong>in</strong>g that <strong>in</strong> the presence <strong>of</strong> EpGOBP2, EpOR1<br />

express<strong>in</strong>g cells were able to b<strong>in</strong>d more odorants at lower concentrations compared to<br />

when DMSO was used. This was true for both methyl salicylate and geranyl acetate<br />

(Table 3.1). The response was not a spontaneous response to EpGOBP2 as assays<br />

carried out with EpGOBP2 alone did not show any response (Figure 3.7), <strong>in</strong>dicat<strong>in</strong>g<br />

that EpGOBP2 without be<strong>in</strong>g exposed to its b<strong>in</strong>d<strong>in</strong>g components is unresponsive to<br />

EpOR1. A possible mechanism <strong>of</strong> EpGOBP2 function <strong>in</strong> the EpOR1 characterisation<br />

assay could be that a complex is formed between the EpGOBP2 and the ligand, and<br />

this complex <strong>in</strong>teract with EpOR1. This is supported by the decrease <strong>in</strong> EC50 values<br />

for both geranyl acetate and methyl salicylate b<strong>in</strong>d<strong>in</strong>g to EpOR1. One hypothesis is<br />

that EpGOBP2 might b<strong>in</strong>d the ligands and release them at the membrane surface to<br />

<strong>in</strong>crease the local concentration <strong>of</strong> ligand <strong>in</strong> close vic<strong>in</strong>ity <strong>of</strong> the OR. It could also be<br />

that the observed decrease <strong>in</strong> EC50 is a result <strong>of</strong> EpGOBP2 act<strong>in</strong>g as an activated<br />

ligand, as observed for the Drosophila OBP LUSH, which upon b<strong>in</strong>d<strong>in</strong>g its ligand,<br />

11-cis vaccenyl acetate undergoes a conformational change (Laughl<strong>in</strong> et al., 2008).<br />

LUSH alone <strong>in</strong> this altered state is able to b<strong>in</strong>d to and activate pheromone sensitive<br />

neurons, <strong>in</strong>dicat<strong>in</strong>g the role <strong>of</strong> a PBP as an activated ligand. EpGOBP2 upon b<strong>in</strong>d<strong>in</strong>g<br />

to geranyl acetate and methyl salicylate might become activated and act as a ligand<br />

itself and <strong>in</strong>teract with the OR, hence lead<strong>in</strong>g to lower EC50 values.<br />

The comb<strong>in</strong>ed effect <strong>of</strong> either ligand b<strong>in</strong>d<strong>in</strong>g to EpOR1 <strong>in</strong> the presence <strong>of</strong> both<br />

DMSO and EpGOBP2 is similar to each b<strong>in</strong>d<strong>in</strong>g EpOR1 <strong>in</strong> the presence <strong>of</strong> EpGOBP2<br />

alone. These data <strong>in</strong>dicate that the change <strong>in</strong> the EC50 values <strong>of</strong> EpOR1 b<strong>in</strong>d<strong>in</strong>g to<br />

methyl salicylate and geranyl acetate could be attributed to the presence <strong>of</strong> EpGOBP2


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 77<br />

<strong>in</strong> the assay. Thus raises the possibility that it could be act<strong>in</strong>g as a solubiliser as well<br />

as an activated ligand to confer <strong>in</strong>creased sensitivity on EpOR1.<br />

EpGOBP2 seem to have <strong>in</strong>creased selectivity <strong>of</strong> EpOR1 to geranyl acetate, as a 100-<br />

fold difference <strong>in</strong> EC50 value was observed upon replacement <strong>of</strong> DMSO with geranyl<br />

acetate. In comparison, the EC50 <strong>of</strong> EpOR1 for methyl salicylate decreased by only<br />

ten-fold. This suggests that EpGOBP2 <strong>in</strong>creases the selectivity <strong>of</strong> EpOR1 to geranyl<br />

acetate over methyl salicylate. However, this change is not sufficient to make the<br />

assay system more sensitive for geranyl acetate over methyl salicylate, as the EC50 <strong>of</strong><br />

EpOR1 for methyl salicylate is still a 1000-fold more than for geranyl acetate. Unlike<br />

PBPs, which can have a significant impact on ligand specificity <strong>of</strong> PRs when organic<br />

solubilisers are replaced by the PBPs, for example, the response <strong>of</strong> BmOR3 to<br />

bombykal is obliterated when DMSO is replaced with BmPBP1; EpGOBP2 has a<br />

broader range <strong>of</strong> ligands and does not have a significant impact on ligand specificity<br />

(Große-Wilde et al., 2006). This may be attributed to moths hav<strong>in</strong>g a similar number<br />

<strong>of</strong> PBPs as the number <strong>of</strong> components that make up the sex pheromones, hence the<br />

ratio <strong>of</strong> PBP:pheromone component is anywhere between 1:1 or 2:1, thereby allow<strong>in</strong>g<br />

the PBP to be selective for its ligands. On the other hand, only two families <strong>of</strong> GOBPs<br />

have been identified <strong>in</strong> moths so far (GOBP1 and GOBP2), while the receptive range<br />

<strong>of</strong> moths for plant volatiles can be <strong>in</strong> the hundreds. Hence the need for GOBPs to b<strong>in</strong>d<br />

a broad range <strong>of</strong> odorants is <strong>in</strong>evitable if the <strong>in</strong> vitro suggested roles <strong>in</strong> this discussion<br />

are applied <strong>in</strong> vivo by the moth. GOBPs thus become biologically relevant for the<br />

moth‟s success <strong>in</strong> detect<strong>in</strong>g host plants. Specificity for pheromone detection has been<br />

shown to be coded by PBP/PR/pheromone comb<strong>in</strong>ation. No such evidence for plant<br />

volatile specificity can be deduced with<strong>in</strong> the scope <strong>of</strong> this study. Thus EpGOBP2<br />

either acts as an activated ligand itself when bound by odorants, or acts as a<br />

solubilis<strong>in</strong>g agent that forms a complex with the ligands and aids its transport and<br />

release near to the membrane surface to <strong>in</strong>crease the local concentration <strong>of</strong> the ligands<br />

near the receptor. The sensitivity <strong>of</strong> methyl salicylate and geranyl acetate b<strong>in</strong>d<strong>in</strong>g to<br />

EpOR1 was <strong>in</strong>creased <strong>in</strong> the presence <strong>of</strong> EpGOBP2, <strong>in</strong>dicat<strong>in</strong>g the <strong>in</strong>volvement <strong>of</strong><br />

EpGOBP2 <strong>in</strong> odorant b<strong>in</strong>d<strong>in</strong>g by EpOR1. A similar approach can perhaps be used for<br />

characteris<strong>in</strong>g ORs <strong>in</strong> heterologous expression systems which are sensitive to organic<br />

solvents where the organic solvent used for solubilis<strong>in</strong>g odorants can be replaced by<br />

b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s.


The roles <strong>of</strong> Epiphyas postvittana GOBP2 <strong>in</strong> odour detection 78<br />

This research has demonstrated the role <strong>of</strong> EpGOBP2 as a solubilis<strong>in</strong>g agent for plant<br />

volatiles and shown that it can be used to reconstitute a Sf9 cell assay <strong>of</strong> EpOR1. This<br />

paves the way for further studies where the sensilla lymph can be replicated <strong>in</strong> tissue<br />

culture. The OR has been expressed <strong>in</strong> vitro, a role for the GOBP has been<br />

demonstrated, and an ODE can be added to this exist<strong>in</strong>g system to complete the<br />

sensilla lymph replica <strong>in</strong> tissue culture. Such a system would give further <strong>in</strong>sights <strong>in</strong>to<br />

the roles <strong>of</strong> other players <strong>in</strong> the moth olfactory system. Recent research has<br />

demonstrated that B. mori GOBP2 b<strong>in</strong>ds the sex pheromone components <strong>of</strong> B. mori,<br />

rais<strong>in</strong>g the question <strong>of</strong> a possible role <strong>of</strong> moth GOBPs <strong>in</strong> pheromone reception (Zhou<br />

et al., 2009; He et al., 2010). However, B. Mori GOBP2 is not expressed <strong>in</strong><br />

pheromone responsive sensilla (Maida et al., 2005). EpGOBP2 can be further tested<br />

with the sex pheromone components <strong>of</strong> E. postvittana to confirm whether a<br />

pheromone b<strong>in</strong>d<strong>in</strong>g function is unique to the Bombycidae or has been conserved <strong>in</strong> the<br />

Lepidoptera.


4.1 Introduction<br />

4<br />

Identification <strong>of</strong><br />

putative odorant<br />

receptors from Epiphyas<br />

postvittana<br />

The significant role that several members <strong>of</strong> the order Lepidoptera play <strong>in</strong> agriculture<br />

as pests has sparked great <strong>in</strong>terest <strong>in</strong> study<strong>in</strong>g their olfactory system as this is essential<br />

for their detection and location <strong>of</strong> mat<strong>in</strong>g partners and host plants. ORs play a pivotal<br />

role <strong>in</strong> the recognition <strong>of</strong> different odorants and their activation results <strong>in</strong> neuronal<br />

signals which the <strong>in</strong>sect‟s bra<strong>in</strong> <strong>in</strong>terprets (refer to section 1.5.2 for an overview).<br />

Decod<strong>in</strong>g the receptor repertoire <strong>of</strong> the olfactory system will enable the identification<br />

<strong>of</strong> specific receptor–ligand comb<strong>in</strong>ations <strong>of</strong> functional importance to various species<br />

<strong>of</strong> moth. Several different approaches can be employed <strong>in</strong> the isolation <strong>of</strong> ORs and<br />

PRs. These approaches are discussed below.<br />

4.1.1 Transcriptome sequenc<strong>in</strong>g – EST approach<br />

A library <strong>of</strong> cDNAs, synthesised from RNA typically from a specific tissue, is made<br />

by shotgun clon<strong>in</strong>g <strong>of</strong> cDNAs <strong>in</strong>to a plasmid vector, followed by transformation <strong>in</strong>to<br />

suitable host, selection <strong>of</strong> clones, plasmid isolation and sequenc<strong>in</strong>g. A high proportion<br />

<strong>of</strong> DNA <strong>in</strong> higher organisms is non-cod<strong>in</strong>g, with only a small proportion <strong>in</strong>volved <strong>in</strong>


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 80<br />

the cod<strong>in</strong>g RNA and prote<strong>in</strong> (Sterky and Lundeberg, 2000) and the references with<strong>in</strong>).<br />

Therefore, sequenc<strong>in</strong>g <strong>of</strong> cDNA, which is transcribed from mRNA elim<strong>in</strong>ates the bulk<br />

<strong>of</strong> the non-cod<strong>in</strong>g sequences.<br />

The differ<strong>in</strong>g expression levels <strong>of</strong> genes can be problematic when identification <strong>of</strong><br />

lowly expressed genes is desired. cDNA sequenc<strong>in</strong>g can result <strong>in</strong> many copies <strong>of</strong> the<br />

highly expressed genes be<strong>in</strong>g sequenced and few or no copies <strong>of</strong> the lowly expressed<br />

genes. One way to mitigate this problem is to normalise the expression levels <strong>of</strong> all<br />

genes <strong>in</strong> a cDNA sample by reduc<strong>in</strong>g the levels <strong>of</strong> highly expressed genes. A method<br />

developed by Zhulidov et al. (2004) and Zhulidov et al. (2005) permits the<br />

identification <strong>of</strong> rare transcripts <strong>in</strong> a cDNA population. Briefly this method utilises the<br />

specificity <strong>of</strong> duplex-specific nuclease (DSN) for double stranded DNA as a substrate.<br />

Double stranded cDNA (dsDNA) is synthesised from mRNA <strong>of</strong> <strong>in</strong>terest by us<strong>in</strong>g<br />

specific adaptors. The dsDNA is then melted and allowed to rehybridise under<br />

controlled conditions. A higher tendency <strong>of</strong> abundant transcripts to rehybridise is<br />

observed compared to rare transcripts. The dsDNA and s<strong>in</strong>gle-stranded DNA<br />

(ssDNA) population is then treated with DSN, which specifically targets dsDNA for<br />

degradation leav<strong>in</strong>g beh<strong>in</strong>d a population that has normalised levels <strong>of</strong> transcripts. PCR<br />

is then used to generate dsDNA from the ssDNA (Zhulidov et al., 2004; Zhulidov et<br />

al., 2005).<br />

4.1.2 Genome sequenc<strong>in</strong>g<br />

Genomic sequenc<strong>in</strong>g provides sequence data for all the genetic <strong>in</strong>formation conta<strong>in</strong>ed<br />

with<strong>in</strong> the cells <strong>of</strong> the organism <strong>of</strong> <strong>in</strong>terest. It is very useful <strong>in</strong> identify<strong>in</strong>g the genes<br />

conta<strong>in</strong>ed with<strong>in</strong> an organism regardless <strong>of</strong> the expression level <strong>of</strong> the gene (Bork et<br />

al., 1998). However, gene prediction becomes a challenge as cod<strong>in</strong>g regions <strong>of</strong> genes<br />

are flanked by <strong>in</strong>trons and repeats (Gilbert, 1978). The challenge <strong>in</strong> genome<br />

sequenc<strong>in</strong>g is not obta<strong>in</strong><strong>in</strong>g the sequence data itself but assembly <strong>of</strong> the data and<br />

accurate gene prediction. Most eukaryotic cells have two sources <strong>of</strong> DNA, the nucleus<br />

and the mitochondria. A larger quantity <strong>of</strong> mitochondrial DNA (mtDNA) than nucleic<br />

DNA is found <strong>in</strong> cells hence when isolat<strong>in</strong>g genomic DNA for sequenc<strong>in</strong>g and<br />

identification <strong>of</strong> nuclear genes, it is important to reduce the mtDNA content <strong>of</strong> the


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 81<br />

total DNA sample. Two broad approaches have been taken to sequence genomes, a<br />

shotgun approach or a clone-by-clone approach.<br />

In the shotgun approach, the genomic DNA <strong>of</strong> <strong>in</strong>terest is fragmented, then size<br />

selected for library construction (Green, 1997; Weber and Myers, 1997). The ends <strong>of</strong><br />

the DNA are end repaired followed by selection <strong>of</strong> fragments <strong>of</strong> a certa<strong>in</strong> size class<br />

which are cloned <strong>in</strong>to plasmid vectors and sequenced. The <strong>in</strong>serts, depend<strong>in</strong>g on the<br />

plasmid used can range anywhere from 2 kb to 200 kb. In the clone-by-clone<br />

approach, large overlapp<strong>in</strong>g clones (<strong>in</strong>sert size greater than 50 kb) that cover the<br />

whole genome are constructed usually with Bacterial Artificial Chromosomes (BACs)<br />

(Olson et al., 1989; Kim et al., 1996). These BACs or libraries are then <strong>in</strong>dividually<br />

fragmented and sequenced by shotgun clon<strong>in</strong>g, after which each BAC is assembled<br />

locally. The <strong>in</strong>dividual BACs are then assembled to get the genome sequence <strong>of</strong> the<br />

organism. This is helpful when assembl<strong>in</strong>g large eukaryotic genomes with repeats<br />

with limited computational power.<br />

4.1.3 Recent progress <strong>in</strong> sequenc<strong>in</strong>g field<br />

A turn<strong>in</strong>g po<strong>in</strong>t <strong>in</strong> the sequenc<strong>in</strong>g field came from the development <strong>of</strong> low-cost high-<br />

throughput sequenc<strong>in</strong>g mach<strong>in</strong>es. Several sequenc<strong>in</strong>g platforms are now available for<br />

generat<strong>in</strong>g millions <strong>of</strong> bases at a fraction <strong>of</strong> the price <strong>of</strong> traditional Sanger sequenc<strong>in</strong>g<br />

methods, although the low cost is <strong>of</strong>fset by the generation <strong>of</strong> only short read lengths, a<br />

limitation <strong>of</strong> new sequenc<strong>in</strong>g technologies. Intense computer power is required for the<br />

correct assembly <strong>of</strong> thousands <strong>of</strong> short reads, yet another limitation <strong>of</strong> new sequenc<strong>in</strong>g<br />

technologies. Of the available sequenc<strong>in</strong>g platforms <strong>of</strong> Roche 454 (Ronaghi et al.,<br />

1996; Dressman et al., 2003; Margulies, 2005), Illum<strong>in</strong>a Genome Analyzer (Adessi,<br />

2000; Fedurco et al., 2006; Turcatti et al., 2008), ABI/Solid (Shendure, 2005;<br />

McKernan et al., 2006) and Helicos (Braslavsky et al., 2003; Harris, 2008) the first<br />

two platforms have been used for sequenc<strong>in</strong>g the E. postvittana antennal<br />

transcriptome and genome respectively, and will be discussed further.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 82<br />

4.1.3.1 454 Sequenc<strong>in</strong>g<br />

The Roche 454 sequenc<strong>in</strong>g platform uses a sequenc<strong>in</strong>g-by-synthesis approach<br />

(Ronaghi et al., 1996; Dressman et al., 2003; Margulies, 2005). dsDNA is<br />

fragmented, then hydrolysed to give ssDNA and adapters are ligated to both ends <strong>of</strong><br />

the fragments. Each fragment is then attached to a streptavid<strong>in</strong> bead via biot<strong>in</strong>. The<br />

bead is then captured <strong>in</strong> a droplet that has the necessary nucleotides and buffers for<br />

amplification to take place. Each bead with approximately 10 7 copies <strong>of</strong> the orig<strong>in</strong>al<br />

fragment is then deposited <strong>in</strong>to a well <strong>of</strong> a picotitre plate. The wells‟ dimensions are<br />

such that each one can accommodate only a s<strong>in</strong>gle bead, together with the reaction<br />

mix. Pyrosequenc<strong>in</strong>g <strong>of</strong> all the wells occurs <strong>in</strong> parallel, with the release <strong>of</strong> <strong>in</strong>organic<br />

pyrophosphate detected by chemilum<strong>in</strong>escence, the <strong>in</strong>tensity <strong>of</strong> the light <strong>in</strong>dicat<strong>in</strong>g the<br />

number <strong>of</strong> each nucleotide <strong>in</strong>corporated. This be<strong>in</strong>g an estimated value gives rise to<br />

slight errors <strong>in</strong> the sequence reads.<br />

4.1.3.2 Illum<strong>in</strong>a Genome Analyzer<br />

The Solexa technology by Illum<strong>in</strong>a also uses sequenc<strong>in</strong>g-by-synthesis where the<br />

fragmented, adaptor ligated ssDNA is attached to the surface <strong>of</strong> a flow cell which has<br />

a dense lawn <strong>of</strong> primers that b<strong>in</strong>d to the adaptor sequences (Adessi, 2000; Fedurco et<br />

al., 2006; Turcatti et al., 2008). Solid phase bridge amplification is <strong>in</strong>itiated by the<br />

addition <strong>of</strong> nucleotides and enzymes, and the fragment bends over as its<br />

complementary strand is synthesized. The amplification reactions result <strong>in</strong> cluster<br />

formation, each cluster hav<strong>in</strong>g 1000 clonal copies <strong>of</strong> each fragment. 40 million<br />

clusters are generated <strong>in</strong> the amplification <strong>in</strong> each flow cell which are then sequenced<br />

<strong>in</strong> parallel to give reads between 35 and 75bp. This is an effective system but a<br />

drawback is the relative short reads produced.<br />

4.1.3.3 Bio<strong>in</strong>formatics<br />

With billions <strong>of</strong> bases generated from the new sequenc<strong>in</strong>g platforms each day, the<br />

need for improved, powerful computational resources is <strong>in</strong>evitable. The bio<strong>in</strong>formatic<br />

analysis follow<strong>in</strong>g sequence data generation can be divided <strong>in</strong>to the follow<strong>in</strong>g four<br />

processes; assembly, annotation, gene sequence identification and function prediction<br />

<strong>of</strong> the prote<strong>in</strong>s (Sterky and Lundeberg, 2000). The major target <strong>of</strong> the assembly<br />

process is to decrease the number <strong>of</strong> fragments generated by the sequenc<strong>in</strong>g


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 83<br />

technology by merg<strong>in</strong>g the short, overlapp<strong>in</strong>g fragments <strong>in</strong>to longer, mean<strong>in</strong>gful<br />

consensus sequences. The challenge <strong>in</strong> assembl<strong>in</strong>g millions to billions <strong>of</strong> short reads<br />

is the detection and discrim<strong>in</strong>ation <strong>of</strong> sequence ambiguity from regions <strong>of</strong> repeats.<br />

Due to the short read length and error associated with the pyrosequenc<strong>in</strong>g, Sanger<br />

assembly tools cannot be used for assembl<strong>in</strong>g these high throughput sequence data<br />

(Ronaghi et al., 1996). Many different assemblers are been designed to overcome the<br />

bottleneck between sequence data generation and analysis <strong>of</strong> these data to form a<br />

mean<strong>in</strong>gful assembly. The Roche 454 sequenc<strong>in</strong>g platform comes packaged with<br />

Newbler for de novo assembly <strong>of</strong> data generated us<strong>in</strong>g this platform (www.roche-<br />

applied-science.com). This assembler can be used for de novo assembly, as has been<br />

shown for bacteria, if sufficient sequence coverage (25 to 30 times) is achieved (Poly<br />

et al., 2007). Short oligonucleotide analysis package (SOAPdenovo) is specifically<br />

designed for assembl<strong>in</strong>g Illum<strong>in</strong>a reads (soap.genomics.cs.org.cn). Some other<br />

programs that have been designed for de novo assembly <strong>of</strong> 30 to 40 bp fragments are<br />

SSAKE (Warren et al., 2007), VCAKE (Jeck, 2007) and SHARCGS (Dohm et al.,<br />

2007). These utilise a greedy approach, whereby reads are chosen to form seeds and<br />

these are extended <strong>in</strong> both directions by merg<strong>in</strong>g overlapp<strong>in</strong>g reads <strong>in</strong>to the seeds. The<br />

drawback <strong>of</strong> such an assembly process is the short length <strong>of</strong> the contigs generated.<br />

Fragmented assembly and sequenc<strong>in</strong>g errors become a problem <strong>in</strong> gene annotation <strong>of</strong><br />

contigs. Sequence alignment with homologous genes from other organisms can be<br />

used to annotate genes or parts <strong>of</strong> genes even if the full length <strong>of</strong> the gene cannot be<br />

predicted due to sequenc<strong>in</strong>g errors or <strong>in</strong>-frame stop codons (Pop and Salzberg, 2008)<br />

and the references with<strong>in</strong>). Open read<strong>in</strong>g frames (ORFs) <strong>of</strong> 300 bp or more can be<br />

identified as potential genes and annotated to some extend by compar<strong>in</strong>g to<br />

homologous genes <strong>in</strong> databases. The function <strong>of</strong> a potential gene can be predicted<br />

from sequence similarity <strong>in</strong> homologous genes, however, specific functional studies<br />

still have to be conducted <strong>in</strong> order to confirm the function <strong>of</strong> these predicted genes.<br />

4.1.4 Odorant Receptor Identification <strong>in</strong> Moths<br />

The first moth odorant receptors were identified <strong>in</strong> H. virescens through BLAST<br />

search <strong>of</strong> Heliothis low coverage genomic sequence with candidate OR sequences<br />

from Drosophila used as queries (Krieger et al., 2002). The genomic sequences<br />

obta<strong>in</strong>ed from the BLAST search were used as probes to obta<strong>in</strong> the exonic regions <strong>of</strong>


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 84<br />

the sequences from a H. virescens cDNA library. This identified n<strong>in</strong>e OR genes <strong>in</strong> H.<br />

virescens and further screen<strong>in</strong>g <strong>of</strong> the genomic sequences as well as transcriptome<br />

sequenc<strong>in</strong>g have led to the identification <strong>of</strong> a total <strong>of</strong> 21 ORs <strong>in</strong> H. virescens so far<br />

(Krieger et al., 2004). ORs have also been identified from transcriptome sequences <strong>of</strong><br />

A. polyphemus, A. pernyi, and M. sexta. In B. mori, 68 ORs have been identified from<br />

transcriptome and genomic sequences, the highest number <strong>of</strong> ORs identified <strong>in</strong> a moth<br />

so far (Krieger et al., 2005; Nakagawa et al., 2005; Wanner et al., 2007; Tanaka et al.,<br />

2009). Degenerate PCR has also been used to identify moth ORs, for example,<br />

Mitsuno et al. (2008) identified 10 ORs from three different moth species us<strong>in</strong>g<br />

degenerate PCR. However, due to the highly divergent nature <strong>of</strong> ORs, with only 10-<br />

75% am<strong>in</strong>o acid sequence identity between and with<strong>in</strong> them, degenerate PCR has<br />

been least successful <strong>in</strong> identify<strong>in</strong>g new ORs, <strong>in</strong>dicat<strong>in</strong>g the need for more high<br />

throughput sequenc<strong>in</strong>g <strong>in</strong> order to identify more members <strong>of</strong> this highly divergent<br />

group.<br />

From glomeruli studies <strong>in</strong> mammals, it has been observed that each OSN very likely<br />

express only a s<strong>in</strong>gle OR. OSNs express<strong>in</strong>g the same OR are scattered throughout the<br />

epithelium, however they merge to the same glomerulus <strong>in</strong> the olfactory bulb. Hence a<br />

close estimation <strong>of</strong> the number <strong>of</strong> ORs <strong>in</strong> a given organism can be predicted from the<br />

number <strong>of</strong> glomeruli it possess. Glomeruli studies <strong>in</strong> tortricids have shown the<br />

presence <strong>of</strong> anywhere between 48 and 72 glomeruli (Masante-Roca et al., 2005;<br />

Varela et al., 2009). No glomeruli studies have yet been done <strong>in</strong> E. postvittana so the<br />

number <strong>of</strong> ORs could be anywhere between 48 and 72. From an EST library <strong>of</strong> male<br />

E. postvittana antennae, three ORs have been identified (Jordan et al., 2009) (refer to<br />

section 1.7 for more details). ORs and PRs are expressed at low levels <strong>in</strong> tissues<br />

therefore are under-represented <strong>in</strong> EST collections, especially <strong>in</strong> “shallow” Sanger<br />

based collections.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 85<br />

4.1.5 Aims<br />

The aim <strong>of</strong> this research chapter is to <strong>in</strong>crease the receptor repertoire <strong>of</strong> E. postvittana.<br />

Does E. postvittana conta<strong>in</strong> a similar set <strong>of</strong> ORs and PRs compared with B. mori <strong>in</strong><br />

terms <strong>of</strong> the number and sequences <strong>of</strong> ORs, that is, are the E. postvittana ORs one-to-<br />

one orthologs <strong>of</strong> the B. mori ORs; or are the Tortricidae ORs dist<strong>in</strong>ct and different<br />

from those <strong>of</strong> the Bombycidae, hav<strong>in</strong>g radiated <strong>in</strong>dependently? Based on the 65%<br />

am<strong>in</strong>o acid identity (90% similarity) <strong>of</strong> BmOR49J and EpOR3, we might expect other<br />

E. postvittana ORs to follow a similar pattern. Does E. postvittana PR also fall <strong>in</strong> the<br />

PR clade, as has been shown for some other PRs? Do some E. postvittana ORs show<br />

no sex-bias expression while some are more highly expressed <strong>in</strong> males and some <strong>in</strong><br />

females, as is the case with B. mori ORs?<br />

To address these questions multiple approaches are undertaken, each with <strong>in</strong>creas<strong>in</strong>g<br />

power to isolate ORs and PRs. First further screen<strong>in</strong>g <strong>of</strong> the Sanger 1800 ESTs is<br />

done to identify potential PRs by differential screen<strong>in</strong>g with male and female antennal<br />

mRNA on a microarray chip. Such an approach will allow the identification <strong>of</strong><br />

differentially expressed genes even if their sequences are only represented by 3‟UTRs<br />

on the array. An attempt is be made at design<strong>in</strong>g degenerate primers to the PR clade<br />

<strong>of</strong> moths and attempt<strong>in</strong>g to amplify homologous genes from E. postvittana antennal<br />

cDNA. If the above does not result <strong>in</strong> the identification <strong>of</strong> more receptors, then high<br />

throughput sequenc<strong>in</strong>g <strong>of</strong> E. postvittana will be conducted. ORs are expressed at very<br />

low levels so even deep transcriptome sequenc<strong>in</strong>g might not identify them all. It is<br />

hypothesized that normalisation <strong>of</strong> cDNA before sequenc<strong>in</strong>g <strong>in</strong>creases the chance <strong>of</strong><br />

identify<strong>in</strong>g lowly expressed genes. Further ESTs will be obta<strong>in</strong>ed from transcriptome<br />

sequenc<strong>in</strong>g <strong>of</strong> male antennal cDNA and genomic DNA will also be sequenced for<br />

potential ORs us<strong>in</strong>g a low coverage whole genome sequenc<strong>in</strong>g method.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 86<br />

4.2 Materials and Methods<br />

4.2.1 Materials<br />

All the PCR reagents were from Invitrogen unless otherwise stated.<br />

4.2.2 Total RNA Extraction<br />

RNA was extracted from two pools <strong>of</strong> 100 antennae pairs each for male and female E.<br />

postivttana and from two pools <strong>of</strong> moth female body without the head, with three<br />

bodies per pool. TRIzol Reagent (Invitrogen) was used for isolat<strong>in</strong>g RNA follow<strong>in</strong>g<br />

manufacturer‟s <strong>in</strong>structions, and the result<strong>in</strong>g pellet resuspended <strong>in</strong> 20 µL <strong>of</strong> TE<br />

buffer. The concentration <strong>of</strong> the samples was measured on a nanodrop at an optical<br />

density <strong>of</strong> 260-280 (OD260/280) (Thermo Sce<strong>in</strong>tific) and the samples stored at -80°C<br />

until use.<br />

4.2.3 mRNA Isolation<br />

mRNA from male and female E. postvittana antennae was isolated us<strong>in</strong>g the Ambion<br />

MicroPoly(A)Purist Kit (Applied Biosystems) and the quality assessed on an<br />

Agilent 2100 Bioanalyzer, follow<strong>in</strong>g manufacturer‟s <strong>in</strong>structions.<br />

4.2.4 cDNA Synthesis<br />

cDNA was synthesised from 1µg <strong>of</strong> RNA us<strong>in</strong>g either Superscript III reverse<br />

transcript (Invitrogen) or the iscript cDNA synthesis kit (Bio-Rad), follow<strong>in</strong>g<br />

manufacturer‟s <strong>in</strong>structions. A no reverse transcriptase control was <strong>in</strong>cluded for every<br />

sample, and the synthesised cDNA stored at -20°C until required.<br />

4.2.5 Degenerate PCR<br />

To identify conserved regions between E. postvittana OR1 and ORs from three other<br />

<strong>in</strong>sect species belong<strong>in</strong>g to the same phylogenetic clade (M. sexta, B. mori and H.<br />

virescens), multiple sequence alignment <strong>of</strong> the follow<strong>in</strong>g OR prote<strong>in</strong> sequences were<br />

conducted us<strong>in</strong>g clustalX 1.81 (MsOR1, BmOR3, HvOR6, HvOR16, HvOR14,


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 87<br />

HvOR15, HvOR11, BmOR5, BmOR7, HvOR13, BmOR4, BmOR1, EpOR1,<br />

BmOR6) with default sett<strong>in</strong>gs. Two regions each conta<strong>in</strong><strong>in</strong>g the highest levels <strong>of</strong><br />

conservation with<strong>in</strong> the alignment were selected to design outer and <strong>in</strong>ner nested<br />

degenerate primers for degenerate PCR us<strong>in</strong>g the Ro and Ri primers, given <strong>in</strong> Table<br />

4.1 (See appendix B for sequence alignment).<br />

Table 4.1: Outer and nested degenerate primers used <strong>in</strong> degenerate PCR <strong>of</strong> male E.<br />

postvittana antennal cDNA.<br />

Primer Name Primer Sequence (5’ – 3’)<br />

Outer ARNYTNATHGAYGCNGTNTA<br />

Nested GGNYTNCCNTGGGARTRYATGGA<br />

Ro (outer reverse) ATCGATGGTCGACGCATGCGGATCC<br />

Ri (nested reverse) GGATCCAAAGCTTGAATTCGAGCTC<br />

Five microlitres <strong>of</strong> ten-fold-diluted cDNA was used as template <strong>in</strong> the first PCR<br />

together with 1x PCR buffer <strong>in</strong>clud<strong>in</strong>g 1.5 mM magnesium, 0.2 mM <strong>of</strong> each dNTP, 2<br />

units <strong>of</strong> Taq DNA polymerase and 0.5 µM <strong>of</strong> each outer primer. The f<strong>in</strong>al volume was<br />

made up to 50 µL with sterile water. The PCR cycl<strong>in</strong>g conditions were as follows:<br />

<strong>in</strong>itial denature at 94°C for 2 m<strong>in</strong>utes, followed by 35 cycles <strong>of</strong> 94°C for 20 second,<br />

55°C for 30 seconds and 72°C for 1 m<strong>in</strong>utes, and a f<strong>in</strong>al extension at 72°C for 10<br />

m<strong>in</strong>utes. One microlitre <strong>of</strong> product from the first round <strong>of</strong> PCR was used as template<br />

<strong>in</strong> a second round <strong>of</strong> PCR us<strong>in</strong>g the nested primers. The same PCR conditions were<br />

used as <strong>in</strong> the first round <strong>of</strong> PCR. 10 µL <strong>of</strong> PCR product was analysed on 1% agarose<br />

gel with 1x SYBR safe DNA gel sta<strong>in</strong> (Invitrogen) and visualised on the ImageQuant<br />

300 system (GE Healthcare Life Sciences).<br />

4.2.6 Clon<strong>in</strong>g and Sequenc<strong>in</strong>g<br />

Products greater than 500 bp were excised from the gel and DNA extracted us<strong>in</strong>g the<br />

gel purification kit (Qiagen), follow<strong>in</strong>g manufacturer‟s <strong>in</strong>structions, after which they<br />

were ligated <strong>in</strong>to pGEM-T Easy vector (Promega) us<strong>in</strong>g T4 DNA ligase follow<strong>in</strong>g<br />

manufacturer‟s <strong>in</strong>structions. The plasmids were transformed <strong>in</strong>to chemically<br />

competent DH5α cells. Briefly, 2 µL <strong>of</strong> the plasmid was mixed gently with 25 µL


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 88<br />

competent cells <strong>in</strong> a 1.5 mL microcentrifuge tube and <strong>in</strong>cubated on ice for 30 m<strong>in</strong>utes.<br />

The cells were heat shocked at 42°C for 1 m<strong>in</strong>ute, and immediately placed on ice.<br />

After three m<strong>in</strong>utes on ice, 225 µL Luria Broth media was added to the tubes and left<br />

to <strong>in</strong>cubate for one hour at 37°C with shak<strong>in</strong>g at 150 rpm. Fifty microlitres <strong>of</strong> media<br />

solution was then plated on LB agar plates conta<strong>in</strong><strong>in</strong>g 100 µg/mL ampicill<strong>in</strong> and 40<br />

µg/mL X-galactose (for blue/white screen<strong>in</strong>g). After overnight <strong>in</strong>cubation at 37°C<br />

white colonies were screened by PCR with plasmid specific M13 primers by<br />

transferr<strong>in</strong>g a small portion <strong>of</strong> the colony with a pipette tip and dipp<strong>in</strong>g it <strong>in</strong> a PCR<br />

reaction mix conta<strong>in</strong><strong>in</strong>g 1x PCR buffer <strong>in</strong>clud<strong>in</strong>g 1.5 mM magnesium, 0.2 mM <strong>of</strong><br />

each dNTP, 2 units <strong>of</strong> Taq DNA polymerase and water up to 20 µL. The PCR cycl<strong>in</strong>g<br />

conditions were as follows: <strong>in</strong>itial denature at 96°C for 2 m<strong>in</strong>utes, then 30 cycles <strong>of</strong><br />

94°C for 30 seconds, 58°C for 30 seconds, 72°C for one m<strong>in</strong>ute and a f<strong>in</strong>al extension<br />

<strong>of</strong> 72°C for 10 m<strong>in</strong>utes. The result<strong>in</strong>g products were visualised on agarose gel as<br />

before and colonies conta<strong>in</strong><strong>in</strong>g an <strong>in</strong>sert greater than 500 bp were picked for<br />

sequenc<strong>in</strong>g. Briefly, the colonies were grown <strong>in</strong> 10 mL LB media supplemented with<br />

100 µg/mL ampicill<strong>in</strong> overnight at 37°C <strong>in</strong> a shak<strong>in</strong>g <strong>in</strong>cubator at 220 rpm. The<br />

plasmid was isolated us<strong>in</strong>g the QIAprep Sp<strong>in</strong> M<strong>in</strong>iprep Kit (Qiagen) and sequenced <strong>in</strong><br />

both directions us<strong>in</strong>g BigDye TM Term<strong>in</strong>ator Version 3.1 Ready Reaction Cycle<br />

Sequenc<strong>in</strong>g kit (Applied Biosystems) with capillary analysis performed on the<br />

ABI3730 DNA Analyzer (Applied Biosystems) at the Allan Wilson Centre<br />

Sequenc<strong>in</strong>g Services, Palmerston North. The sequences obta<strong>in</strong>ed were analysed <strong>in</strong><br />

Sequencher v4.2 s<strong>of</strong>tware (Gene Codes) by align<strong>in</strong>g both the sequence strands and to<br />

confirm that the overlapp<strong>in</strong>g regions matched. These sequences were then searched<br />

aga<strong>in</strong>st all the available sequences <strong>in</strong> the NCBI BLAST server (Altschul et al., 1990)<br />

us<strong>in</strong>g the tblastx algorithm to identify homologous genes and confirm the identity <strong>of</strong><br />

the cloned genes.<br />

4.2.7 Microarray Screen<strong>in</strong>g<br />

For differential microarray screen<strong>in</strong>g <strong>of</strong> the male antennal EST library to identify<br />

potential PR candidates, each <strong>of</strong> the 4472 male antennal EST oligonucleotides were<br />

checked manually to see if the oligonucleotide was <strong>in</strong> an acceptable region <strong>of</strong> the<br />

sequence trace file. Candidates for spott<strong>in</strong>g onto a microarray chip were selected<br />

based on their trace file be<strong>in</strong>g good and the putative function <strong>of</strong> the EST from


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 89<br />

homology searches <strong>in</strong> NCBI databases yield<strong>in</strong>g prote<strong>in</strong>s with olfactory function or for<br />

whom a function could not be determ<strong>in</strong>ed. 1809 candidates were selected for which<br />

oligos were designed, synthesized and spotted onto a microarray chip for<br />

hybridisation screen<strong>in</strong>g with male and female E. postvittana antennae mRNA. 250 ng<br />

<strong>of</strong> male and female antennae mRNA each, were labelled with both Cy3 and Cy5 to be<br />

used <strong>in</strong> a dye swap hybridisation (Cy3 male vs Cy5 female and Cy3 female vs Cy5<br />

male). The microarray hybridisation <strong>of</strong> the oligonucleotides spotted on the chip with<br />

male and female antennae mRNA was performed at the microarray facility at Plant<br />

and Food Research, Mt Albert, Auckland, as outl<strong>in</strong>ed <strong>in</strong> Janssen et al. (2008).<br />

4.2.8 Microarray Data Analysis<br />

The fluorescence scores from the microarray hybridisation were checked for male-<br />

biased expression and confirmed by check<strong>in</strong>g the correspond<strong>in</strong>g spots by eye. An<br />

oligo was considered to have male antennae biased expression if the same results were<br />

obta<strong>in</strong>ed for the dye swap. Genes with oligos show<strong>in</strong>g differential (if fluorescence for<br />

male was ≥ to 2 times that <strong>of</strong> female) expression were further tested for male-biased<br />

expression by qRT-PCR. The primer design, primer test<strong>in</strong>g and qRT-PCR is outl<strong>in</strong>ed<br />

<strong>in</strong> sections 4.2.12 – 4.2.14 and the primer pairs are given <strong>in</strong> Table 4.2.<br />

4.2.9 Transcriptome Sequenc<strong>in</strong>g<br />

5 µg <strong>of</strong> male E. postvittana antennal total RNA with an OD260/280 <strong>of</strong> 1.96 was sent to<br />

Evrogen (www.evrogen.com) for synthesis<strong>in</strong>g directionally normalised cDNA, us<strong>in</strong>g<br />

the SMART approach and the enzyme DSN, as outl<strong>in</strong>ed <strong>in</strong> section 4.1.1. The<br />

normalised sample was then sequenced on the Roche 454 GS FLX Titanium at the<br />

Department <strong>of</strong> Anatomy and Structural Biology at The University <strong>of</strong> Otago. The<br />

sequence data was assembled us<strong>in</strong>g the Newbler assembler by the Bio<strong>in</strong>formatics<br />

department at Plant and Food Research Ltd, Mt Albert, Auckland. The contigs and<br />

s<strong>in</strong>gletons were deposited <strong>in</strong>to BioView, a privately held database at Plant and Food<br />

Research Ltd. B. mori OR am<strong>in</strong>o acid sequences were used to perform tblastn<br />

searches <strong>of</strong> the E. postvittana Insect Database (conta<strong>in</strong><strong>in</strong>g the male antennal contigs<br />

and s<strong>in</strong>gletons) and best hit ESTs with e-value less than 0.5 were taken as significant<br />

and selected for further analysis.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 90<br />

4.2.10 Genome Sequenc<strong>in</strong>g<br />

For genomic sequenc<strong>in</strong>g <strong>of</strong> E. postvittana, the genomic DNA (gDNA) was extracted<br />

from purified nuclei. This was done to reduce the amount <strong>of</strong> mitochondrial DNA<br />

(mtDNA) be<strong>in</strong>g sequenced. The nuclei was isolated first before gDNA extraction as<br />

outl<strong>in</strong>ed <strong>in</strong> section 4.2.11. Library construction (us<strong>in</strong>g the Paired-End DNA sample<br />

prep kit VI from Illum<strong>in</strong>a), cluster generation and genomic analysis was carried out at<br />

the Allan Wilson Centre for Genome Sequence (AWCGS, Palmerston North). A total<br />

<strong>of</strong> 8 lanes <strong>of</strong> 75 bp paired-end genome sequenc<strong>in</strong>g were done on the Illum<strong>in</strong>a 1G<br />

Genome Analyzer (Illum<strong>in</strong>a) at AWCGS. Pipel<strong>in</strong>e analysis was done at the<br />

Bio<strong>in</strong>formatics department at Plant and Food Research Ltd by Ross Crowhurst. The<br />

sequence data was assembled us<strong>in</strong>g SOAPdenovo assembler (Li et al., 2008; Li et al.,<br />

2009). Result<strong>in</strong>g scaffolds were deposited <strong>in</strong>to the genome server held at Plant and<br />

Food Research Ltd. The putative OR sequences identified from the transcriptome<br />

data, together with known B. mori OR sequences (Krieger et al., 2005; Nakagawa et<br />

al., 2005; Wanner et al., 2007) were used for search<strong>in</strong>g the genomic scaffolds for their<br />

E. postivttana homologues us<strong>in</strong>g tblastn. Scaffold hits with e-value less than 0.5 were<br />

taken as significant.<br />

4.2.11 Nuclear DNA isolation and mitochondrial DNA<br />

contam<strong>in</strong>ation test<br />

[This protocol has been modified from the method by (Guo; Lopez-Gomez and<br />

Gomez-Lim., 1992)].<br />

4.2.11.1 Isolation <strong>of</strong> <strong>in</strong>sect nuclei<br />

Refer to Appendix C for buffer recipes used <strong>in</strong> this section. Three grams <strong>of</strong> female E.<br />

postvittana pupae were ground <strong>in</strong> a mortar and pestle <strong>in</strong> liquid nitrogen and<br />

transferred to a beaker with 300 mL <strong>of</strong> ice-cold nucleic extraction (NEB) complete<br />

buffer. The result<strong>in</strong>g homogenate was filtered through 4-6 layers <strong>of</strong> cheesecloth <strong>in</strong>to a<br />

sterile glass beaker on ice. The filtration step was repeated through 2-4 layers <strong>of</strong><br />

miracloth <strong>in</strong>to a 500 mL sterile glass cyl<strong>in</strong>der. The volume was adjusted to 294 mL<br />

with NEB complete buffer and 6 mL <strong>of</strong> 25% Triton X-100 <strong>in</strong> NEB complete buffer.<br />

The cyl<strong>in</strong>der was sealed with parafilm and mixed very gently by <strong>in</strong>version 10-20


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 91<br />

times. The homogenate was divided <strong>in</strong>to six 50 mL Falcon tubes and spun down at<br />

588 rpm at 4-10°C, for 2 m<strong>in</strong>utes. The supernatant was transferred to a new set <strong>of</strong> 50<br />

mL Falcon tubes and centrifuged at 3300 rpm, 4-10°C for 15 m<strong>in</strong>utes. A reddish<br />

pellet was clearly visible. The supernatant was transferred to the waste conta<strong>in</strong>er and<br />

each pellet resuspended with 50 mL NEB– no -mercaptoethanol and mixed gently<br />

by <strong>in</strong>version until the pellet was resuspended. The samples were spun down as before<br />

and the supernatant transferred to the waste conta<strong>in</strong>er. Each pellet was resuspended<br />

with 5 mL NEB– no -mercaptoethanol and all pellets collected <strong>in</strong>to one weighed<br />

Falcon tube. More NEB– no -mercaptoethanol was added to a f<strong>in</strong>al volume <strong>of</strong> 50<br />

mL, spun down aga<strong>in</strong> as before and the supernatant transferred to waste conta<strong>in</strong>er.<br />

The tube was weighed to f<strong>in</strong>d out the total amount <strong>of</strong> nuclei isolated and the tube kept<br />

on ice.<br />

4.2.11.2 Genomic DNA extraction<br />

0.2 grams <strong>of</strong> nuclei pellet was resuspended <strong>in</strong> 14 mL lysis buffer and mixed by<br />

<strong>in</strong>version. 14 µL <strong>of</strong> RNase A 50 mg/mL (Qiagen) was added and mixed by <strong>in</strong>version,<br />

followed by <strong>in</strong>cubation at 37°C with gentle shak<strong>in</strong>g for 10 m<strong>in</strong>utes. 1.4 mL 5 M<br />

potassium acetate pH 7 was then added to it followed by addition <strong>of</strong> 3.5 mL 100%<br />

ethanol. The sample was vortexed at maximum speed for 30 seconds and extracted<br />

with equal volume <strong>of</strong> chlor<strong>of</strong>orm:isoamyl alcohol (24:1). The tube was then placed<br />

horizontally <strong>in</strong> a platform shaker at room temperature and shook gently for 5 m<strong>in</strong>utes.<br />

The sample was spun down at 3000 rpm for 10 m<strong>in</strong>utes at room temperature and the<br />

supernatant collected and placed <strong>in</strong> a new 50 mL Falcon tube. Equal volume <strong>of</strong> ice<br />

cold isopropanol was added to the sample and mixed by <strong>in</strong>version, followed by<br />

<strong>in</strong>cubation at -20°C for 30 m<strong>in</strong>utes. The sample was spun down as before, the pellet<br />

washed with same volume <strong>of</strong> 70% ethanol. The DNA pellet was air dried at room<br />

temperature for 20-30 m<strong>in</strong>utes and resuspended <strong>in</strong> 100–200 µL TE buffer pH 7.5. The<br />

DNA quality was checked at OD260/280 on the nanodrop (Thermo Scientific).<br />

4.2.11.3 Mitochondrial DNA contam<strong>in</strong>ation test<br />

A ten-fold serial dilution <strong>of</strong> the gDNA was made from 1 <strong>in</strong> 10 to 1 <strong>in</strong> 100 million with<br />

TE buffer. The primer pairs MT6 forward 5‟-<br />

GGAGGATTTGGAAATTGATTAGTTCC-3‟ and MT9 reverse 5‟-


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 92<br />

CCCGGTAAAATTAAAATATAAACTTC-3‟, specific for the cytochrome oxidase I<br />

gene <strong>of</strong> the mitochondrial genome <strong>of</strong> E. postvittana, were used <strong>in</strong> a PCR reaction to<br />

test for mtDNA contam<strong>in</strong>ation <strong>in</strong> the gDNA sample. Primers for Takeout 3, which is a<br />

nuclear gene, were used for a direct comparison <strong>of</strong> the nuclear versus mtDNA (5‟-<br />

AAAGGCGAGGGTCACTACAAG-3‟ TO3 forward primer and 5‟-<br />

GGTTCTTCAGTGCGTACACCT-3‟ TO3 reverse primer). The PCR reaction mix<br />

was prepared and the first round <strong>of</strong> PCR conducted, as stated <strong>in</strong> section 4.2.5. The<br />

products were analysed on 1% agarose gel with 1x SYBR safe DNA gel sta<strong>in</strong><br />

(Invitrogen) and visualised on the ImageQuant 300 system (GE Healthcare Life<br />

Sciences). One lane <strong>of</strong> 75 bp paired-end genome sequenc<strong>in</strong>g was done as stated <strong>in</strong><br />

section 4.2.10 and Bowtie, a short read aligner (Langmead et al., 2009) was used for<br />

mapp<strong>in</strong>g mitochondrial sequences <strong>in</strong> GenBank and a 2,216 base pair region <strong>of</strong> the E.<br />

postvittana mitochondrial genome conta<strong>in</strong><strong>in</strong>g the cytochrome oxidase I and II genes<br />

to the 13 million paired-end reads obta<strong>in</strong>ed from one lane <strong>of</strong> Illum<strong>in</strong>a sequenc<strong>in</strong>g.<br />

4.2.12 qRT-PCR primer design<br />

Primers were designed us<strong>in</strong>g Oligo Explorer 1.1.0 (Kuulasmaa, 2000) and where<br />

possible primers were designed us<strong>in</strong>g the follow<strong>in</strong>g str<strong>in</strong>gent conditions: optimal<br />

temperature between 50-60°C, GC% content between 40-60%, and at least 3 GC<br />

d<strong>in</strong>ucleotides at the 3‟end for primer stability. The primers were between 20–23 bp <strong>in</strong><br />

length and the amplicon size was between 100–200 bp. The primers were made such<br />

that the formation <strong>of</strong> primer dimers are m<strong>in</strong>imised and the primers do not self anneal.<br />

The primer pairs for the candidate OR sequences identified from the microarray<br />

screen<strong>in</strong>g and transcriptome sequenc<strong>in</strong>g are given <strong>in</strong> Table 4.2.


Microarray OR<br />

454 Transcriptome OR candidate ESTs<br />

Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 93<br />

Table 4.2: Primer pairs for qRT-PCR amplification <strong>of</strong> products from OR candidates<br />

identified from microarray screen<strong>in</strong>g and 454 transcriptome sequenc<strong>in</strong>g <strong>of</strong> male E.<br />

postvittana antennae.<br />

candidate ESTs<br />

EST Number Primer Forward (5’ – 3’) Primer Reverse (5’ – 3’)<br />

23893 AGGCATCACCAAAGGTAAG CGATAGAAACGCTGTCATTG<br />

25116 GTAATACGCCTTCTGAATGC CCAAAGCGTATCAAAGTTCC<br />

25174 GCAACTCATCAGAAAACTCG AGACAGCAGAGTGAAATCCAG<br />

25214 GCCCCTTGTATTATTATTGGTC AGATAGGTCTGCCTTTTACACG<br />

1032444 CACTGCGGGTATGTTTCCTG TACCCTGCCCTTGATTTCG<br />

1037459 CGTGCTCATCCAAAGATCC ATGCCTTGCCGAAGTCATC<br />

1051140 GCGAGCTGCGTAAAGTTGG TGGAATACAGGGGTCTGGTC<br />

1041667 CTCGGCAACAAATAAAACGC ACCATTTGCTCTCATAAGGC<br />

1043845 AGACGCAGCGTACAAAAGC AATAACCCATGGCAGACAAC<br />

1042257 TTTATTCGGGGCACAAACG GGATTCCGCTGAAAGGATTC<br />

1034291 AACTGGGCTTAAGTAAGTGC TAGCCCTGGAATGAAACAC<br />

1037304 ATGCCTTTCGCTTTGATGC TCCCCACTTGCTATCGTAAC<br />

1040652 CGAGACGTTTTCTTGTGTC ATAGGTCAGCCGACTGTAG<br />

1034929 ACTACGATGCCACAGAAAGG GCCCAGAGGTAGCGATATG<br />

1218954 TCCGTTTCCTTGACATAGG ATTGAGGCGAGGCTGATAG<br />

1212362 TCAATGGGGGGAAGTAATAG TGAGCTGCGGTATATTTCTG<br />

1209721 TTGGGAGTCGACATCTGTG CCACCGACATTGTTTTCTGC<br />

1037495 CCAATACAATCAGGTGAAGC CGTTGTCTGTCGGTCTTAG<br />

1234206 CAAACGCCCCAGTACCAAG CTCTCGCTCACCACTTTCAC<br />

1233473 TAGCACGCTGACATTGATG TACTTCACTGGTTCTGTGGC<br />

1039296 AACGAGGTTGCTGTTGAAG ACATACGCGCCTTACTACG<br />

1212480 CAAACGGCATCCAAGAACTG GTGACGGTCACCACAACAC<br />

1207127 GGACGAGTCACTGAACTGC CGGGATACCCATTAGAGAG<br />

1226006 GCCCTTCTCTACACTCACTG CTGCGTCTAATCTTCGTTG<br />

1043233 CTTCCCCAAGATGAAATGCC GACCCCTAACCCAACTGATG<br />

1201411 TGGCTTCGCTTATATGGAC CCCTAATGGTGGTTGATGG<br />

1049412 AGGTCTGGCAATCATCTCAG GGTGCCGTGTCTTATTATG<br />

1041985 GCTCCAAGGTGAGTACAAC CTGGCTGAGCTAAATCTGC<br />

1046227 ACCAAGGCCAAAGAGAGAAC CTTCGACTGGCAATAAGGTG<br />

1039491 GCCAGCTGCACTCATCAAG GCTGGCGATTCTCCTATTC


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 94<br />

4.2.13 qRT-PCR primer test<br />

Each set <strong>of</strong> primers were used with the cDNA samples from the three different tissues<br />

<strong>in</strong> a gradient cycler to optimise the PCR conditions and obta<strong>in</strong> the optimal anneal<strong>in</strong>g<br />

temperature for each. The standard PCR reaction components were as follows: 1x<br />

PCR buffer <strong>in</strong>clud<strong>in</strong>g 1.5 mM magnesium, 0.2 mM <strong>of</strong> each dNTP, 2 units <strong>of</strong> Taq<br />

DNA polymerase, 0.5 µM <strong>of</strong> each primer, 30 ng cDNA template and water added to a<br />

reaction volume <strong>of</strong> 20 µL. The PCR cycl<strong>in</strong>g conditions were as follows: 95 o C for 2<br />

m<strong>in</strong>utes, followed by 40 cycles at 94 o C for 15 seconds, gradient from 50 to 60 o C for<br />

20 seconds, 72 o C for 30 seconds and a f<strong>in</strong>al elongation at 72 o C for 10 m<strong>in</strong>utes. The<br />

gradient tested was from 50 o C <strong>in</strong> column 1 through to 60 o C <strong>in</strong> column 12. PCR<br />

products were visualised on 1% agarose gel with 1x SYBR safe DNA gel sta<strong>in</strong><br />

(Invitrogen) and visualised on the ImageQuant 300 system (GE Healthcare Life<br />

Sciences).<br />

4.2.14 Quantitative Real-Time PCR<br />

Quantitative real-time PCR were performed <strong>in</strong> 384 well plates on either Applied<br />

Biosystems 7900HT Fast Real-Time PCR System (Applied Biosystems) or on the<br />

Roche LightCycler 480 <strong>in</strong>strument (Roche, Germany), with SYBR green as the<br />

fluorescent reporter. Each primer template comb<strong>in</strong>ation was set up <strong>in</strong> triplicate and a<br />

no template water control was <strong>in</strong>cluded for each primer pair, (two pools <strong>of</strong> each<br />

template giv<strong>in</strong>g a total <strong>of</strong> six different templates). The housekeep<strong>in</strong>g genes Ef1α and<br />

α-tubul<strong>in</strong> (Turner et al., 2006) were used due to their consistent expression <strong>in</strong><br />

different E. postvittana tissues.<br />

For qRT-PCR performed on Applied Biosystems 7900HT Fast Real-Time PCR<br />

System, the plate and <strong>in</strong>strument set up were designed us<strong>in</strong>g the SDS 2.1 s<strong>of</strong>tware<br />

(Applied Biosystems). The assay type performed was absolute quantification followed<br />

by a dissociation curve analysis at the end <strong>of</strong> the run to check for s<strong>in</strong>gle PCR<br />

products. The reaction conditions were same as that for primer tests except for the<br />

addition <strong>of</strong> 0.2 µL <strong>of</strong> 1/1000 SYBR green dye (Molecular Probes). The PCR cycl<strong>in</strong>g<br />

conditions were 95 o C for 2 m<strong>in</strong>utes, followed by 40 cycles at 94 o C for 15 seconds,<br />

55 o C for 20 seconds, 72 o C for 30 seconds followed by the generation <strong>of</strong> a dissociation


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 95<br />

curve with the follow<strong>in</strong>g thermal pr<strong>of</strong>ile; 95°C for 15 seconds, 60°C for 15 seconds,<br />

95°C for 15 seconds, with a ramp rate <strong>of</strong> 2%. The result files obta<strong>in</strong>ed from the run<br />

were loaded <strong>in</strong>to the SDS 2.1 s<strong>of</strong>tware and the amplification and dissociation curves<br />

were checked for consistency between the triplicates. Wells that had more than one<br />

peak <strong>in</strong> the dissociation curve, hence more than one PCR product were omitted from<br />

further analysis. Primer PCR efficiencies were calculated us<strong>in</strong>g the L<strong>in</strong>RegPCR<br />

(Ramakers et al., 2003) program and the Cycle Threshold (Ct) values for all the<br />

samples were converted to quantity <strong>of</strong> product formed <strong>in</strong> each sample by the<br />

follow<strong>in</strong>g equation:<br />

Quantity = (PCR efficiency)<br />

(m<strong>in</strong>imum quantity for a particular treatment – current Ct value)<br />

The mean quantities <strong>of</strong> the housekeep<strong>in</strong>g genes for each tissue type were used with<br />

the s<strong>of</strong>tware geNorm (Vandesompele et al., 2002) for calculat<strong>in</strong>g the normalisation<br />

factor for the <strong>in</strong>dividual tissue types. These normalisation factors were then used for<br />

calculat<strong>in</strong>g the relative expression <strong>of</strong> the samples.<br />

For qRT-PCR performed on a Roche LightCycler 480 <strong>in</strong>strument, the reaction<br />

mixture consisted <strong>of</strong> 5 µL 2x FastStart SYBR Green Master (Roche, Germany), 0.5<br />

µM <strong>of</strong> each primer and 30 ng <strong>of</strong> cDNA template, <strong>in</strong> a total reaction volume <strong>of</strong> 10 µL<br />

made up with nuclease-free water. The thermal pr<strong>of</strong>ile <strong>of</strong> the run was as follows: 95 o C<br />

for 5 m<strong>in</strong>utes at ramp rate ( o C/s) <strong>of</strong> 4.8, then 40 cycles at 95 o C for 15 seconds at ramp<br />

rate <strong>of</strong> 4.8, 55 o C for 30 seconds at ramp rate <strong>of</strong> 2.5 and 72 o C for 20 seconds at ramp<br />

rate <strong>of</strong> 4.8. A melt curve was then generated at the end <strong>of</strong> the amplification to verify<br />

s<strong>in</strong>gle products with the follow<strong>in</strong>g pr<strong>of</strong>ile: 95 o C for 5 seconds at a ramp rate <strong>of</strong> 4.8,<br />

65 o C for 1 m<strong>in</strong>ute at ramp rate <strong>of</strong> 2.5 followed by cont<strong>in</strong>uous acquisition at 97 o C at a<br />

ramp rate <strong>of</strong> 0.11. Analysis <strong>of</strong> the raw data was carried out with the LightCycler 480<br />

s<strong>of</strong>tware version 1.5 (Roche). PCR efficiencies <strong>of</strong> the primers were calculated us<strong>in</strong>g<br />

the exported amplification data at every cycle us<strong>in</strong>g the program L<strong>in</strong>RegPCR<br />

(Ramakers et al., 2003). The relative quantification <strong>of</strong> the test genes was calculated<br />

based on relative expression <strong>of</strong> target versus housekeep<strong>in</strong>g genes, as stated <strong>in</strong> Pfaffl<br />

(2001).


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 96<br />

4.2.15 Rapid Amplification <strong>of</strong> cDNA Ends (RACE)<br />

For each 3‟RACE-PCR reaction, 3 µL <strong>of</strong> ten-fold diluted male E. postvittana antennal<br />

cDNA was used together with a gene specific forward primer (designed from a known<br />

region on the EST <strong>of</strong> <strong>in</strong>terest) and the reverse oligo dT primer (RoRidT16) that is<br />

specific for the mRNA poly A + tail. The gene specific forward primers used are given<br />

<strong>in</strong> Table 4.3. The reaction mix consisted <strong>of</strong> 1x PCR buffer, 1.5 mM magnesium, 0.2<br />

mM <strong>of</strong> each dNTP, 2 units <strong>of</strong> Taq DNA polymerase, 0.5 µM <strong>of</strong> each primer. The f<strong>in</strong>al<br />

volume was made up to 25 µL with sterile water. Touchdown PCR was performed as<br />

follows: <strong>in</strong>itial denature at 94 o C for 4 m<strong>in</strong>utes followed by 20 cycles at 94 o C for 30<br />

seconds, 60-50 o C for 30 seconds and 72 o C for 30 seconds. This was followed by<br />

another 10 cycles at 94 o C for 30 seconds, 50 o C for 30 seconds and 72 o C for 30<br />

seconds. A f<strong>in</strong>al extension was carried out at 72 o C for 7 m<strong>in</strong>utes. The products were<br />

analysed on 1% agarose gel. Products greater than 300 bp were selected and<br />

sequenced.<br />

Table 4.3: Gene specific primers used for amplify<strong>in</strong>g 3‟RACE products from OR<br />

candidates from the 454 transcriptome assembled contigs.<br />

Putative OR Gene Gene Specific Primer (5’ – 3’)<br />

EpOR7 GGATCCTAACGCATTGGCAAAGT<br />

EpOR23 GCAGCGTACAAAAGCCTGTGGTA<br />

EpOR24 GAGTGGAATTCGCTCCCATCGTC<br />

EpOR30 TATTTACAGCTTGATCGCTGTCG<br />

EpOR32 AACTTTACGCAGCTCGCCGCCAT<br />

EpOR33 GATACGAGCGACGAGGTCCCTCT<br />

EpOR34 GCTTATTCAGCCTACGCTGGAGA<br />

EpOR39 GGAGGATTGGAGCACACGTATAT<br />

EpOR41 GATGGGGCCTTATGAGAGCAAAT<br />

EpOR48 GACATTTCTTCATAGGTCCGTCG<br />

EpOR49 ATGCGCGAGGACTACAGTCGGCT<br />

EpOR50 GTTTTGATTTGTATCTGTGCGCC<br />

EpOR51 GGAGTTTTAGGGCAAACATAGTT<br />

EpOR52 TGATTGCCGTGCTGCTGGCGATT


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 97<br />

5‟ RACE was used to isolate the 5‟ ends <strong>of</strong> the cod<strong>in</strong>g regions <strong>of</strong> the genes given <strong>in</strong><br />

Table 4.4, each PR candidate from the microarray analysis and 454 transcriptomics<br />

sampl<strong>in</strong>g based on their tissue specific expressions <strong>in</strong> male E. postvittana antennae (as<br />

shown <strong>in</strong> Figures 4.1 and 4.4). The 5‟RACE System for Rapid Amplification <strong>of</strong><br />

cDNA ends, Version 2.0 (Invitrogen) was employed accord<strong>in</strong>g to manufacturer‟s<br />

<strong>in</strong>structions.<br />

First strand cDNA was synthesized from male antennae E. postvittana RNA us<strong>in</strong>g the<br />

outer gene specific primers (GSP) given <strong>in</strong> Table 4.4. This was done for each <strong>of</strong><br />

<strong>in</strong>dividual ESTs given <strong>in</strong> Table 4.4 so as to synthesise cDNA specific for these ESTs,<br />

thereby enrich<strong>in</strong>g the cDNA for the particular gene. The cDNA synthesis was done<br />

us<strong>in</strong>g either superscript III reverse transcriptase (Invitrogen), or with iscript reverse<br />

transcriptase (Bio-Rad) follow<strong>in</strong>g manufacturer‟s <strong>in</strong>structions, except that the oligo<br />

dT primer was replaced by the primer GSP. The synthesized cDNA was then purified<br />

us<strong>in</strong>g the PCR purification kit (Roche) follow<strong>in</strong>g manufacturer‟s <strong>in</strong>structions. The<br />

cDNA was tailed at the 3‟end with Term<strong>in</strong>al Transferase (TdT) (Roche). This adds a<br />

homopolymeric tail <strong>of</strong> cyste<strong>in</strong>e residues at the 3‟end <strong>of</strong> the cDNA hence a primer that<br />

b<strong>in</strong>ds to these residues can be used to amplify up the target gene with GSP at the other<br />

end. A 20 µL tail<strong>in</strong>g reaction was carried out as follows: 9 µL purified cDNA, 4 µL <strong>of</strong><br />

5x tail<strong>in</strong>g buffer, 2 µL <strong>of</strong> 10 mM dCTP, 4 µL <strong>of</strong> 25 mM CoCl2 and 1 µL <strong>of</strong> TdT<br />

enzyme were mixed and <strong>in</strong>cubated for 1 hour at 37°C. This was followed by heat<br />

<strong>in</strong>activation <strong>of</strong> the TdT at 65°C for 15 m<strong>in</strong>utes. Each <strong>of</strong> the tailed cDNA were then<br />

used <strong>in</strong> 5‟RACE reaction together with their respective outer GSPs given <strong>in</strong> Table 4.4.<br />

A 50 µL PCR reaction mix was setup as <strong>in</strong> section 4.2.5 with 5 µL <strong>of</strong> the tailed cDNA<br />

with 0.5 µM <strong>of</strong> its correspond<strong>in</strong>g outer GSP and 0.5 µM <strong>of</strong> AAP primer (Table 4.4).<br />

The follow<strong>in</strong>g PCR cycl<strong>in</strong>g conditions were used: 94°C for 2 m<strong>in</strong>utes, followed by 30<br />

cycles <strong>of</strong> 94°C for 10 seconds, 50°C for 30 seconds, 72°C for 1 m<strong>in</strong>ute and a f<strong>in</strong>al<br />

extension was carried out at 72°C for 10 m<strong>in</strong>utes. One microlitre <strong>of</strong> this PCR product<br />

was used as template <strong>in</strong> a second round <strong>of</strong> amplification with the nested GSP and<br />

AUAP primer (Table 4.4), us<strong>in</strong>g the same reaction components as before. Products<br />

from this second round <strong>of</strong> PCR amplification were analysed on 1% agarose gel and<br />

products greater than 500 bp were excised from the gel, the DNA extracted and<br />

cloned and sequenced as stated <strong>in</strong> section 4.2.6.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 98<br />

Table 4.4: Primers used for amplify<strong>in</strong>g 5‟ RACE products from the two candidate<br />

PRs identified from microarray screen<strong>in</strong>g and the three PR candidates identified from<br />

the 454 transcriptome sequenc<strong>in</strong>g <strong>of</strong> E. postvittana.<br />

EST Identified<br />

from:<br />

Microarray<br />

screen<strong>in</strong>g<br />

454<br />

Transcriptome<br />

sequenc<strong>in</strong>g<br />

Primer Nucleotide Sequence (5’ – 3’)<br />

25174 (outer) AGATTTCTCTGATTATTTACACAATGTATA<br />

25174 (nested) GTTGCACAGTGGCGACAGACCGGCACTTTT<br />

25214 (outer) TTATTACTAAAATTACCCTCGTAACAACAA<br />

25214 (nested) CCCGAATGATATTCCAAATGCTTACAAACT<br />

1032444 (outer) GTGGCCCAAAAGGCGCAAAGCCCGAAT<br />

1032444 (nested) GCCATAACTTCGATACCCACTGCCCTC<br />

1037459 (outer) GCGGGAACAGCGCGCAGTTCACACAAC<br />

1037459 (nested) GAAGTCATCCCCTTTATGTCGACAGCG<br />

1042257 (outer) GTATGAGTATTGTACCGTTTGTGCCCCG<br />

1042257 (nested) CTTCCAGGCAGACAGCCCTGAACACCG<br />

Abridged anchor primer (AAP) GGCCACGCGTCGACTAGTACGGGIIGGGIIGGGIIG<br />

Abridged universal amplification<br />

primer (AUAP)<br />

4.2.16 Bio<strong>in</strong>formatics<br />

GGCCACGCGTCGACTAGTAC<br />

The predicted OR prote<strong>in</strong>s from the low coverage genomic scaffolds were translated<br />

<strong>in</strong>to the six read<strong>in</strong>g frames and compared with the B. mori OR exons for determ<strong>in</strong><strong>in</strong>g<br />

the exons <strong>of</strong> the prote<strong>in</strong>s. The program SplicePredictor<br />

(http://deepc2.psi.iastate.edu/cgi-b<strong>in</strong>/sp.cgi) was also used for predict<strong>in</strong>g the<br />

exon/<strong>in</strong>tron boundaries. A phylogenetic tree was constructed us<strong>in</strong>g the PHYLIP3.6a3<br />

package (Felsenste<strong>in</strong>, 2005) <strong>of</strong> all the known and putative E. postvittana ORs together<br />

with the ORs from B. mori (Warren et al., 2007; Tanaka et al., 2009), H. virescens<br />

(Krieger et al., 2002; Krieger et al., 2004), P. xylostella, M. separata, and D. <strong>in</strong>dica<br />

(Mitsuno et al., 2008).<br />

The prote<strong>in</strong> sequence <strong>of</strong> EpOR34 was submitted to seven different transmembrane<br />

prediction programs <strong>in</strong>clud<strong>in</strong>g HMMTOP (Tusnady and Simon, 2001), TMPred<br />

(H<strong>of</strong>mann and St<strong>of</strong>fel, 1993), TMHMM (Krogh et al., 2001), DAS (Cserzo et al.,<br />

1997), SPLIT (Juretic et al., 1993), TMMOD (Kahsay et al., 2005) and TOPPRED<br />

(Claros and von Heijne, 1994). The predicted TM regions were compared and a


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 99<br />

consensus <strong>of</strong> the different doma<strong>in</strong>s was made manually. A topology map <strong>of</strong> EpOR34<br />

was then drawn <strong>in</strong> TOPO2 (http://www.sacs.ucsf.edu/TOPO-run/wtopo.pl) us<strong>in</strong>g the<br />

consensus sequence above.<br />

4.3 Results<br />

4.3.1 Microarray Analysis<br />

A microarray analysis <strong>of</strong> 1809 ESTs from a male antennal cDNA library <strong>of</strong> E.<br />

postvittana was carried out to identify male-specific genes. From the microarray<br />

screen<strong>in</strong>g, four ESTs had more than two times higher expression <strong>in</strong> male than female<br />

antennae (EST numbers 23893, 25116, 25174 and 25214). The sex-biased expression<br />

<strong>of</strong> two (ESTs 25174 and 25214) <strong>of</strong> the four ESTs were confirmed by qRT-PCR<br />

(Figure 4.1).<br />

Figure 4.1: Quantitative RT-PCR <strong>of</strong> four candidates PRs as identified from a<br />

differential microarray screen<strong>in</strong>g <strong>of</strong> EST oligos. The error bars represent standard<br />

errors <strong>of</strong> three technical replicates for each tissue type. ND = not detectable.<br />

3‟RACE-PCR was not done on ESTs 25174 and 25214 as the orig<strong>in</strong>al sequences<br />

already had the poly A + tail sequence. The 5‟RACE-PCR was only able to extend the<br />

orig<strong>in</strong>al EST sequences by 350–400 bp. Further 5‟RACE to extend the sequences


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 100<br />

further were unsuccessful. Tblastx <strong>of</strong> the sequences <strong>of</strong> the 5‟RACE-PCR products <strong>of</strong><br />

ESTs 25174 and 25214 <strong>in</strong> the NCBI BLAST server did not yield any significant hits.<br />

The orig<strong>in</strong>al EST sequences were also used to search the blastable E. postvittana 454<br />

transcripts as well as the assembled genomic scaffolds us<strong>in</strong>g the tblastx tool. For both<br />

the ESTs, the 454 transcripts hit were the same length and sequence as the orig<strong>in</strong>al<br />

sequences hence deep transcriptomics was not able to extend these sequences any<br />

further. Fragments <strong>of</strong> the orig<strong>in</strong>al ESTs were also found <strong>in</strong> the genomic scaffolds,<br />

however the scaffolds were not long enough to extend the EST sequences any further.<br />

4.3.2 Deep Transcriptomics<br />

The male E. postvittana antennal cDNA was normalised prior to deep transcriptomics<br />

<strong>in</strong> an attempt to reduce to levels <strong>of</strong> highly expressed genes and <strong>in</strong>crease the chance <strong>of</strong><br />

sequenc<strong>in</strong>g the lowly expressed ORs. The male E. postvittana antennal cDNA<br />

received from Evrogen after normalisation had an average length <strong>of</strong> 1 kb. Figure 4.2<br />

shows 500ng <strong>of</strong> the cDNA before (lane 1) and after (lane 2) normalisation. The highly<br />

fluoresc<strong>in</strong>g bands <strong>in</strong> lane 1 seemed to have been evened out after undergo<strong>in</strong>g DSN<br />

treatment suggest<strong>in</strong>g that the cDNA population <strong>in</strong> the sample had been successfully<br />

normalised.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 101<br />

Figure 4.2: E. postvittana male antennal SMART-amplified cDNA before (1) and<br />

after (2) normalisation. Approximately 500ng <strong>of</strong> cDNA were loaded on the gel. Lane<br />

M is 1kb DNA marker.<br />

4.3.2.1 454 Sequenc<strong>in</strong>g Statistics<br />

A s<strong>in</strong>gle whole plate <strong>of</strong> 454 titanium EST sequenc<strong>in</strong>g <strong>of</strong> the normalised E. postvittana<br />

male antennae cDNA yielded 1,091,459 EST sequences. This data underwent<br />

validation before any analysis was conducted. The validation process <strong>in</strong>volved<br />

filter<strong>in</strong>g the ESTs and trimm<strong>in</strong>g possible vector sequences, removal <strong>of</strong> low quality<br />

sequences as well as sequences shorter than 50 bp. After the validation process, the<br />

number <strong>of</strong> ESTs rema<strong>in</strong><strong>in</strong>g were 903,121 totall<strong>in</strong>g 188,511,228 bp. Contig assembly<br />

was done us<strong>in</strong>g Newbler s<strong>of</strong>tware under default sett<strong>in</strong>gs. This resulted <strong>in</strong> 21,627<br />

contigs and 94,579 s<strong>in</strong>gletons. The longest contig was 1994 bp and the N50 (the<br />

contig length that produces half the bases <strong>in</strong> the genome) <strong>of</strong> the comb<strong>in</strong>ed contigs and<br />

s<strong>in</strong>gletons was 279 bp. An analysis <strong>of</strong> the 454 sequence data is given <strong>in</strong> Table 4.5.<br />

The contigs and s<strong>in</strong>gletons were deposited <strong>in</strong>to a privately held Bioview Insect<br />

Database at Plant and Food Research Ltd and automatically annotated us<strong>in</strong>g Gene<br />

Ontology (GO) analysis.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 102<br />

Table 4.5: Analysis <strong>of</strong> the 454 transcriptome sequence data from the raw data<br />

obta<strong>in</strong>ed (before contig assembly) through to the processed data, assembled contigs<br />

and s<strong>in</strong>gletons.<br />

Before contig<br />

assembly<br />

No. <strong>of</strong><br />

ESTs<br />

Bases (bp) N50 Average<br />

length (bp)<br />

Median<br />

(bp)<br />

Longest<br />

(bp)<br />

903121 188511228 208.7 192 594<br />

Contigs 21627 6884597 369 318.3 311 1994<br />

S<strong>in</strong>gletons 94579 16837407 234 178.0 152 579<br />

Contigs +<br />

s<strong>in</strong>gletons<br />

116206 23722001 279 204.1 175 1994<br />

4.3.3 Mitochondrial DNA contam<strong>in</strong>ation<br />

Before perform<strong>in</strong>g whole genome sequenc<strong>in</strong>g, an attempt was made at m<strong>in</strong>imis<strong>in</strong>g the<br />

sequenc<strong>in</strong>g <strong>of</strong> mitochondrial DNA thereby <strong>in</strong>creas<strong>in</strong>g sequenc<strong>in</strong>g <strong>of</strong> nucleic DNA (as<br />

the OR genes are nuclear), the nucleic was isolated first, then gDNA extracted from it.<br />

E. postvittana DNA prepared from nuclei conta<strong>in</strong>ed less mitochondrial DNA than<br />

nuclear DNA based on semi-quantitative PCR <strong>of</strong> the mitochondrial and nuclear genes<br />

(Figures 4.3). The DNA prepared from nuclei was then used as a template for<br />

Illum<strong>in</strong>a sequenc<strong>in</strong>g. Just to be cautious that the gDNA sample has m<strong>in</strong>imal mtDNA<br />

contam<strong>in</strong>ation, one lane <strong>of</strong> Illum<strong>in</strong>a sequenc<strong>in</strong>g was <strong>in</strong>itially done on the gDNA.<br />

Bowtie, which is an ultrafast short read aligner (Langmead et al., 2009) was used for<br />

mapp<strong>in</strong>g the mitochondrial sequences <strong>in</strong> genbank and a 2,216 base pair region <strong>of</strong> the<br />

E. postvittana mitochondrial genome conta<strong>in</strong><strong>in</strong>g the cytochrome oxidase I and II<br />

genes to the 13 million paired end reads. The test parameters for bowtie allows up to<br />

only 3 mismatches. Only 92 out <strong>of</strong> the 13 million paired end reads were predicted to<br />

be mitochondrial, suggest<strong>in</strong>g that mtDNA contam<strong>in</strong>ation is m<strong>in</strong>imal. Another 7 lanes<br />

<strong>of</strong> Illum<strong>in</strong>a sequenc<strong>in</strong>g was then done <strong>in</strong> order to obta<strong>in</strong> higher coverage <strong>of</strong> the<br />

genome and an attempt at assembl<strong>in</strong>g longer scaffolds for use as a sequence set able<br />

to be searched us<strong>in</strong>g blast.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 103<br />

850bp<br />

650bp<br />

1kb+<br />

ladder<br />

Figure 4.3: Semi-quantitative PCR on nuclear (Takeout 3) and mitochondrial<br />

(cytochrome oxidase I) genes us<strong>in</strong>g total DNA extracted from purified nuclei as<br />

template. Lanes A to G show PCR performed on a 10 fold serial dilution <strong>of</strong> the DNA<br />

template from 1 <strong>in</strong> 100 to 1 <strong>in</strong> 100 million with E. postvittana Takeout 3 primers.<br />

Lanes H to N are PCR performed on the same template serial dilution with<br />

cytochrome oxidase I primers.<br />

4.3.4 Illum<strong>in</strong>a Sequenc<strong>in</strong>g<br />

The assembly <strong>of</strong> the 8 lanes <strong>of</strong> Illum<strong>in</strong>a sequenc<strong>in</strong>g resulted <strong>in</strong> 299,968,860 bases<br />

be<strong>in</strong>g assembled <strong>in</strong>to 843,963 scaffolds with the longest scaffold be<strong>in</strong>g 19,645 bp,<br />

with an N50 <strong>of</strong> 511 bp. The sequences were deposited <strong>in</strong>to a privately held genome<br />

server at Plant and Food Research Ltd for blast query<strong>in</strong>g and data m<strong>in</strong><strong>in</strong>g.<br />

4.3.5 Data M<strong>in</strong><strong>in</strong>g<br />

Nucleic DNA products Mitochondrial DNA products<br />

A B C D E F G H I J K L M N<br />

Tblastn <strong>of</strong> the E. postvittana 454 transcripts and genomic sequences with known B.<br />

mori OR sequences resulted <strong>in</strong> the identification <strong>of</strong> a further 49 putative ORs,<br />

br<strong>in</strong>g<strong>in</strong>g the total number <strong>of</strong> ORs identified from E. postvittana to 52. The ORs with<br />

shortest and longest predicted cod<strong>in</strong>g region lengths are EpOR49 (40 am<strong>in</strong>o acids)<br />

and EpOR21 (475 am<strong>in</strong>o acids) respectively. The full length <strong>of</strong> only one putative OR<br />

was identified (EpOR34). The highest level <strong>of</strong> am<strong>in</strong>o acid identity shared between E.<br />

postvittana and B. mori ORs is 88% between EpOR2 and BmOR2. Interest<strong>in</strong>gly,<br />

EpOR8 shares 71% am<strong>in</strong>o acid identity with BmOR8, a larval-specific OR <strong>in</strong> B. mori.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 104<br />

Table 4.6 summarises the properties <strong>of</strong> the E. postvittana ORs, with the predicted<br />

cod<strong>in</strong>g region length available for each OR, its closest B. mori homologue based on<br />

pairwise distance matrix calculated <strong>in</strong> ClustalX and the method(s) that identified each<br />

OR. TMPRED was used for predict<strong>in</strong>g the topology <strong>of</strong> the ORs for which more than<br />

350 am<strong>in</strong>o acids <strong>of</strong> the predicted cod<strong>in</strong>g regions were available. Of these, EpOR2, 32,<br />

34 and 47 were predicted to have <strong>in</strong>tracellular N-term<strong>in</strong>us (shaded <strong>in</strong> orange and<br />

purple on Figure 4.4). EpOR1 and EpOR6 share the most am<strong>in</strong>o acid identity with<br />

BmOR1, a PR <strong>of</strong> B. mori. EpOR1 has been demonstrated to be a receptor for plant<br />

volatiles <strong>in</strong> Chapter 2; hence EpOR6 (shaded <strong>in</strong> yellow on Figure 4.4) is predicted to<br />

be a candidate PR <strong>of</strong> E. postvittana.<br />

Table 4.6: E. postvittana putative OR repertoire to date. The ORs were identified<br />

either by Sanger sequenc<strong>in</strong>g (S), 454 transcriptome sequenc<strong>in</strong>g (T) or low coverage<br />

whole genome sequenc<strong>in</strong>g (G). The B. mori homologues <strong>of</strong> the correspond<strong>in</strong>g E.<br />

postvittana ORs are given, together with the percentage am<strong>in</strong>o acid identity shared<br />

between the homologues. This data is based on the <strong>in</strong>complete sequences <strong>of</strong> the<br />

putative E. postvittana ORs and can change as the full length sequences become<br />

available.<br />

Putative<br />

OR<br />

Length<br />

(am<strong>in</strong>o<br />

acids)<br />

Closest B. mori OR<br />

identified with the<br />

available length<br />

% am<strong>in</strong>o acid identity with<br />

B. mori ORs with the<br />

current sequence length<br />

Method<br />

EpOR1* 415 BmOR1 38 S, T, G<br />

EpOR2* 474 BmOR2 88 S, T, G<br />

EpOR3* 410 BmOR49J 65 S, T, G<br />

EpOR4 268 BmOR56 60 G<br />

EpOR5 304 BmOR68 22 G<br />

EpOR6 241 BmOR1 17 T, G<br />

EpOR7 308 BmOR4 26 T, G<br />

EpOR8 324 BmOR8 71 G<br />

EpOR9 225 BmOR4 14 G<br />

EpOR10 349 BmOR27 22 G<br />

EpOR11 113 BmOR26 19 G<br />

EpOR12 293 BmOR15 45 G<br />

EpOR13 271 BmOR13 52 T, G<br />

EpOR14 118 BmOR14 20 T, G<br />

EpOR15 319 BmOR13 20 T, G<br />

EpOR16 206 BmOR41 23 G<br />

EpOR17 243 BmOR49 60 G


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 105<br />

EpOR18 312 BmOR30 37 G<br />

EpOR19 209 BmOR28 17 G<br />

EpOR20 131 BmOR14 37 T, G<br />

EpOR21 475 BmOR49 18 T, G<br />

EpOR22 84 BmOR6 32 T, G<br />

EpOR23 206 BmOR23 33 T, G<br />

EpOR24 114 BmOR24 28 T, G<br />

EpOR25 257 BmOR18 47 G<br />

EpOR26 188 BmOR26 22 G<br />

EpOR27 329 BmOR27 57 T, G<br />

EpOR28 170 BmOR28 15 G<br />

EpOR29 185 BmOR29 57 T, G<br />

EpOR30 257 BmOR30 28 T, G<br />

EpOR31 282 BmOR16 38 G<br />

EpOR32 359 BmOR32 41 T, G<br />

EpOR33 118 BmOR68 22 T, G<br />

EpOR34 395 BmOR33 30 T, G<br />

EpOR35 324 BmOR36 49 G<br />

EpOR36 152 BmOR61 18 T, G<br />

EpOR37 219 BmOR55 43 G<br />

EpOR38 162 BmOR38 55 G<br />

EpOR39 120 BmOR39 60 T, G<br />

EpOR40 113 BmOR16 31 G<br />

EpOR41 210 BmOR29 42 T, G<br />

EpOR42 293 BmOR42 33 T, G<br />

EpOR43 175 BmOR63 22 G<br />

EpOR44 211 BmOR44 29 G<br />

EpOR45 77 BmOR15 22 G<br />

EpOR46 221 BmOR47 17 G<br />

EpOR47 350 BmOR60 54 T, G<br />

EpOR48 97 BmOR52 21 T<br />

EpOR49 40 BmOR30 30 T<br />

EpOR50 70 BmOR46 20 T<br />

EpOR51 76 BmOR26 23 T<br />

EpOR52 59 BmOR64 27 T<br />

*These E. postvittana ORs were identified by Jordan et al. (2009), while all the other<br />

ORs have been identified <strong>in</strong> this study.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 106


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 107


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 108


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 109


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 110<br />

Figure 4.4: Multiple sequence alignment <strong>in</strong> ClustalX <strong>of</strong> E. postvittana ORs identified<br />

to date. Transmembrane positions obta<strong>in</strong>ed from consensus <strong>of</strong> EpOR1, 2, 3 and 34 are<br />

shown as light orange bars above the alignment, with the roman numeral <strong>in</strong>dicat<strong>in</strong>g<br />

the TM doma<strong>in</strong> region. The predicted E. postvittana PRs are shaded <strong>in</strong> yellow, the<br />

ORs with predicted <strong>in</strong>tracellular N-term<strong>in</strong>us are shaded <strong>in</strong> purple while EpOR34<br />

which has both these features is shaded <strong>in</strong> dark orange.<br />

4.3.6 Quantitative Real-Time PCR<br />

The expression levels <strong>of</strong> 23 <strong>of</strong> the 49 new putative E. postvittana ORs (these 23<br />

putative ORs were identified from the 454 transcriptome sequenc<strong>in</strong>g; the Illum<strong>in</strong>a<br />

sequenc<strong>in</strong>g data became available towards the end <strong>of</strong> the project and were not able to<br />

be analysed by qRT-PCR due to time restriction) together with EpOR1, 2 and 3 were<br />

exam<strong>in</strong>ed <strong>in</strong> male antennae, female antennae and female body (m<strong>in</strong>us head) tissues<br />

us<strong>in</strong>g qRT-PCR (Figure 4.5). Sex-biased expression was detected <strong>in</strong> three ORs, with<br />

EpOR30, 33 and 34 be<strong>in</strong>g expressed more than 500 times more highly <strong>in</strong> male than<br />

female antennae, therefore are predicted to be the PRs <strong>of</strong> E. postvittana (these are<br />

shaded <strong>in</strong> yellow and orange on Figure 4.4). A further 12 ORs (EpOR7, 13, 23, 24,<br />

36, 39, 41, 47, 49, 50, 51 and 52) had two to ten-fold higher expression <strong>in</strong> male than<br />

female antennae while seven ORs (EpOR14, 20, 21, 27, 32, 42 and 48) had similar<br />

expression levels <strong>in</strong> male and female antennae. Expression <strong>of</strong> eight ORs (EpOR7, 13,<br />

24, 27, 33, 49, 50 and 52) was also detected <strong>in</strong> the body.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 111<br />

Figure 4.5: Expression levels <strong>of</strong> the putative E. postvittana ORs <strong>in</strong> male antennae, female antennae and body tissues. The housekeep<strong>in</strong>g genes<br />

Ef1α and α-tubul<strong>in</strong> were used for normalis<strong>in</strong>g the expression levels. The error bars represent standard errors calculated from two biological<br />

replicates. * The tissue expression <strong>of</strong> these ORs was not determ<strong>in</strong>ed.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 112<br />

A prelim<strong>in</strong>ary phylogenetic tree was constructed from multiple sequence alignment <strong>of</strong><br />

the 52 E. postvittana ORs together with known ORs from five other moth species B.<br />

mori (Wanner et al., 2007; Tanaka et al. 2009), H. virescens (Krieger et al., 2002,<br />

2004), P. xylostella, M. separata, and D. <strong>in</strong>dica (Mitsuno et al., 2008) (Figure 4.6).<br />

The sequences were obta<strong>in</strong>ed from GenBank with accession numbers given <strong>in</strong> the<br />

mentioned studies. The alignment produc<strong>in</strong>g this tree is given <strong>in</strong> Appendix D. This<br />

tree has been constructed us<strong>in</strong>g partial E. postvittana OR sequences therefore the<br />

phylogenetic <strong>in</strong>formation presented here may be <strong>in</strong>accurate and only the availability<br />

<strong>of</strong> the full length sequences <strong>of</strong> these putative ORs will give a true relationship <strong>of</strong> these<br />

ORs with those from the five other moth species. The male-biased sex pheromone<br />

receptor clade to which all the Lepidoptera PRs identified to date belong to, and the<br />

female-biased clade <strong>of</strong> B. mori are <strong>in</strong>dicated, together with the highly conserved<br />

Drosophila OR83b clade on Figure 4.6. The putative E. postvittana PRs are shaded <strong>in</strong><br />

red. Interest<strong>in</strong>gly, the three sex-biased putative PRs (EpOR30, 33 and 34) cluster<br />

together, lead<strong>in</strong>g us to deduce that a different set <strong>of</strong> ORs (as opposed to the PRs <strong>of</strong><br />

other moths form<strong>in</strong>g one clade) have been co-opted <strong>in</strong> E. postvittana for a role <strong>in</strong><br />

pheromone reception.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 113<br />

Figure 4.6: A prelim<strong>in</strong>ary phylogram result<strong>in</strong>g from multiple sequence alignment <strong>of</strong><br />

ORs and PRs from moths, together with EpORs1–52. The tree was constructed us<strong>in</strong>g<br />

Fitch us<strong>in</strong>g John-Thorton distances and rooted with the clade <strong>of</strong> EpOR2 orthologues.<br />

Bootstrap values were calculated from 1000 bootstrap replicates and are given as a<br />

percentage value on the nodes, where possible. The putative E. postvittana PRs are<br />

shaded <strong>in</strong> red.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 114<br />

4.3.7 EpOR34 – a putative PR<br />

The only putative E. postvittana OR for which a full length cod<strong>in</strong>g region has been<br />

obta<strong>in</strong>ed from the 49 new ORs identified is EpOR34. The predicted cod<strong>in</strong>g region <strong>of</strong><br />

EpOR34 (Figure 4.7) was assembled from the partial transcript <strong>of</strong> EST 1032444, a<br />

2033 bp genomic scaffold and 3‟RACE-PCR. The predicted cod<strong>in</strong>g region was 1185<br />

bp and encodes 395 am<strong>in</strong>o acids from the start methion<strong>in</strong>e to the stop codon. From the<br />

genomic sequence, two <strong>in</strong>trons were identified, as <strong>in</strong>dicated by red arrowheads <strong>in</strong><br />

Figure 4.7. The first <strong>in</strong>tron occurred after am<strong>in</strong>o acid 258 and was 87 bp while the<br />

second <strong>in</strong>tron was after am<strong>in</strong>o acid 291 and was 351 bp long. The second <strong>in</strong>tron could<br />

be even longer as the genomic scaffold encod<strong>in</strong>g EpOR34 did not have the 3‟ end <strong>of</strong><br />

the gene and the sequence after the second <strong>in</strong>tron was obta<strong>in</strong>ed from the transcriptome<br />

sequence data and the poly A+ tail obta<strong>in</strong>ed by 3‟RACE-PCR. The am<strong>in</strong>o acid<br />

sequence was also confirmed by homology to known B. mori ORs <strong>in</strong> GenBank.<br />

Seven different freely available membrane prediction programs were used for<br />

predict<strong>in</strong>g the transmembrane doma<strong>in</strong>s <strong>of</strong> EpOR34 from its prote<strong>in</strong> cod<strong>in</strong>g region.<br />

Moth ORs are predicted to have 7TM doma<strong>in</strong>s (Benton et al., 2006), however not all<br />

the prediction programs used here gave 7TMs. S<strong>in</strong>ce the exact placement <strong>of</strong> the start<br />

and f<strong>in</strong>ish <strong>of</strong> the TM doma<strong>in</strong>s and the number <strong>of</strong> TMs predicted varied between the<br />

different prediction programs, a consensus <strong>of</strong> the predicted doma<strong>in</strong> regions was<br />

developed and is shown <strong>in</strong> roman numerals on Figure 4.8. This consensus found<br />

EpOR34 to have 7TMs and TMPred predicted an <strong>in</strong>tracellular N-term<strong>in</strong>us, consistent<br />

with the recent evidence that suggest <strong>in</strong>sect ORs as hav<strong>in</strong>g <strong>in</strong>tracellular N-term<strong>in</strong>us<br />

(Benton et al., 2006; Lund<strong>in</strong> et al., 2007; Smart et al., 2008; Jordan et al., 2009).


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 115<br />

1 M E D K L V F E E V Y K I N M T C L R L N L S H 24<br />

1 ATCTTTCAGTAAATTATAATGGAAGACAAATTAGTTTTTGAAGAAGTCTATAAGATCAACATGACCTGCCTCCGATTAAACCTCTCACAC 90<br />

25 P S V P R D L K W L L T F I L M Q G L Y T A F N A I C L Y N 54<br />

91 CCTTCTGTGCCGAGGGATCTCAAATGGCTCTTGACTTTTATCTTAATGCAGGGACTGTATACCGCATTCAATGCGATCTGTCTCTACAAC 180<br />

55 L I F I N V K E N D F P N A C S N G V Y M V I Y F V V T F K 84<br />

181 TTGATATTTATCAACGTAAAAGAAAATGATTTTCCCAATGCATGCAGCAACGGTGTTTATATGGTTATTTACTTTGTAGTAACATTTAAA 270<br />

85 Y G V M V W Y Q K D I K D V I R Y Q Q E Y F D S F R E Y T V 114<br />

271 TATGGTGTGATGGTGTGGTATCAGAAGGATATAAAGGATGTTATACGATATCAGCAGGAATACTTTGATTCTTTCCGAGAATATACTGTT 360<br />

115 E E Q A V V K D Y I Q R G Q W V S K L W L R S T I V T A G M 144<br />

361 GAAGAGCAAGCTGTAGTAAAAGATTACATCCAAAGAGGGCAGTGGGTATCGAAGTTATGGCTTAGATCCACTATCGTCACTGCGGGTATG 450<br />

145 F P V K S F I D S A Y S A Y A G D F R L H S F N E N S Y G P 174<br />

451 TTTCCTGTTAAGAGTTTCATTGATAGCGCTTATTCAGCCTACGCTGGAGATTTCAGGCTACATAGCTTCAATGAGAACAGTTATGGTCCG 540<br />

175 Y I D E I K G R V D V F I L M Y A I F S V Y T T Y T A I M Y 204<br />

541 TATATTGACGAAATCAAGGGCAGGGTAGATGTATTCATTTTAATGTACGCCATATTCAGTGTGTATACTACTTACACCGCAATCATGTAT 630<br />

205 S G F A P F G P L C I L N A C A Q M D I V M M R V N H L F D 234<br />

631 TCGGGCTTTGCGCCTTTTGGGCCACTTTGTATACTGAACGCCTGTGCTCAGATGGACATAGTAATGATGAGAGTCAATCACCTTTTCGAC 720<br />

235 E G F D K E K S P K Q L Q N L V K F T Q N I Y G ▼ F V D Q I N 264<br />

721 GAAGGTTTCGACAAGGAAAAATCACCGAAGCAGTTACAAAACTTAGTAAAGTTTACACAAAATATATACGGGTTTGTTGATCAAATTAAT 810<br />

265 D I F Q V L Y E M C L K A S A I L I P I S L Y L I I E ▼ G F S 294<br />

811 GATATATTTCAAGTTTTGTACGAGATGTGTTTAAAAGCCTCTGCGATATTAATACCTATTTCACTGTATTTGATAATTGAGGGTTTCAGT 900<br />

295 E G K L Y Y D Y V M F S Y M A S L L C F V P C Y Y S D Y L K 324<br />

901 GAGGGCAAGCTGTACTATGACTACGTGATGTTTTCTTACATGGCGTCTTTACTCTGCTTTGTGCCTTGCTATTATAGCGATTATCTCAAG 990<br />

325 E K G D D L R C A I Y A S G W E K F Y D R N T R V T L R I M 354<br />

991 GAAAAGGGTGACGATCTGCGTTGCGCAATATACGCGTCGGGCTGGGAGAAATTCTACGACAGAAATACCAGAGTCACCCTCCGCATAATG 1080<br />

355 L I R A T R S L S I K T V F R A V C L E A F S D L C K E A Y 384<br />

1081 CTGATACGAGCGACGAGGTCCCTCTCCATAAAGACGGTGTTCAGGGCTGTCTGCCTGGAAGCCTTCTCTGATTTGTGTAAAGAAGCTTAC 1170<br />

385 V I F N M M Y A V L H 396<br />

1171 GTTATTTTCAATATGATGTATGCTGTTTTGCATTAAGTCAGCTTGATGATTTAAATTTTATTCGGGCACAAACGGTACAATACTCATACT 1260<br />

1261 TATAATTGCACAATTTAGTTTGGGTCACCTTGAGAGCTAAAAGCTAAAGAATCCTTTCAGCGGAATCCAACAAACTTTACTTGTACTAAT 1350<br />

1351 TATCCACACACTTTGCAAATTTTATTGTTGTCATAATTCAAGTTCTTCACATATTTTAATAAACTCTTTCACACCACAAAAAAAAAA 1437<br />

Figure 4.7: EpOR34 cDNA nucleotide sequence show<strong>in</strong>g the predicted start<br />

methion<strong>in</strong>e <strong>in</strong> green, the theoretical prote<strong>in</strong> cod<strong>in</strong>g region above the nucleotide<br />

sequence and the stop codon at the end <strong>of</strong> the cod<strong>in</strong>g region as well as upstream <strong>of</strong><br />

the translation <strong>in</strong>itiation site <strong>in</strong> purple. The predicted polyadenylation site and the poly<br />

A+ tail are shown <strong>in</strong> red.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 116<br />

TMMOD<br />

DAS<br />

TMPRED<br />

SPLIT<br />

TOPPRED<br />

HMMTOP<br />

TMHMM<br />

N L1 L2 L3 L4 L5 L6<br />

C<br />

Intracellular<br />

I II III IV V VI VII<br />

0 50 100 150 200 250 300 350<br />

400<br />

Figure 4.8: Schematic representation <strong>of</strong> the consensus <strong>of</strong> the predicted TM regions<br />

from seven TM prediction programs. The black blocks represent the predicted TM<br />

regions by the respective program on the left, and are drawn to scale, correspond<strong>in</strong>g to<br />

the predicted cod<strong>in</strong>g region <strong>of</strong> the gene, as <strong>in</strong>dicated by the ruler at the bottom.<br />

Number I to VII are the seven consensus TM doma<strong>in</strong> regions, N- and C- term<strong>in</strong>us are<br />

<strong>in</strong>dicated at the ends and the <strong>in</strong>tercellular loops <strong>in</strong>dicated by L1 to L6.<br />

The transmembrane prediction program, TMHMM2.0 predicted five TM doma<strong>in</strong>s<br />

(TM1, 2, 4, 5 and 6) for EpOR34 with high probability, TM3 with low probability<br />

while TM 7 was not predicted at all. These results were compared with the TM<br />

prediction from TMMOD program. This predicted TM 1-6 with high probability<br />

Residue<br />

while TM7 was also predicted albeit with lower probability.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 117<br />

I II III IV V VI VII<br />

Figure 4.9: Transmembrane doma<strong>in</strong> prediction <strong>of</strong> EpOR34 with TMHMM 2.0<br />

algorithms. The y-axis shows the probability <strong>of</strong> an am<strong>in</strong>o acid to be part <strong>of</strong> a TM<br />

doma<strong>in</strong>. The numbered red shaded areas <strong>in</strong>dicate predicted TM doma<strong>in</strong>s.<br />

I II III IV V VI VII<br />

Figure 4.10: Transmembrane doma<strong>in</strong> prediction <strong>of</strong> EpOR34 with TMMOD<br />

algorithms. The y-axis shows the probability <strong>of</strong> an am<strong>in</strong>o acid to be part <strong>of</strong> a TM<br />

doma<strong>in</strong>. The numbered red shaded areas <strong>in</strong>dicate predicted TM doma<strong>in</strong>s.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 118<br />

These results were then compared with TM predictions us<strong>in</strong>g five other programs (as<br />

stated <strong>in</strong> section 4.2.16). Not all <strong>of</strong> these programs predicted the classical topology <strong>of</strong><br />

7TMs for ORs. This could <strong>in</strong> part be due to the lack <strong>of</strong> structural <strong>in</strong>formation <strong>in</strong> the<br />

prote<strong>in</strong> database for closely related prote<strong>in</strong>s. Therefore, a consensus <strong>of</strong> the output<br />

from the different prediction programs was taken <strong>in</strong> order to assign TM doma<strong>in</strong>s to<br />

EpOR34. The placement <strong>of</strong> these TM doma<strong>in</strong>s correlate with the TM doma<strong>in</strong>s<br />

predicted for EpOR1, 2 and 3 <strong>in</strong> Jordan et al. (2009), and are shown <strong>in</strong> numerals on<br />

Figure 4.4.<br />

To draw a topology map <strong>of</strong> EpOR34, the TM predictions produced by the consensus<br />

(numbers I to VII on Figure 4.8) was used. A short C-term<strong>in</strong>us was predicted by both<br />

SPLIT and HMMTOP, comparable to the short C-term<strong>in</strong>us predicted for the<br />

topologies <strong>of</strong> EpOR1, EpOR2 and EpOR3, perhaps a feature conserved <strong>in</strong> <strong>in</strong>sects<br />

(Jordon, 2006). Some Drosophila ORs are also predicted to have short C-term<strong>in</strong>i<br />

(Clyne et al., 1999; Gao and Chess, 1999; Dobritsa et al., 2003; Kiely, 2008). This<br />

could just be an artefact <strong>of</strong> the prediction algorithms used and only the solved<br />

structure <strong>of</strong> such a prote<strong>in</strong> will be able to confirm this theoretical observation.<br />

N<br />

Figure 4.11: Theoretical membrane topology <strong>of</strong> EpOR34 from the consensus <strong>of</strong> the<br />

predictions <strong>of</strong> seven different TM prediction algorithms. N = N- term<strong>in</strong>us and C = C-<br />

term<strong>in</strong>us.<br />

C


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 119<br />

4.4 Discussion<br />

A total <strong>of</strong> 52 ORs have so far been identified us<strong>in</strong>g three different methods. Sanger<br />

sequenc<strong>in</strong>g identified three ORs, 454 transcriptome sequenc<strong>in</strong>g identified 29 ORs and<br />

47 ORs were identified from low coverage whole genome sequenc<strong>in</strong>g <strong>of</strong> E.<br />

postvittana. Some ORs were identified by all three sequenc<strong>in</strong>g methods (three ORs),<br />

some by two (21 ORs) and some by only one method (28 ORs), as summarised <strong>in</strong><br />

Table 4.6. Sanger sequenc<strong>in</strong>g gives longer sequence fragments hence contig assembly<br />

is easier and results <strong>in</strong> long contigs be<strong>in</strong>g assembled; however the drawback <strong>of</strong> this<br />

sequenc<strong>in</strong>g technology is the low depth <strong>of</strong> coverage, with only three ORs be<strong>in</strong>g<br />

identified from Sanger sequenc<strong>in</strong>g <strong>of</strong> E. postvittana antennal cDNA (Jordan et al.,<br />

2009). ORs are lowly expressed genes and due to their low copy number <strong>in</strong> cDNA<br />

populations are not efficiently picked up <strong>in</strong> shallow Sanger sequenc<strong>in</strong>g. 454<br />

transcriptomics was more efficient at identify<strong>in</strong>g E. postvittana ORs, as it was able to<br />

identify the three ORs identified by Sanger sequenc<strong>in</strong>g and a further 26 new ORs. The<br />

drawback <strong>of</strong> this system was the short read lengths obta<strong>in</strong>ed hence the OR sequences<br />

were <strong>in</strong> small fragments. Low coverage whole genome sequenc<strong>in</strong>g was able to<br />

identify 47 ORs. This is consistent with recovery <strong>of</strong> ORs from other moths such as B.<br />

mori where whole genome sequenc<strong>in</strong>g recovered most <strong>of</strong> the ORs. However, due to<br />

the presence <strong>of</strong> <strong>in</strong>trons <strong>in</strong> <strong>in</strong>sect ORs, prediction <strong>of</strong> cod<strong>in</strong>g regions <strong>of</strong> genes becomes a<br />

challenge.<br />

The degenerate PCR approach taken to identify new putative ORs from E. postvittana<br />

did not yield any significant results. This may be attributed to the low sequence<br />

homology between <strong>in</strong>sect ORs, as only a few ORs have been identified us<strong>in</strong>g this<br />

method thus far (Mitsuno et al., 2008; Patch et al., 2009). Regions <strong>of</strong> homology<br />

among ORs are limited and result <strong>in</strong> high levels <strong>of</strong> degeneracy <strong>in</strong> designed primers.<br />

This leads to non-specific b<strong>in</strong>d<strong>in</strong>g and amplification <strong>of</strong> undesirable genes, or genes<br />

present <strong>in</strong> high copy numbers <strong>in</strong> cells.<br />

In the Lepidoptera B. mori, 68 ORs (from both adult and larvae) have been identified<br />

so far, which is higher than the number <strong>of</strong> ORs identified <strong>in</strong> E. postvittana to date.<br />

Both these lists may be <strong>in</strong>complete and further ORs might be identified from both<br />

these moths, however, for the purpose <strong>of</strong> this discussion, the 68 ORs identified <strong>in</strong> B.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 120<br />

mori (Wanner et al., 2007; Tanaka et al., 2009) and the 52 ORs identified <strong>in</strong> Jordan et<br />

al. (2009) and <strong>in</strong> this study from E. postvittana will be looked at. From pairwise<br />

sequence alignment <strong>of</strong> E. postvittana ORs with B. mori ORs (Table 4.6), 24 one-to-<br />

one orthologs <strong>of</strong> E. postvittana and B. mori ORs were identified, with pairwise am<strong>in</strong>o<br />

acid identity rang<strong>in</strong>g from 15% to 88%. The lowest pairwise am<strong>in</strong>o acid identity is<br />

shared between EpOR9 and BmOR4 while the highest homology is shared by EpOR2<br />

and BmOR2. Eleven EpORs share more than 50% am<strong>in</strong>o acid identity with their B.<br />

mori homologues suggest<strong>in</strong>g conservation <strong>of</strong> some ORs to a certa<strong>in</strong> extend with<strong>in</strong> this<br />

highly divergent group <strong>of</strong> prote<strong>in</strong>s. Members <strong>of</strong> the highly conserved OR2 clade have<br />

been shown to be co-expressed with other ORs <strong>in</strong> nearly all ORNs and may function<br />

<strong>in</strong> transport <strong>of</strong> the ORs to the dendritic membrane (Larsson et al., 2004; Neuhaus et<br />

al., 2004; Nakagawa et al., 2005). A similar function <strong>of</strong> EpOR2 is predicted based on<br />

its high homology to other members <strong>of</strong> this clade. EpOR3 has previously been shown<br />

to be highly conserved <strong>in</strong> some moth species and share 65% am<strong>in</strong>o acid identity with<br />

its B. mori homologue, BmOR49J (Jordan et al., 2009). Based on the high levels <strong>of</strong><br />

shared am<strong>in</strong>o acid identity and functional conservation <strong>of</strong> EpOR2 and EpOR3 with<br />

their B. mori counterparts, it was hypothesised that more EpORs will share these high<br />

conservation levels with the B. mori ORs. Indeed n<strong>in</strong>e more ORs have been identified<br />

<strong>in</strong> the E. postvittana genome (namely EpOR4, 8, 13, 17, 27, 29, 38, 39 and 47) that<br />

share medium to high levels <strong>of</strong> am<strong>in</strong>o acid identity (between 52% to 71%) with their<br />

B. mori counterparts, suggest<strong>in</strong>g evolution <strong>of</strong> these ORs occurred before divergence<br />

<strong>of</strong> the Tortricidae from the Bombycidae. EpOR8 shares 71% am<strong>in</strong>o acid identity with<br />

BmOR8, which is a larval-specific OR <strong>in</strong> B. mori. Based on this high level <strong>of</strong> am<strong>in</strong>o<br />

acid identity, it can be postulated that EpOR8 is a larval-specific OR <strong>in</strong> E. postvittana.<br />

Interest<strong>in</strong>gly, EpOR3 and BmOR49J (65% shared am<strong>in</strong>o acid identity) b<strong>in</strong>d citral, an<br />

oviposition deterrent <strong>of</strong> E. postvittana (Jordan et al., 2009). If shared am<strong>in</strong>o acid<br />

identity conferred functional conservation on ORs, then EpOR4 (shares 60% am<strong>in</strong>o<br />

acid identity with BmOR56) would b<strong>in</strong>d cis-jasmone, the ligand <strong>of</strong> BmOR56 and<br />

EpOR29 (shares 57% am<strong>in</strong>o acid identity with BmOR29) would b<strong>in</strong>d l<strong>in</strong>alool, citral<br />

and l<strong>in</strong>alyl acetate, the ligands <strong>of</strong> BmOR29 (Tanaka et al., 2009). However, this is<br />

only a postulation and functional analysis <strong>of</strong> these E. postvittana ORs would confirm<br />

their ligands.<br />

Expansions can also be seen to have occurred <strong>in</strong> the E. postvittana OR repertoire, with<br />

13 B. mori ORs hav<strong>in</strong>g 28 homologues <strong>in</strong> E. postvittana. BmORs1, 4, 13, 14, 15, 16,


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 121<br />

27, 28, 29, 49 and 68 have two homologues, while BmORs 26 and 30 have three<br />

homologues <strong>in</strong> E. postvittana OR repertoire, suggest<strong>in</strong>g the E. postvittana ORs with<strong>in</strong><br />

these clades duplicated after diverg<strong>in</strong>g from their B. mori homologues.<br />

Sexually dimorphic expression <strong>of</strong> ORs have been documented <strong>in</strong> some Lepidoptera<br />

species and is one <strong>of</strong> the criteria used for identify<strong>in</strong>g species-specific sex pheromone<br />

receptors. To date sex-specific pheromone receptors have been identified from B.<br />

mori, H. virescens, H. armigera, H. assulta, M. sexta, M. separate, D. <strong>in</strong>dica and P.<br />

xylostella; some based on tissue expression analysis (qRT-PCR and <strong>in</strong> situ<br />

hybridisation experiments) and some based on their phylogenetic relatedness to PRs<br />

<strong>in</strong> other moths that b<strong>in</strong>d pheromone components <strong>in</strong> functional characterisation assays<br />

(Krieger et al., 2004; Sakurai et al., 2004; Krieger et al., 2005; Mitsuno et al., 2008;<br />

Patch et al., 2009; Große-Wilde et al., 2010; Zhang et al., 2010). Most male-specific<br />

moth PRs cluster together on the same phylogenetic l<strong>in</strong>eage based on am<strong>in</strong>o acid<br />

identity. BmOR1, BmOR3, DiOR1, MsOR1, PxOR1 and HvOR13 have been shown<br />

to be expressed highly <strong>in</strong> male moth antennae and b<strong>in</strong>d female released sex<br />

pheromone components. Two EpORs, EpOR1 and EpOR6 also belong to this PR<br />

clade, with both shar<strong>in</strong>g 17-38% am<strong>in</strong>o acid identity with BmOR1. EpOR1 is<br />

expressed at similar levels <strong>in</strong> both male and female antennae and has been shown to<br />

be the receptor for plant volatiles <strong>in</strong> Chapter 2, suggest<strong>in</strong>g that even though it shares<br />

the highest am<strong>in</strong>o acid identity with PRs from other moths, it is not the PR <strong>of</strong> E.<br />

postvittana. This is supported by other non-pheromone receptor members <strong>of</strong> this clade<br />

such as BmOR9 be<strong>in</strong>g expressed <strong>in</strong> similar levels <strong>in</strong> both male and female antennae<br />

(Krieger et al., 2002; Wanner et al., 2007) and BmOR4, BmOR5, and BmOR7 are<br />

expressed <strong>in</strong> both male and female antennae albeit at higher levels <strong>in</strong> male antennae<br />

(Wanner et al., 2007). Functional characterisation <strong>of</strong> BmOR4 and BmOR5 aga<strong>in</strong>st the<br />

sex pheromones <strong>of</strong> B. mori have not yielded any ligands for these two ORs yet<br />

(Nakagawa et al., 2005), <strong>in</strong>dicat<strong>in</strong>g that perhaps these are receptors for plant volatiles<br />

or for unidentified pheromone components, provid<strong>in</strong>g support for EpOR1 be<strong>in</strong>g an<br />

OR for plant volatiles and render<strong>in</strong>g EpOR6 as a possible PR <strong>of</strong> E. postvittana. If the<br />

E. postvittana PR is a member <strong>of</strong> the PR clade, then based on am<strong>in</strong>o acid identity,<br />

EpOR6 can be postulated to be a PR candidate <strong>in</strong> E. postvittana. However, tissue<br />

expression analysis and functional characterisation <strong>of</strong> EpOR6 is required to resolve its<br />

role <strong>in</strong> E. postvittana olfaction.


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 122<br />

Another clade <strong>of</strong> <strong>in</strong>terest is the female-specific odorant receptor clade. Members <strong>of</strong><br />

this l<strong>in</strong>eage have been identified <strong>in</strong> B. mori and M. sexta, and have been shown to<br />

have female-specific tissue expression (Wanner et al., 2007; Anderson et al., 2009;<br />

Große-Wilde et al., 2010). MsOR5 forms a l<strong>in</strong>eage with BmOR19 while BmOR45,<br />

46, 47, 48 and 50 cluster together phylogenetically. EpOR46 shares 17% and EpOR50<br />

shares 20% am<strong>in</strong>o acid identity with BmOR47 and BmOR46 respectively. The low<br />

level <strong>of</strong> homology between the E. postvittana homologues <strong>of</strong> the B. mori female<br />

biased clade suggest that either the E. postvittana OR repertoire is <strong>in</strong>complete and the<br />

female biased ORs <strong>of</strong> this moth have not been identified yet, or the E. postvittana<br />

female biased ORs are not phylogenetically related to the B. mori homologues and<br />

only tissue expression analysis <strong>of</strong> all <strong>of</strong> the E. postvittana ORs will identify the female<br />

biased ORs.<br />

The tissue expression <strong>of</strong> 26 <strong>of</strong> the E. postvittana ORs <strong>in</strong> male antennae, female<br />

antennae and body tissues revealed differ<strong>in</strong>g levels <strong>of</strong> expression both between the<br />

various ORs <strong>in</strong> the same tissue as well as <strong>of</strong> the same OR <strong>in</strong> different tissues. The<br />

expression levels <strong>of</strong> these putative ORs were assessed for male-biased expression as it<br />

was assumed that a sex pheromone receptor would be more highly expressed <strong>in</strong> the<br />

male compared with the female antennae as found <strong>in</strong> various lepidopterans (Krieger et<br />

al., 2004; Krieger et al., 2005; Mitsuno et al., 2008; Große-Wilde et al., 2010). Three<br />

<strong>of</strong> the 26 candidate ORs identified from the deep transcriptomics displayed male-<br />

biased expression. EpOR34 had 1500 times higher expression <strong>in</strong> male antennae than<br />

female antennae, and EpOR30 and EpOR33 had 600 times and 900 times higher<br />

expression, respectively (Figure 4.5). Sex biased expression <strong>of</strong> ORs is a good<br />

<strong>in</strong>dicator <strong>of</strong> their <strong>in</strong>volvement <strong>in</strong> sex specific odour recognition as supported by<br />

evidence from B. mori ORs (Sakurai et al., 2004). BmOR1 and BmOR3 have male<br />

biased expression (with upto10,000 times higher expression <strong>in</strong> male antennae than <strong>in</strong><br />

female) and have been shown <strong>in</strong> a number <strong>of</strong> <strong>in</strong> vivo studies as well as <strong>in</strong> heterologous<br />

assays to be <strong>in</strong>volved <strong>in</strong> sex-specific pheromone recognition (Sakurai et al., 2004;<br />

Nakagawa et al., 2005; Große-Wilde et al., 2006; Syed et al., 2006; Wanner et al.,<br />

2007). It is tempt<strong>in</strong>g to correlate the high expression levels <strong>of</strong> EpOR30, 33 and 34 <strong>in</strong><br />

male antennae with roles <strong>in</strong> recognis<strong>in</strong>g the three components <strong>of</strong> the E. postvittana<br />

sex pheromone, however, further functional analysis <strong>of</strong> these ORs will be crucial <strong>in</strong>


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 123<br />

identify<strong>in</strong>g the roles <strong>of</strong> these three male-biased ORs <strong>in</strong> E. postvittana olfaction. ORs<br />

with high expression levels <strong>in</strong> female antennae have been shown to be <strong>in</strong>volved <strong>in</strong><br />

recognition <strong>of</strong> odorants that elicit female-specific behaviours <strong>in</strong> moths. For example,<br />

<strong>in</strong> B. mori, BmOR19 has 830 times higher expression <strong>in</strong> female antennae and<br />

functional analysis revealed it to be the receptor for l<strong>in</strong>alool, a major constituent <strong>of</strong><br />

mulberry leaves which serve as oviposition sites for B. mori (Anderson et al., 2009).<br />

No such female-specific ORs have been identified from tissue expression analysis <strong>of</strong><br />

E. postvittana ORs as yet. This could be due to the fact that the tissue expression <strong>of</strong><br />

the putative E. postvittana ORs have only been carried out on the ORs identified from<br />

the deep transcriptomics <strong>of</strong> male antennal cDNA. This sample is rich <strong>in</strong> RNA<br />

expressed highly <strong>in</strong> male antennae compared with female antennae hence the<br />

identification <strong>of</strong> any female-specific ORs <strong>in</strong> this sample will be m<strong>in</strong>imal. The female-<br />

specific ORs would be more likely identified from low coverage whole genome<br />

sequences. Tissue expression analysis <strong>of</strong> the 23 putative ORs identified exclusively by<br />

the low coverage whole genome sequenc<strong>in</strong>g should reveal the female-specific ORs <strong>of</strong><br />

E. postvittana, if any. Twelve ORs showed expression <strong>in</strong> both male and female<br />

antennae, although the expression level <strong>in</strong> the male antennae was higher than <strong>in</strong><br />

female. These male biased ORs might have a role <strong>in</strong> detect<strong>in</strong>g odorants that control<br />

em<strong>in</strong>ent behaviours <strong>in</strong> male moths. Female moths have been shown to be able to<br />

detect female released sex pheromones hence the expression <strong>of</strong> male biased ORs <strong>in</strong><br />

females could be <strong>in</strong>volved <strong>in</strong> the detection <strong>of</strong> sex pheromones (Widmayer et al.,<br />

2009). Eleven <strong>of</strong> the 26 putative E. postvittana ORs have similar levels <strong>of</strong> expression<br />

<strong>in</strong> both male and female antennae, <strong>in</strong>dicat<strong>in</strong>g their role <strong>in</strong> detect<strong>in</strong>g odorants such as<br />

plant volatiles to locate host plants. Eight ORs (EpOR7, 13, 24, 27, 33, 49, 50 and 52)<br />

were also detected <strong>in</strong> the body. Studies <strong>in</strong> H. virescens has revealed the expression <strong>of</strong><br />

ORs (HvOR2, 6 and 13) <strong>in</strong> the abdomen <strong>of</strong> this moth. One hypothesis is that the<br />

female moth may be detect<strong>in</strong>g its own sex pheromone <strong>in</strong> a negative feedback<br />

mechanism to control its release <strong>in</strong>to the environment. ORs such as EpOR27, 39 and<br />

48–52 have upto 100-fold lower expression levels as compared with EpOR1, 2, 3, 13,<br />

33, 34, 36 and 42 <strong>in</strong> the moth tissues, probably due to their expression restricted to a<br />

few sensilla, or they are more highly expressed at other developmental life stages<br />

(Tanaka et al., 2009). This varied range <strong>of</strong> expression levels <strong>of</strong> different ORs are also<br />

seen <strong>in</strong> B. mori (Wanner et al., 2007).


Identification <strong>of</strong> putative odorant receptors from Epiphyas postvittana 124<br />

Of the four potential candidate PRs identified from E. postvittana based on am<strong>in</strong>o<br />

acid identity and sex-biased expression, the full length cDNA was obta<strong>in</strong>ed for only<br />

EpOR34. Seven TM doma<strong>in</strong>s are predicted for EpOR34 by a consensus us<strong>in</strong>g seven<br />

different prediction programs, comparable to the predicted TM doma<strong>in</strong>s obta<strong>in</strong>ed for<br />

EpOR1, 2 and 3; and keep<strong>in</strong>g with the notion <strong>of</strong> seven TM doma<strong>in</strong>s for moth ORs. An<br />

alignment <strong>of</strong> the predicted TM doma<strong>in</strong>s <strong>of</strong> these four E. postvittana ORs show that the<br />

TM regions are conserved across the four ORs. Furthermore the general locations <strong>of</strong><br />

the predicted TMs <strong>of</strong> the E. postvittana ORs align with the predicted TM doma<strong>in</strong><br />

regions <strong>of</strong> the H. virescens ORs (Krieger et al., 2002) and a consensus <strong>of</strong> the B. mori<br />

homologues. This <strong>in</strong>dicates that the relative positions <strong>of</strong> the TM doma<strong>in</strong>s <strong>in</strong> ORs may<br />

be conserved across the Lepidoptera. A short C-term<strong>in</strong>us was predicted by both<br />

SPLIT and HMMTOP, comparable to the short C-term<strong>in</strong>us predicted <strong>in</strong> topologies <strong>of</strong><br />

EpOR1, EpOR2 and EpOR3, perhaps a feature conserved <strong>in</strong> <strong>in</strong>sects (Jordon, 2006).<br />

Some Drosophila ORs have also been shown to have short C-term<strong>in</strong>i (Clyne et al.,<br />

1999; Gao and Chess, 1999; Dobritsa et al., 2003; Kiely, 2008). This could just be an<br />

artefact <strong>of</strong> the prediction algorithms used and only the solved structure <strong>of</strong> such a<br />

prote<strong>in</strong> will be able to confirm this theoretical observation.<br />

F<strong>in</strong>ally, the stage has been set for identification and characterisation <strong>of</strong> the PRs <strong>of</strong> E.<br />

postvittana, with four candidates PRs be<strong>in</strong>g identified. Further work <strong>in</strong>to the tissue<br />

localisation through <strong>in</strong> situ hybridisation experiments and functional characterisation<br />

either us<strong>in</strong>g <strong>in</strong> vivo or <strong>in</strong> vitro characterisation assays will confirm the roles <strong>of</strong> these<br />

receptors <strong>in</strong> E. postvittana olfaction.


5.1 Introduction<br />

5<br />

Conclud<strong>in</strong>g Discussion<br />

Moths <strong>of</strong> the order Lepidoptera have implications as pests <strong>of</strong> the agricultural and<br />

horticultural <strong>in</strong>dustries <strong>of</strong> countries the world over. Feed<strong>in</strong>g by larvae on and<br />

damag<strong>in</strong>g leaves and fruits <strong>of</strong> important agricultural crops have vast economic impact.<br />

These pests have thus far been dealt with us<strong>in</strong>g chemical sprays, biological controls<br />

and mat<strong>in</strong>g disruption, however, the never end<strong>in</strong>g resistance to these agents and<br />

impact <strong>of</strong> <strong>in</strong>secticides on human health call for the development <strong>of</strong> improved,<br />

susta<strong>in</strong>able control measures. Studies on <strong>in</strong>sect reproduction and host selection have<br />

shed light on the <strong>in</strong>volvement <strong>of</strong> the olfactory system <strong>in</strong> the survival and reproduction<br />

<strong>of</strong> <strong>in</strong>sects. For example, mate attraction studies have shown pheromones can be used<br />

for pest control (Suckl<strong>in</strong>g and Karg, 1999). <strong>Olfaction</strong> has thus become the forefront <strong>in</strong><br />

the study <strong>of</strong> <strong>in</strong>sect pests for development <strong>of</strong> new control strategies. The mechanisms<br />

<strong>in</strong>volved <strong>in</strong> odorant reception at the molecular level are still not fully understood,<br />

however, <strong>in</strong>dividual prote<strong>in</strong>s <strong>in</strong>volved <strong>in</strong> this system are beg<strong>in</strong>n<strong>in</strong>g to be identified<br />

and their specific roles deduced. A better understand<strong>in</strong>g <strong>of</strong> the function<strong>in</strong>g <strong>of</strong> the<br />

olfactory system as a whole will lead to susta<strong>in</strong>ed pest control. The light brown apple<br />

moth is a very important agricultural pest <strong>of</strong> the New Zealand economy as a zero<br />

tolerance <strong>of</strong> egg, larvae, pupae and adult moth is required for all exports. With an<br />

agricultural based economy, appropriate control measures have to be <strong>in</strong> place to meet<br />

this str<strong>in</strong>gent control measure.


Conclud<strong>in</strong>g Discussion 126<br />

The aims <strong>of</strong> this thesis were to decode the b<strong>in</strong>d<strong>in</strong>g partners <strong>of</strong> EpOR1, a member <strong>of</strong> a<br />

group <strong>of</strong> receptors shown to be essential <strong>in</strong> odorant perception <strong>in</strong> <strong>in</strong>sects; to show the<br />

role, if any, <strong>of</strong> E. postvittana GOBP2 <strong>in</strong> odorant b<strong>in</strong>d<strong>in</strong>g by EpOR1 <strong>in</strong> a heterologous<br />

expression system; and to expand the known OR repertoire <strong>of</strong> E. postvittana and<br />

identify potential candidates for the PR(s) amongst these.<br />

5.2 Summary <strong>of</strong> results and discussion<br />

5.2.1 EpOR1 characterisation<br />

EpOR1 was previously identified from a male antennal EST library <strong>of</strong> E. postvittana<br />

by Jordan et al. (2009). EpOR1 forms a clade with ORs from other moths, <strong>in</strong>clud<strong>in</strong>g<br />

PRs from B. mori, H. virescens, P. xylostella, D. <strong>in</strong>dica and M. separate (Krieger et<br />

al., 2002; Krieger et al., 2004; Krieger et al., 2005; Nakagawa et al., 2005; Mitsuno et<br />

al., 2008). However qRT-PCR analysis did not show male-biased expression for<br />

EpOR1 so a functional analysis <strong>of</strong> EpOR1 was conducted to identify its b<strong>in</strong>d<strong>in</strong>g<br />

repertoire. Transient expression <strong>of</strong> EpOR1 <strong>in</strong> <strong>in</strong>sect Sf9 cells revealed that EpOR1<br />

b<strong>in</strong>ds a range <strong>of</strong> plant volatiles <strong>in</strong> vitro (Table 2.1). Some <strong>of</strong> these volatiles have<br />

implications <strong>in</strong> plant based <strong>in</strong>sect repellents (geraniol), act as oviposition deterrents<br />

(citral), are released by plants under herbivore attack (methyl salicylate) and are<br />

natural constituents <strong>of</strong> plant aroma (geranyl acetate). This characterisation <strong>of</strong> EpOR1<br />

reveals it to be a receptor for general odorants and shows that at least some moth ORs,<br />

like the Drosophila ORs are broadly tuned, recognis<strong>in</strong>g a range <strong>of</strong> compounds. The<br />

decod<strong>in</strong>g <strong>of</strong> b<strong>in</strong>d<strong>in</strong>g partners <strong>of</strong> ORs has implications <strong>in</strong> deal<strong>in</strong>g with <strong>in</strong>sect pests by<br />

target<strong>in</strong>g their chemosensory systems. This will aid to the development <strong>of</strong> susta<strong>in</strong>ed<br />

and environment friendly <strong>in</strong>secticides.<br />

5.2.2 EpGOBP2 reconstitution <strong>of</strong> the Sf9 cell assay<br />

E. postvittana GOBP2, identified by Newcomb et al. (2002) was successfully used for<br />

reconstitut<strong>in</strong>g the Sf9 cell assay system for functional characterisation <strong>of</strong> EpOR1.<br />

Recomb<strong>in</strong>ant EpGOBP2 was able to b<strong>in</strong>d seven <strong>of</strong> the ten compounds tested <strong>in</strong> a


Conclud<strong>in</strong>g Discussion 127<br />

VOBA sett<strong>in</strong>g (Figure 3.5). These results <strong>in</strong>dicate that EpGOBP2 is broadly tuned<br />

recognis<strong>in</strong>g a range <strong>of</strong> compounds. EpGOBP2 was further tested for its ability to<br />

solubilise two ligands <strong>of</strong> EpOR1, methyl salicylate and geranyl acetate <strong>in</strong> the Sf9 cell<br />

assay. These experiments showed that EpGOBP2 could be used to replace DMSO as a<br />

solubilisation agent for both methyl salicylate and geranyl acetate (Figures 3.7 and<br />

3.8). EpGOBP2 also enhanced the sensitivity with which EpOR1 recognise these<br />

compounds <strong>in</strong> vitro as determ<strong>in</strong>ed by compar<strong>in</strong>g EC50 values (Table 3.1). The<br />

<strong>in</strong>creased sensitivity observed for EpOR1 <strong>in</strong> the presence <strong>of</strong> EpGOBP2 could be a<br />

comb<strong>in</strong>ation <strong>of</strong> a role as a solubilis<strong>in</strong>g agent and/or as an activated ligand and suggest<br />

that species-specific OBPs can be used to reconstitute the Sf9 cell assay system for<br />

improv<strong>in</strong>g the sensitivity <strong>of</strong> the system. This is the first documented role <strong>of</strong> a<br />

lepidopteran GOBP <strong>in</strong> odorant b<strong>in</strong>d<strong>in</strong>g <strong>in</strong> an <strong>in</strong> vitro OR characterisation assay.<br />

5.2.3 Identification <strong>of</strong> putative ORs from E. postvittana<br />

The <strong>in</strong>itial development <strong>of</strong> an antennal EST library at Plant and Food Research Ltd<br />

enabled the identification <strong>of</strong> three ORs from E. postvittana (Jordan et al., 2009). A<br />

further 49 ORs have been identified as part <strong>of</strong> this project us<strong>in</strong>g 454 transcriptome<br />

and Illum<strong>in</strong>a genome sequenc<strong>in</strong>g (Table 4.6). These ORs share some common<br />

features with B. mori ORs identified so far. Twenty-four E. postvittana ORs have one-<br />

to-one orthologs with B. mori ORs. Expansions seem to have occurred <strong>in</strong> 13 <strong>of</strong> the B.<br />

mori ORs, with E. postvittana hav<strong>in</strong>g between 2 to 4 ORs for each <strong>of</strong> the 13 B. mori<br />

ORs. EpOR1 and EpOR6 fall <strong>in</strong>to the moth PR clade. However, EpOR1 has been<br />

shown to be the receptor for plant volatiles, with similar expression levels <strong>in</strong> male and<br />

female antennae. For EpOR6 tissue expression analysis by qRT-PCR has yet to be<br />

performed. The other possibility is that some <strong>of</strong> the phylogenetics is not correct as not<br />

all ORs are full length and members <strong>of</strong> the PR clade may be placed elsewhere<br />

currently. Only the full lengths <strong>of</strong> all the identified E. postvittana ORs will determ<strong>in</strong>e<br />

their true phylogeny. It could also be that the PRs are miss<strong>in</strong>g from the list <strong>of</strong><br />

identified ORs, as the E. postvittana OR repertoire is approximately twenty ORs short<br />

compared to B. mori OR repertoire, thereby <strong>in</strong>dicat<strong>in</strong>g that not all E. postvittana ORs<br />

have been identified. S<strong>in</strong>ce the PRs from six other moth species fall <strong>in</strong>to the moth PR<br />

clade, we assume the E. postvittana PR(s) to be part <strong>of</strong> this clade also. On the other<br />

hand, not all members <strong>of</strong> this clade are PRs (for example BmOR4, 5, 7 and 9, as


Conclud<strong>in</strong>g Discussion 128<br />

discussed <strong>in</strong> section 4.4.1) <strong>in</strong>dicat<strong>in</strong>g that phylogenetic relatedness does not<br />

necessitate functional conservation. The expression levels <strong>of</strong> 26 <strong>of</strong> the ORs were<br />

determ<strong>in</strong>ed <strong>in</strong> male and female antennae by qRT-PCR as shown <strong>in</strong> Figure 4.5.<br />

EpOR30, 33 and 34 show high expression levels <strong>in</strong> male than female antennae (more<br />

than a ten-fold difference), mak<strong>in</strong>g these good candidates for be<strong>in</strong>g PRs. These ORs,<br />

however, are not members <strong>of</strong> the PR clade hence it could be that the Tortricidae and<br />

Bombycidae ORs diverged early <strong>in</strong> lepidopteran evolution. If <strong>in</strong>deed these three ORs<br />

are the E. postvittana PRs, then it could be assumed that different ORs have been co-<br />

opted <strong>in</strong>to a role <strong>in</strong> pheromone reception <strong>in</strong> E. postvittana. Twelve ORs were<br />

expressed <strong>in</strong> both male and female though the level <strong>of</strong> expression was higher <strong>in</strong> the<br />

male antennae, while 11 ORs were expressed at approximately similar levels <strong>in</strong> male<br />

and female antennae. No ORs with female antennae specific expression were<br />

detected. This could be because the ORs that were analysed for tissue expression were<br />

all from male antennae specific ESTs, while the ORs isolated from the genome<br />

scann<strong>in</strong>g were not tested. Analysis <strong>of</strong> these ORs could possibly reveal the female-<br />

specific ORs. From a comb<strong>in</strong>ation <strong>of</strong> the transcriptome, light genomics and 3‟RACE<br />

PCR, the full length cod<strong>in</strong>g region <strong>of</strong> EpOR34, one <strong>of</strong> the male antennal specific ORs<br />

was identified. Phylogenetic analysis showed that it shares 25 to 30% am<strong>in</strong>o acid<br />

identity with BmOR18, 30, 33 and 34. Tissue expression analysis <strong>of</strong> the B. mori<br />

homologues <strong>of</strong> EpOR34 show that BmOR18 is expressed at similar levels <strong>in</strong> male and<br />

female antennae, BmOR30 is a female-specific receptor, and BmOR33 and 34 are<br />

expressed <strong>in</strong> both male and female antennae albeit at higher levels <strong>in</strong> male antennae<br />

(Wanner et al., 2007). However, due to the limited structural knowledge available for<br />

<strong>in</strong>sect ORs and their low levels <strong>of</strong> homology, OR function based solely on homology<br />

cannot be predicted.<br />

5.3 Current hypothesis and future directions<br />

The deorphan<strong>in</strong>g <strong>of</strong> EpOR1 revealed it to be a receptor for plant volatiles. It will be<br />

<strong>in</strong>terest<strong>in</strong>g to see if and how genetic modification <strong>of</strong> EpOR1 affects recognition <strong>of</strong> its<br />

ligands by E. postvittana. EpOR1 and EpOR3 have been shown to share some ligands<br />

(Jordan et al., 2009), suggest<strong>in</strong>g that <strong>in</strong> vitro one compound is recognised by more<br />

than one OR and vice versa, albeit with different levels <strong>of</strong> sensitivities. These ORs are


Conclud<strong>in</strong>g Discussion 129<br />

redundant for some <strong>of</strong> their ligands, a phenomenon observed <strong>in</strong> Drosophila also<br />

(Hallem and Carlson, 2006). Perhaps the modification <strong>of</strong> EpOR1 will have no<br />

significant effect on odorant recognition by the moth. Such an observation would<br />

suggest a „backup‟ mechanism employed by moths; a survival advantage <strong>of</strong> hav<strong>in</strong>g<br />

more than one prote<strong>in</strong> recognis<strong>in</strong>g a particular compound. Or it could happen that the<br />

modification <strong>of</strong> one receptor affects the recognition <strong>of</strong> odorants by another receptor <strong>in</strong><br />

vivo. Either way, such an experiment will contribute towards answer<strong>in</strong>g questions<br />

about olfactory mechanisms <strong>in</strong> <strong>in</strong>sects. Gene silenc<strong>in</strong>g us<strong>in</strong>g RNA <strong>in</strong>terference can be<br />

used for study<strong>in</strong>g host localisation <strong>of</strong> E. postivittana, any impact will help <strong>in</strong><br />

develop<strong>in</strong>g novel pest control strategies for this moth. EpOR1 could also be studied <strong>in</strong><br />

the empty-neuron system and the results obta<strong>in</strong>ed from the <strong>in</strong> vitro Sf9 cell assays<br />

could be compared with the <strong>in</strong> vivo results to study if/how the ligand b<strong>in</strong>d<strong>in</strong>g range <strong>of</strong><br />

EpOR1 is affected.<br />

A number <strong>of</strong> studies have shown the <strong>in</strong>volvement <strong>of</strong> moth PBPs <strong>in</strong> pheromone<br />

b<strong>in</strong>d<strong>in</strong>g, however little data for <strong>in</strong>sect GOBP role <strong>in</strong> odorant recognition exists. This<br />

study has demonstrated a role <strong>of</strong> EpGOBP2 <strong>in</strong> odorant solubilisation, and is the first<br />

documented study show<strong>in</strong>g the <strong>in</strong>volvement <strong>of</strong> a moth GOBP <strong>in</strong> odorant recognition<br />

by ORs <strong>in</strong> a heterologous expression system. The results are also <strong>in</strong>dicative <strong>of</strong> a role<br />

<strong>of</strong> EpGOBP2 <strong>in</strong> odorant transport for example, a 100-fold decrease is observed <strong>in</strong> the<br />

concentration <strong>of</strong> geranyl acetate required to activate EpOR1 <strong>in</strong> the Sf9 cell assay<br />

(Figures 3.8B and C) <strong>in</strong> the presence <strong>of</strong> EpGOBP2. These results suggest that s<strong>in</strong>ce<br />

OBPs have a role <strong>in</strong> OR-ligand b<strong>in</strong>d<strong>in</strong>g, then other prote<strong>in</strong>s identified <strong>in</strong> the olfactory<br />

system could play significant roles also. Odorant degrad<strong>in</strong>g enzymes could be<br />

<strong>in</strong>troduced <strong>in</strong>to the cell assay system <strong>in</strong> the presence and absence <strong>of</strong> EpGOBP2. If<br />

EpGOBP2 has a role <strong>in</strong> protect<strong>in</strong>g ligands from degradation, then the <strong>in</strong>troduction <strong>of</strong><br />

ODE <strong>in</strong> the presence <strong>of</strong> EpGOBP2 will not have any effect on EpOR1 activation by<br />

its ligands. The <strong>in</strong>teraction <strong>of</strong> EpGOBP2/ligand complex with OR can also be studied.<br />

Just like the reconstitution <strong>of</strong> the Sf9 cell assay system for EpOR1 with EpGOBP2,<br />

the effect <strong>of</strong> EpGOBP2 on ligand b<strong>in</strong>d<strong>in</strong>g by EpOR3 could also be studied. This will<br />

reveal if and how the EpGOBP2/ligand complex <strong>in</strong>teracts with the different ORs. The<br />

<strong>in</strong>volvement <strong>of</strong> GOBPs <strong>in</strong> deactivation <strong>of</strong> ligands can also be <strong>in</strong>vestigated by study<strong>in</strong>g<br />

effects <strong>of</strong> pH changes on GOBP structure. Future studies can look at decod<strong>in</strong>g the<br />

roles played by PDEs and SNMPs <strong>in</strong> <strong>in</strong> vitro assay systems for ORs. Some cell


Conclud<strong>in</strong>g Discussion 130<br />

components or test odorants may be toxic to the OR be<strong>in</strong>g tested and an<br />

understand<strong>in</strong>g <strong>of</strong> the roles <strong>of</strong> other olfactory players <strong>in</strong> these <strong>in</strong> vitro systems may<br />

help overcome these limitations.<br />

Functional studies <strong>of</strong> EpOR6, 30, 33 and 34 will enable the determ<strong>in</strong>ation <strong>of</strong> their<br />

roles <strong>in</strong> E. postvittana olfaction. The phylogenetic relatedness <strong>of</strong> EpOR6 to the<br />

pheromone receptors from other moths confers it as a PR candidate, however, tissue<br />

expression analysis and functional characterisation <strong>of</strong> this candidate OR will deduce<br />

its role <strong>in</strong> E. postvittana olfaction. The relatively high expression <strong>of</strong> EpOR30, 33 and<br />

34 <strong>in</strong> male moth antennae, as shown by qRT-PCR is suggestive <strong>of</strong> a plausible role <strong>in</strong><br />

pheromone recognition <strong>of</strong> these ORs. In situ hybridisation <strong>of</strong> these candidates can be<br />

used to test if they are likely PRs. ORs that are expressed at the base <strong>of</strong> pheromone<br />

sensitive trichoid sensilla <strong>of</strong> male moths will likely be PRs. Functional analysis by<br />

clon<strong>in</strong>g, expression and characterisation <strong>in</strong> an assay system aga<strong>in</strong>st E. postvittana<br />

pheromone components will confirm their b<strong>in</strong>d<strong>in</strong>g partners. The functional<br />

characterisation <strong>of</strong> PRs from moths such as B. mori and H. virescens has <strong>in</strong>dicated the<br />

presence <strong>of</strong> multiple PRs with<strong>in</strong> each species which are tuned to recognis<strong>in</strong>g one or<br />

more <strong>of</strong> the sex pheromone components. It is tempt<strong>in</strong>g to postulate the existence <strong>of</strong><br />

multiple PRs <strong>in</strong> E. postvittana also and one or more <strong>of</strong> the four candidates could be<br />

the PRs <strong>of</strong> E. postvittana, however, further functional analysis will have to be done to<br />

test the roles <strong>of</strong> these ORs.<br />

If any <strong>of</strong> these four candidates are found to be the PR for E. postvittana, then further<br />

strategies can be developed <strong>in</strong> deal<strong>in</strong>g with this agricultural pest. Mimics that can act<br />

as irreversible block<strong>in</strong>g agents for the b<strong>in</strong>d<strong>in</strong>g site <strong>of</strong> the PR could be developed. A<br />

mimic that will not be cleared from the receptor after signal attenuation would be<br />

ideal (native pheromones are cleared from the receptor hence large amounts are<br />

needed <strong>in</strong> mat<strong>in</strong>g disruption studies, render<strong>in</strong>g this pest control method expensive). Its<br />

release <strong>in</strong>to E. postvittana <strong>in</strong>fested areas could see a reduction <strong>in</strong> the number <strong>of</strong><br />

successful mat<strong>in</strong>gs as the mimic will flood the PR and confuse the male moth.<br />

Another strategy would be to genetically modify the PR itself so that it no longer<br />

recognises the pheromone components. This can be done by completely shutt<strong>in</strong>g <strong>of</strong>f<br />

the PR by RNAi silenc<strong>in</strong>g. Small <strong>in</strong>terfer<strong>in</strong>g RNA or double-stranded RNA (dsRNA)<br />

could be sprayed on host plants, or genetically modified plants express<strong>in</strong>g dsRNA


Conclud<strong>in</strong>g Discussion 131<br />

aga<strong>in</strong>st the receptor could be <strong>in</strong>troduced (perta<strong>in</strong><strong>in</strong>g ethical approvals). Larvae feed<strong>in</strong>g<br />

on these plants will have their PRs turned <strong>of</strong>f or modified, hence will be unable to<br />

recognise sex pheromones and the number <strong>of</strong> successful mat<strong>in</strong>gs will decrease. Turner<br />

et al. (2006) fed dsRNA to larvae and demonstrated the effective knockdown <strong>of</strong> genes<br />

expressed <strong>in</strong> the adult antennae (Turner et al., 2006). This method will however<br />

require the GM plants, siRNA and dsRNA sprays to undergo various tests to render<br />

them safe for the environment and humans.<br />

Dur<strong>in</strong>g the course <strong>of</strong> this project, 49 putative ORs were identified, however, the full<br />

length sequences <strong>of</strong> 48 <strong>of</strong> these still need to be completed. Towards the end <strong>of</strong> this<br />

study, six more lanes <strong>of</strong> Solexa genome sequences was obta<strong>in</strong>ed. However the data<br />

was not available on time to <strong>in</strong>clude <strong>in</strong> this thesis. This new data comb<strong>in</strong>ed with the<br />

previous genomic sequence dataset will hopefully give longer scaffolds with longer, if<br />

not full length cod<strong>in</strong>g regions <strong>of</strong> the putative ORs. Further novel ORs could also be<br />

identified from the new data, thereby <strong>in</strong>creas<strong>in</strong>g the receptor repertoire <strong>of</strong> E.<br />

postvittana. Functional characterisation <strong>of</strong> these full length ORs can be done to<br />

identify male and female specific ORs and their b<strong>in</strong>d<strong>in</strong>g components. A detailed<br />

functional characterisation <strong>of</strong> these ORs will facilitate the decod<strong>in</strong>g <strong>of</strong> all these ORs,<br />

and an odorant–OR map can be built for E. postvittana. Such a study will provide data<br />

that can be used as a basis to create a computer model to predict b<strong>in</strong>d<strong>in</strong>g partners for<br />

ORs <strong>in</strong> related moths. A computer model that can for example narrow down on „x‟<br />

number <strong>of</strong> candidate odorants for test<strong>in</strong>g an OR with can be used as a first screen<strong>in</strong>g<br />

towards characteris<strong>in</strong>g novel ORs. This will save the researcher physically test<strong>in</strong>g tens<br />

to hundreds <strong>of</strong> compounds per receptor to identify its b<strong>in</strong>d<strong>in</strong>g components.<br />

The only lepidopteran for which a draft genome sequence is publically available to<br />

date is the silkmoth B. mori. This is a domesticated moth which is dependent on<br />

humans for survival and reproduction. Currently it is used as the model for<br />

Lepidoptera studies, <strong>in</strong>clud<strong>in</strong>g tortricid moths however, the availability <strong>of</strong> genome<br />

data for an <strong>in</strong>sect pest will be <strong>in</strong>valuable to Tortricidae genomic biology, especially <strong>in</strong><br />

understand<strong>in</strong>g the underly<strong>in</strong>g molecular mechanisms <strong>of</strong> <strong>in</strong>sect pests and combat<strong>in</strong>g<br />

the disruptive members <strong>of</strong> this family. Keep<strong>in</strong>g this <strong>in</strong> m<strong>in</strong>d, the genome sequenc<strong>in</strong>g<br />

<strong>of</strong> E. postvittana has been <strong>in</strong>itiated. E. postvittana is perhaps one <strong>of</strong> the most<br />

disruptive members <strong>of</strong> this family, <strong>in</strong> that it is highly polyphagous and can survive <strong>in</strong>


Conclud<strong>in</strong>g Discussion 132<br />

a wide range <strong>of</strong> temperatures. The availability <strong>of</strong> a draft genome <strong>of</strong> E. postvittana will<br />

be the first genome for a pest member <strong>of</strong> this family and will become the model<br />

organism for tortricid pests. The availability <strong>of</strong> the genome sequence <strong>of</strong> E. postvittana<br />

should enable functional genomic studies <strong>of</strong> the moth; provide a wealth <strong>of</strong><br />

<strong>in</strong>formation on genes <strong>in</strong>volved <strong>in</strong> biological processes and <strong>in</strong> the regulation <strong>of</strong> gene<br />

expression as well as facilitate the study <strong>of</strong> the complexity <strong>of</strong> gene families. Studies <strong>of</strong><br />

host-plant specialisation for example, E. postvittana is polyphagous so genes <strong>in</strong>volved<br />

<strong>in</strong> its survival on such a wide host range, as well as prote<strong>in</strong>s <strong>in</strong>volved <strong>in</strong> food<br />

digestion <strong>in</strong> the gut <strong>of</strong> the larvae can be identified. Other target processes <strong>in</strong>clude the<br />

regulation <strong>of</strong> genes <strong>in</strong>volved <strong>in</strong> the survival <strong>of</strong> E. postvittana, such as for juvenile<br />

hormone, which have implications <strong>in</strong> physiological processes <strong>of</strong> moths can also be<br />

studied <strong>in</strong> detail. Other genes associated with the olfactory system and their regulation<br />

mechanisms could also be studied <strong>in</strong> detail. One <strong>of</strong> the major drawbacks <strong>of</strong> us<strong>in</strong>g<br />

chemical <strong>in</strong>secticides is development <strong>of</strong> resistance by the <strong>in</strong>sects. These resistance<br />

mechanisms can also be studied and better <strong>in</strong>secticide targets developed. With the<br />

availability <strong>of</strong> genomic data for more moths, the evolution <strong>of</strong> ORs can be studied. An<br />

improved set <strong>of</strong> ORs from lepidopterans can perhaps be used <strong>in</strong> phylogenetic studies<br />

to test for associations, if any, between phylogenetic relatedness and odorant b<strong>in</strong>d<strong>in</strong>g.<br />

The problem with any ab <strong>in</strong>itio whole genome assembly is the feasibility <strong>of</strong> the<br />

assembly process. E. postvittana does not have any closely related genome sequence<br />

available for mapp<strong>in</strong>g its genome to. The low cost <strong>of</strong> sequenc<strong>in</strong>g with high throughput<br />

sequenc<strong>in</strong>g platforms has the drawback <strong>of</strong> produc<strong>in</strong>g significantly short reads that are<br />

impossible to assemble ab <strong>in</strong>itio. Another problem fac<strong>in</strong>g E. postvittana genome<br />

assembly is that its genome size is unknown. B. mori genome has been shown to have<br />

large repeats and it is highly likely to be the case for E. postvittana. Assembl<strong>in</strong>g such<br />

a genome will be highly challeng<strong>in</strong>g.<br />

Fortunately for the genome sequenc<strong>in</strong>g project, a bacterial artificial chromosome<br />

(BAC) library <strong>of</strong> E. postvittana has already been constructed. BAC-end sequenc<strong>in</strong>g<br />

will enable local, smaller assemblies <strong>of</strong> <strong>in</strong>dividual BACs which can be used further to<br />

assemble longer scaffolds, thereby facilitat<strong>in</strong>g the assembly <strong>of</strong> the draft genome.<br />

Several BACs that potentially have genes for several <strong>of</strong> the putative ORs have already<br />

been identified. These BACs can be used as the start<strong>in</strong>g po<strong>in</strong>t for BAC shotgun


Conclud<strong>in</strong>g Discussion 133<br />

sequenc<strong>in</strong>g and assembly. The genome sequence dataset that we currently have has<br />

lots <strong>of</strong> gaps, as shown by the high number <strong>of</strong> scaffolds, approximately 843000<br />

scaffolds. This coverage is be<strong>in</strong>g improved by further sequenc<strong>in</strong>g <strong>of</strong> the genome.<br />

Another complementary approach to BAC sequenc<strong>in</strong>g is mate pair end sequenc<strong>in</strong>g<br />

(Illum<strong>in</strong>a, 2010), which will aga<strong>in</strong> give the ends <strong>of</strong> 5kb long fragments that can be<br />

used as a guide for assembl<strong>in</strong>g the shotgun sequenc<strong>in</strong>g fragments.<br />

The challenges faced with<strong>in</strong> the context <strong>of</strong> this thesis <strong>in</strong> identify<strong>in</strong>g new ORs from E.<br />

postvittana is reflected by the low levels <strong>of</strong> homology between the members <strong>of</strong> this<br />

group and the importance <strong>of</strong> high throughput sequenc<strong>in</strong>g technologies to such<br />

projects. With the identification and decod<strong>in</strong>g <strong>of</strong> <strong>in</strong>creas<strong>in</strong>g number <strong>of</strong> moth ORs,<br />

important questions about correlation <strong>of</strong> the phylogenetic relatedness <strong>of</strong> these prote<strong>in</strong>s<br />

with their ligands, as well as the occurrence and significance <strong>of</strong> multiple PRs <strong>in</strong> moths<br />

can be addressed.


Appendix A 134<br />

A.<br />

Appendix A – Dose<br />

response pr<strong>of</strong>ile <strong>of</strong> pIB-<br />

V5 His empty vector<br />

control to methyl<br />

salicylate<br />

Figure A.1: Dose response <strong>of</strong> the empty pIB-V5 His vector to methyl salicylate <strong>in</strong> the<br />

concentration range <strong>of</strong> 10 -5 M to 10 -13 M. Error bars represent the standard error <strong>of</strong> 6<br />

respond<strong>in</strong>g cells.


Appendix B 135<br />

Buffered TB (Sambrook et al., 1989)<br />

B.<br />

Appendix B –<br />

EpGOBP2 expression<br />

buffer recipe and<br />

prote<strong>in</strong> details<br />

Per Litre: K phosphate<br />

900mL water Per Litre:<br />

12g Tryptone 23.1g KH2PO4<br />

24g yeast Extract 125.4g K2HPO4<br />

4mL glycerol Autoclave<br />

The TB was autoclaved and cooled to room temperature. 100 mL <strong>of</strong> sterile K<br />

phosphates was then added to it.<br />

His6-EpGOBP2 has a molecular weight <strong>of</strong> 22597.20 Daltons and the am<strong>in</strong>o acid<br />

sequence is as follows:<br />

M H H H H H H S S G L V P R G S G M K E T A A A K F E R Q H M D S P D L G<br />

T D D D D K A M A D I G S E F E N L Y F Q T A E V M S H V T A H F G K A L<br />

E Q C R E E S G L S T A V L E E F Q H F W R D D F E V V H R E L G C A I L C<br />

M S N K F S L M Q D D A R M H H E N M H D Y V K S F P Q G E V L S A K M<br />

V E L I H N C E K P Y D D I K D D C E R V V K V A A C F K V D A K K A G I A<br />

P E V A M I E A V M E K Y<br />

Native EpGOBP2 has a molecular weight <strong>of</strong> 16094.51 Daltons and the am<strong>in</strong>o acid<br />

sequence is as follows:<br />

TAEVMSHVTAHFGKALEQCREESGLSTAVLEEFQHFWRDDFEVVHRELGCAI<br />

LCMSNKFSLMQDDARMHHENMHDYVKSFPQGEVLSAKMVELIHNCEKPYD<br />

DIKDDCERVVKVAACFKVDAKKAGIAPEVAMIEAVMEKY


Appendix B 136<br />

Figure B.1: Chromatogram show<strong>in</strong>g the elution pr<strong>of</strong>ile <strong>of</strong> His6-EpGOBP2 from<br />

HiTrap chelat<strong>in</strong>g HP column. The numbers <strong>in</strong> red <strong>in</strong>dicate the elution fractions and<br />

His6-EpGOBP2 eluted as a s<strong>in</strong>gle peak <strong>in</strong> fractions 16–21.<br />

Figure B.2: Elution pr<strong>of</strong>ile <strong>of</strong> His6-EpGOBP2 on Q-sepharose HP column. The<br />

numbers <strong>in</strong> red <strong>in</strong>dicate the elution fractions and His6-EpGOBP2 eluted <strong>in</strong> a s<strong>in</strong>gle<br />

peak <strong>in</strong> fractions 25–31.


Appendix C 137<br />

C.<br />

Appendix C – ClustalX multiple<br />

sequence alignment <strong>of</strong> moth PR<br />

clade<br />

MsOR1 ------MIFMDDPLSKSIKDPRDYRYMKLFRSTLRLIGSWPGR------DLKEEGATKYE<br />

BmOR-3 ------MIFVDDAVIG-IKDPREYRHLRVLRTSLRLLGAWPGH------YLGEETGSKYE<br />

HR6 -MNLRKFLFENEAVEG-INSPADYLYTRILRFNLDFIRTWPRK------ELGEPENLAFT<br />

HR16 -MGLRQFLFENEAVEG-INTASDYLYIKILRFTLVIVNSWPRK------EIGEPESPRLS<br />

HR14 MTGIRDFFFNYEAKDG-VTNPTEYPYMIMSRHLLTVITCWPKKPKEGLNARAKLRAKIWV<br />

HR15 MTGFRDFVFNYQPKDG-ITNPVDYPYLIIARYLLTFISMWPKKSVVYHSARAELKARIWL<br />

HR11 ------MHLAGNAVTG-ITGPMDYKYMKVLRFVLRIISGWPGK------ALGEKTLRIEG<br />

BmOR-5 ----MLLYYPNTQVKEKVNNVEEFTYIKFLKSFCKIMDFWP--------EREEKNSKTRI<br />

BmOR-7 ----MLLYHPNTQVEEKVNNVEEFTYMKFLKSFCKIMDFWP--------EREEKNSKTRI<br />

HR13 ----MKILSDGSDLEG-VEKVEDIFYINLARKSMWILDSWP--------KAFNASSKYRY<br />

BmOR-4 MFKIIKNIIVENDALKQVEKPQEFQYMKWVQYHLKYIDGWPNM------DMNKKNVSKIR<br />

BmOR-1 ----MLLSFKDDSRSPDIQKPQNFQYMKILRFNLKIICAWPE-------KQLNEIRSLGH<br />

EposOR1 ------------------MDVFNLKYMRMIRFTLRSIGAWPSHEFED--VPATKLTLLSS<br />

BmOR-6 -----------MKEEYYLQHPRTQLFYKVLAHVSTIESTIDLTWWG---YTFPKYVGWFY<br />

.<br />

MsOR1 IAPLYWVLVIKITCFVLTIIYLIENTNKLGFFEIGHVYITVFMTMITLSRSITLSLNPKY<br />

BmOR-3 CAPMFLLMFIKIACLYLTIVYLRNNADVLGFFELGHVYLTIFMTFVTLSRGFSLTWNPNY<br />

HR6 VFMQYFYLILNIVTVMGSTSYIVVRGSELSFIEAGLMYLIFLIGIVDTLTVVCLTFSEKF<br />

HR16 AFAKYFYLFLTVLAAIGSIAYVAVHNRELTFLETGHMYIVVLMSLVDVSRVATLTMSTTY<br />

HR14 IVQKIFHMSLCLLTTLGMAMYIGLHKKSMSFLELGHLYISLLMTVVIFSRITTLCLNPKY<br />

HR15 WVQKFYHLLLCAVAFFGGVLYITLHKKSMTFYELGHLYISLLMMACTFSRITTLCFNDEY<br />

HR11 MGHAYYNTILSLVYLALGIAYLKKNFHRFDFLELGQLYIVLLMNMLSTSRAFTLCLSQKY<br />

BmOR-5 FRLRYILVLQFCFTLVAGVLYLTNSVGKQTFYDLGHTIITVLMNVVSLSRLILR-CFKKY<br />

BmOR-7 FRLRYILVLQFCFTLVAGVLYLKNNFGKKTFYDLGHTIITVVMNVVSVSRLILR-CFKKY<br />

HR13 F----VLALN-VATLIGGAIYLRNNTGVLSSFELGHTYITVFMNCITCSRCLMI-LSKDY<br />

BmOR-4 FHKRHLLVVEQTITFLSQMFYIVKNYGKLSFFEIGHSYITALMTIVIFSRSVVT-ALGRY<br />

BmOR-1 SIHRVILPIQSVVCLACGILYIHFHFNEIPFFILASTFITVMMNLVTCSRTALVMLFERY<br />

EposOR1 YSYTCFLCIICCFGTIAQIAYLITNQGILGFIDLGQTYLTVLMCFVYIQRTMLP-LQTSY<br />

BmOR-6 HLQCNVVRLFGKCVVVSQILFIILNYQTIDKSVFIIAITITPLGALVGIKAESA-KAECY<br />

:: : :<br />

MsOR1 RRVMTKYITKMHLFYYK---DMSDIALKTHIRVHKLSHFFTMYLSTQVVLGTVTFNIVPM<br />

BmOR-3 HKVVKKFITEMHLLYFK---DNSEYAMKTHRRVHKISHFYTVFLKVQMIAGLTLFNVIPM<br />

HR6 RVLAKDFLTKTHLFYYK---DRSKHAMEIHKKIHLISHLFSLWILFQMLSGLSLFNIIPM<br />

HR16 REVARDFLTKIHLFYYK---DRSKHAMETHRAVHKISHLFTLWLVGQMLSGLSLFNLIPM<br />

HR14 RAVSTEFLTKIHLFYYK---DDSEFSMQIHKQVHKISHLFTLYLTGQMIAGLSLFNLTPM<br />

HR15 RVVAKDFVTKIHLFFYK---NRSDYSMQIHKKVHMISHVFTLYLSGQMMLGLFLFNVTPM<br />

HR11 REVAKIFIQKIHLFYFK---EKSDYAMKIHVIVHKISFISAVYLSVLLFIAAVMFNLIPM<br />

BmOR-5 DVVGQQFINKIHLYHYR---NDSEYAMKIHTVVHKISHNMTYIFSFCIIFGTVTFNLTPI<br />

BmOR-7 DVVGQQFINKIHLYHFR---NDSEYSMKTYKAVHKISNNMTYIFSFSIFVCVVTFNLNPV<br />

HR13 NHVMTLFVQKIHLFHHK---HKSDYAYLTHIFIHKISHFYTVYLLGLALNGLFLFNMIPF<br />

BmOR-4 RKIARYFVSSLHLYHYK---DISEYALQTHLLVHRLSHYYTVYLISLVVTGMLLFNITPL<br />

BmOR-1 LVLTGRFITVMHLFNFQ---KNSDYAYKLCTFVNRMSHFYTLYVLFSMFMGLGLFNLLPL<br />

EposOR1 QAEIIEFSTKFHLMYHK---NETEFAAKMHNKVKRICEIVTGIQHLQIYYVLVMYNIAPL<br />

BmOR-6 VNLMKNFMDKVHIHSIYRKNENNEFVKKKVIQIERVSRFTAYFLVILIAINCLSWMLKPT<br />

: *: . .. :. :. : : : *<br />

MsOR1 YNNYKVGRFEN-NILVNDSYELSIYFKTPTKFLSTLNGYIAITTFNWYSSYICSNFFCMF<br />

BmOR-3 YNNYRQGNYAS-DRPANITYDLSIYYET-FDILNTPNGYIFICVFNWFASYICCSFFCSF<br />

HR6 YSNLAAGKYRK-GGLQNSTFEHSLYYLYPFNTSTDITGYIIACILHWIISYLCSCWFCII<br />

HR16 YSNYAAGRYSG-DVSKNSTFEHSLYYSYPFDTSTDIRGYSIACVIHWVLSYLCSTWFCMF<br />

HR14 YNNFSAGKYKK-GGLKNSTFEHSLYYSYPFNASSDVGGYIVSNICDWIISYLCSTWFCTL<br />

HR15 YNNYSAGKYKS-GGLKNSTYEHSLYFSWPFNASTDMRGYIVSNILNWMLSYTCSSWFCVI<br />

HR11 YNNYSAGRYSSFDNLENTTYEQAISCLYPWNFETNFNGYLVATLSGWYGTMLCGSSVSMF<br />

BmOR-5 FNNIGSDAYKN-PRPDNVTLQQCVYYALPFDYTGNFKWYLLVAIFNVQKTFFCTSLFILF<br />

BmOR-7 FNNIGSGAYKN-PRPDNVTLQQCVYYALPFDYTGDFKWYMLVAIFNVQKTFFCTSLFILF<br />

HR13 YNCYSRGMFRD-VIPANATYDHAVFYSVPFDYTTKFKGYLAMTSFNVFISYTCTSYFCVV<br />

BmOR-4 YNNISSGVFNS-PRPENMTFQHAVYLGLPFDYTTDIKGYFVVFILNWHLSHIAASYFCTF<br />

BmOR-1 YNNYVSGAFSD-PYGPNVTFFHSVYFAFPFDYSHNFRGYIIMALFNSYVSVTCSIGLVMF


Appendix C 138<br />

EposOR1 YSNFKSGMLSS-EKPINGTYEHSVYYVLPFDHNNEV-WYPVVGLYNFYVSYNLGAMFSCH<br />

BmOR-6 LHNIK---HFEEIMNKSMEFQYYIYFWTPLDYKYNLRDYIIIHTLCIYLGATAVTVIVTF<br />

. : . * .<br />

MsOR1 DLALSLLIFTVSGHFKILIHNLNNFPLPAVVSDSSKVLKTDEIQ----------APLYNK<br />

BmOR-3 DLILSLMISTVSGHFRILIHNLLTFPLPEAITASKKFVDKHRCNGNRSEFVLEEAKLYSP<br />

HR6 NLFLSLLVFNLWGHFKILISTLNEFPRPSSKSVDTQESP----------------YKYTE<br />

HR16 DLFLSLMVFHLWGHFKILINTLNDFPRPSSKVEGAQ---------------------FSD<br />

HR14 DLFLSIMVFHVWGHFKILLHDLDHFPRPANLTTFKLDNSN--IT--------LTSEKFSS<br />

HR15 DFFLSLMVFHIWGHFKILLHDLDHFPRPLNKVNSVIEDS---IT--------ITNEMYSQ<br />

HR11 DLFLCLMIFNLWGHFKILIHNLEHFPRPASEIVDAEGAERSGRI--------IGSEMYSQ<br />

BmOR-5 ELSLSLMIICLWGHLRIFIHNLNHIPAPRNSFE------------------------YTK<br />

BmOR-7 DLLLSMMIIHLWGHIRIFIHNLNHIPAPRNSLE------------------------YTR<br />

HR13 DLTISLVIFHLWGHMRLLTYHLANFKKPASVLESNDNNKDEIKD-----------HSYTE<br />

BmOR-4 DLFLSLLILHLWGHLRIILNNLKTFPKPYTNNS-----------------------MYTE<br />

BmOR-1 DLLMCLMVMHVWGHLKILSHNLINFPRPKASHVITTPNGPTN-V-----------ETYTE<br />

EposOR1 DLLISVYVFHIWGHLNICEHNLNNFPRPSITRNSKTVPLR-----------------YSA<br />

BmOR-6 DIFNFIAVFHVVAHIQILKNNVKSNWSDD----------------------------FNE<br />

:: : : : .*:.: : :.<br />

MsOR1 TEKKDITLRLKQCIDYHREVLEFTQDISEAFGPMLFVYYLFHQVSGCLLLLECSQMDAAA<br />

BmOR-3 AEMWQVTDRLRQCIDYHRKLVEFTGDISEAFGPMLFVYYLFHQVSGCLLLLECSQLNTAA<br />

HR6 EELIEVAEKLKDCINYHREIKIFTNRMSDVFGPMLFIYYAFHQASGCLLLLECSQMTARA<br />

HR16 EELVDVAARLKDCIVYHREITLFTDRMSNVFGPMLFVYYSFHQASGCLLLLECSQMTAQA<br />

HR14 IELGQVSEKLKKCIEYHRKIVSFTDEMSEVFGPMLFVYYGFHQTSGCLLLLECSQMTVAA<br />

HR15 TELDQVFDRLGKCIDYHREIVSFTDKMSEVFGPMLFAYYGFHQASGCLLLLECSQMTVAA<br />

HR11 AELEQVAVLLRECIQYHMLIFDFTNNMSDAFGMALFIYYSFHQITGCLLLLECSQMTAAA<br />

BmOR-5 EERQEVDDTLKKCIQHHTLIIGFVRIMSETYGLAVLIYYAFQQVVGCLLLLQCSQMELKT<br />

BmOR-7 EERQEVDNTLKKCIQHHTLIIGFVRIMSETYGLAVLIYYAFQQVVGCLLLLQCSRLDLKT<br />

HR13 EELKEVFSKLREYIQHHNLILEFSSEMSNAFGPALLAYMVFHQVSGCILLLECSQLDTKT<br />

BmOR-4 EENQVVLLKLQECIRYHNFIISFTVMMSNVYDVVIIVYYLFHQVTGCLLLLQCSTLDWES<br />

BmOR-1 EESKEVFARLRECIKHYGTVDDFANDMSETFGVILLVYYGFHQVSLCMLLLECSDLSTKA<br />

EposOR1 EENKKVAAGLKEIIIHYIMIKKFVEKTSNTYSVTLCFYYGFHMVAECILLLQCSTLEVEA<br />

BmOR-6 SEKK---GYLVSILEYHAYIIRIFGEVQSAFGLNVASNYLQNLIEDGLFLYQIMNGEKEN<br />

* * . : :: : : ...:. : : ::* :<br />

MsOR1 LMRYGLLTAVLFQQLIQLSVVVESVGTVTGYLKDAVYNVPWEYMDTQDRKTVCIFLMNVQ<br />

BmOR-3 LVRYGVLTVVLYQQLIQLSVIVESVGTVTGRLKDAVYEVPWEYMDTSNRKTVAIFLMNVQ<br />

HR6 LMRYLPLTIIMLQQLIQLSVIFELVGTESEKLKDAVYGVPWDCMDTKNRKVVMFFLMNVQ<br />

HR16 LMRYVPLTIILTQQLIQLSVIFELVGSESDKLKHAVYGLPWECMDVKNRRVVLIFLANTQ<br />

HR14 LVCYLPLTIMLFQQLIQLSIIFELVGSVSDKLKDAVYSLPWEAMDIKNKKTVAIFLMNVQ<br />

HR15 LVRYLPLTIILFQQLIQMSIIFELVGSVTDKLRDAVYGLPWEAMDTKNRKTVAFFLMNVQ<br />

HR11 LTRYLPLTIIMFGELVLLSIIFETIGTMSEKLKDAVYKVPWEYMDTKNRRTLLIFLIKVQ<br />

BmOR-5 VTRFGFLTLVLNQQLIQISVIFELLGYMSDKLQDAVYCVPWEYMDTSHRKMVYMMFRQSQ<br />

BmOR-7 ITRFGFLTTMVNQQLIQISVIFELLGYMNDKLQEAVYCVPWEYMDTSHRKMVYMMFRQSQ<br />

HR13 LVRYGPLTIVIFQQLIQISVIFELLGSSNDKLIDGVYLVPWEYMDTKNRKLVFTMLRQSH<br />

BmOR-4 LSRYGPLTLIIFQQLIQVSMIFEILGFLSDKLPNAVYSIPWEAMNVTNRKLVQVLLQKSQ<br />

BmOR-1 MLRYGPLTLIMIQQLIQISIIFELLGSVADRIPDAVYQLPWECMDVKNRRVVYGFLRRTQ<br />

EposOR1 LAKYGFLTVAVYQELIQLSVVFELIYAKGTSLIDAVYGLPWECMDNSSRRTVLILLQIVQ<br />

BmOR-6 VLMYGLMIILYLGGLIFLSIVLEEIRRQNYDLCEYVYALPWEGMSLENQKIFVVFLQRTQ<br />

: : : *: :*::.* : : . ** :**: *. :: . :: :<br />

MsOR1 EPVHINALGLAKVGVQAMAGILKTSFSYFAFLRTVSN---<br />

BmOR-3 EPLHVNALGLAKVGVQSMAAILKTSFSYFTFLRTVSE---<br />

HR6 EPVHVKAMGLANVGVTTMASILKTSLSYFTFLLSQTKEE-<br />

HR16 EPVHVKAMGVANVGVTSMAAILKTSMSYFTFLRSM-----<br />

HR14 EPVHVKALGLAEVGVTSMTAILKTSMSYFTFLRSK-----<br />

HR15 EPVHVKALGLAEVGVTSMTAILKTSMSYFAFLRSM-----<br />

HR11 EPIHVKAGGLVDVGVTTMASILKTSFSYFAFLRTF-----<br />

BmOR-5 IPLQLKAMNMLSIGVKTMVSILKTSVTYYLILKTVTTD--<br />

BmOR-7 IPLQLKAMNMLSIGVKTMASILKTSVTYYLMLKTITANEA<br />

HR13 RSINLTMMSMVTVGVQTMTAILKTSFSYFVMLKTVAEEE-<br />

BmOR-4 KPIQFKAMNMMSVGVQTMASIIKTSISYFIMLRTIARD--<br />

BmOR-1 NPVRFKAMGMLDVGVQTMASILKTSISYFVMLRTVAT---<br />

EposOR1 QPLSLKACGMVPVGIQTMQAILKGSFSYFLMLRTFANQ--<br />

BmOR-6 PDLEFETVCGMKAGVKPAFSIVKSMFSYYVMINSRF----<br />

: . *: . .*:* .:*: :: :<br />

Figure C.1: ClustalX multiple sequence alignment <strong>of</strong> the moth PR clade, <strong>in</strong>clud<strong>in</strong>g<br />

members from B. mori, M. sexta, E. postvittana and H. virescens (Krieger et al., 2002;<br />

Krieger et al., 2004; Nakagawa et al., 2005; Wanner et al., 2007; Jordan et al., 2009;<br />

Patch et al., 2009). Am<strong>in</strong>o acids <strong>in</strong> red <strong>in</strong>dicate region <strong>of</strong> degenerate primer design for<br />

identify<strong>in</strong>g PR genes <strong>in</strong> E.postvittana.


Appendix D 139<br />

D.<br />

Appendix D – Solutions for<br />

nuclear DNA isolation<br />

Table D.1: Reagents for mak<strong>in</strong>g 1 L <strong>of</strong> Nuclei Extraction Buffer, pH 6.<br />

Reagent 1 L F<strong>in</strong>al concentration<br />

Mannitol 91g 0.5 M<br />

PIPES-KOH 3.78 g 10 mM<br />

MgCl2 7H2O 2.03 10 mM<br />

PVP K30 20 g 2%<br />

L-lys<strong>in</strong>e monohydrochloride 36.52 g 200 mM<br />

EGTA 2.28 g 6 mM<br />

Sodium meta bisulphite 1.9 g 10 mM<br />

800 mL <strong>of</strong> deionized water was heated up <strong>in</strong> the microwave for 2-4 m<strong>in</strong>utes and<br />

poured <strong>in</strong>to a glass beaker with a magnetic stirrer. PVP K30 was added to the beaker<br />

with vigorous stirr<strong>in</strong>g. Mannitol was added and completely dissolved before addition<br />

<strong>of</strong> PIPES, magnesium, lys<strong>in</strong>e, and EGTA. Volume was adjusted to 1 L, pH 6. The<br />

buffer was split <strong>in</strong>to two 1 L flasks, autoclaved and cooled to room temperature. The<br />

buffer was kept at 4°C until required.<br />

Just before start<strong>in</strong>g the nuclei isolation, 0.9 g sodium meta-bisulphite was added to<br />

each flask. To one flask 0.2 mL -mercaptoethanol was added and this flask labelled<br />

as NEB complete buffer, while the other flask is NEB– no -mercaptoethanol. Both<br />

the flasks were kept on ice until used.<br />

The lysis buffer <strong>in</strong> part II <strong>of</strong> nuclear DNA isolation consists <strong>of</strong> 0.5% SDS, 5 mM<br />

EDTA, 150 mM Tris-borate pH 7.4. To prepare this solution the Tris buffer was<br />

titrated with boric acid, <strong>in</strong>stead <strong>of</strong> the usual HCl, autoclaved and kept at room<br />

temperature.


Appendix E 140<br />

E.<br />

Appendix E – Multiple<br />

sequence alignment <strong>of</strong> E.<br />

postvittana ORs together with<br />

ORs from five other moth<br />

species


Appendix E 141<br />

I II


Appendix E 142<br />

III IV


Appendix E 143<br />

V


Appendix E 144<br />

VI VII


Appendix E 145<br />

VII cont<strong>in</strong>ued<br />

Figure E.1: Multiple sequence alignment <strong>in</strong> ClustalX <strong>of</strong> E. postvittana ORs with<br />

those <strong>of</strong> five other species B. mori (Wanner et al., 2007), H. virescens (Krieger et al.,<br />

2002, Krieger et al., 2004) and P. xylostella, M. separata, and D. <strong>in</strong>dica (Mitsuno et<br />

al., 2008). The sequences were obta<strong>in</strong>ed from GenBank. Transmembrane positions<br />

obta<strong>in</strong>ed from consensus <strong>of</strong> EpOR1, 2, 3 and 34 are shown as orange bars above the<br />

alignment.


References<br />

Abraham, D., L<strong>of</strong>stedt, C. and Picimbon, J.-F. (2005). Molecular characterization and<br />

evolution <strong>of</strong> pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> genes <strong>in</strong> Agrotis moths. Insect<br />

Biochemistry and Molecular Biology 35(10): 1100-1111.<br />

Adessi, C. (2000). Solid phase DNA amplification: characterisation <strong>of</strong> primer<br />

attachment and amplification mechanisms. Nucleic Acids Res. 28: e87.<br />

Altner, H., Sass, H. and Altner, I. (1977). Relationship between structure and function<br />

<strong>of</strong> antennal chemo-, hygro-, and thermoreceptive sensilla <strong>in</strong> Periplaneta<br />

americana. Cell and Tissue Research 176: 389-405.<br />

Altschul, S. F., Gish, W., Miller, W., Myers, E. W. and Lipman, D. J. (1990). Basic<br />

local alignment search tool. Journal <strong>of</strong> Molecular Biology 215(3): 403-410.<br />

Anderson, A. R., Wanner, K. W., Trowell, S. C., Warr, C. G., Jaqu<strong>in</strong>-Joly, E., Zagatti,<br />

P., Robertson, H. and Newcomb, R. D. (2009). Molecular basis <strong>of</strong> female-<br />

specific odorant responses <strong>in</strong> Bombyx mori. Insect Biochemistry and<br />

Molecular Biology 39(3): 189-197.<br />

Anderson, P., Hallberg, E. and Subchev, M. (2000). Morphology <strong>of</strong> antennal sensilla<br />

auricillica and their detection <strong>of</strong> plant volatiles <strong>in</strong> the Herald moth,<br />

Scoliopteryx libatrix L. (Lepidoptera: Noctuidae) Arthropod Structure and<br />

Development 29: 33-41.<br />

Anderson, P., Hansson, B. S. and L<strong>of</strong>qvist, J. (1995). Plant-odour-specific receptor<br />

neurones on the antennae <strong>of</strong> female and male Spodoptera littoralis.<br />

Physiological Entomology 20: 189-198.<br />

Angioy, A. M., Desogus, A., Barbarossa, I. T., Anderson, P. and Hansson, B. S.<br />

(2003). Extreme Sensitivity <strong>in</strong> an Olfactory System. Chemical Senses 28(4):<br />

279-284.


References 147<br />

Ansebo, L., Ignell, R., Löfqvist, J. and Hansson, B. S. (2005). Responses to sex<br />

pheromone and plant odours by olfactory receptor neurons housed <strong>in</strong> sensilla<br />

auricillica <strong>of</strong> the codl<strong>in</strong>g moth, Cydia pomonella (Lepidoptera: Tortricidae).<br />

Journal <strong>of</strong> Insect Physiology 51(10): 1066-1074.<br />

Arn, H., Toth, M. and Priesner, E. (1986). List <strong>of</strong> Sex Pheromones <strong>of</strong> Lepidoptera and<br />

Related Attractants. International Organisation for Biological Control, West<br />

Palearctic Regional Section.<br />

Arn, H., Toth, M. and Priesner, E. (1992). List <strong>of</strong> Sex Pheromones <strong>of</strong> Lepidoptera and<br />

Related Attractants. International Organisation for Biological Control,<br />

Montfavet, France.<br />

Baker, T. C. (2009). Nearest neural neighbors: moth sex pheromone receptors HR11<br />

and HR13. Chemical Senses 34: 465-468.<br />

Bakke, A. and Lie, R. (1989). Evaluation and use <strong>of</strong> behaviour-modify<strong>in</strong>g chemicals.<br />

Insect Pheromones <strong>in</strong> Plant Protection. A. R. Jutsum and R. F. S. Gordon,<br />

Wiley-Interscience.<br />

Bellas, T. E., Bartell, R. J. and Hill, A. (1983). Identification <strong>of</strong> two components <strong>of</strong><br />

the sex pheromone <strong>of</strong> the moth, Epiphyas postvittana (Lepidoptera,<br />

Tortricidae). Journal <strong>of</strong> Chemical Ecology 9(4): 503-512.<br />

Benton, R. (2006). On the ORig<strong>in</strong> <strong>of</strong> smell: odorant receptors <strong>in</strong> <strong>in</strong>sects. Cellular and<br />

Molecular Life Sciences 63(14): 1579-1585.<br />

Benton, R., Sachse, S., Michnick, S. W. and Vosshall, L. B. (2006). Atypical<br />

Membrane Topology and Heteromeric Function <strong>of</strong> Drosophila Odorant<br />

Receptors In Vivo. PLoS Biol 4(2): e20.<br />

Benton, R., Vannice, K. and Vosshall, L. (2007). An essential role for a CD36-related<br />

receptor <strong>in</strong> pheromone detection <strong>in</strong> Drosophila. Nature 450(7167): 289.


References 148<br />

Bordier, C. (1980). Phase separation <strong>of</strong> <strong>in</strong>tegral membrane prote<strong>in</strong>s <strong>in</strong> triton X-114<br />

solution. Journal <strong>of</strong> Biological Chemistry 256: 1604–1607.<br />

Bork, P., Dandekar, T., Diaz-Lazcoz, Y., Eisenhaber, F., Huynen, M. and Yuan, Y.<br />

(1998). Predict<strong>in</strong>g function: from genes to genomes and back. Journal <strong>of</strong><br />

Molecular Biology 283(4): 707-725.<br />

Braslavsky, I., Hebert, B., Kartalov, E. and Quake, S. R. (2003). Sequence<br />

<strong>in</strong>formation can be obta<strong>in</strong>ed from s<strong>in</strong>gle DNA molecules. Proc. Natl. Acad.<br />

Sci. USA 100: 3960-3964.<br />

Breer, H. (1997). Molecular mechanisms <strong>of</strong> pheromone reception <strong>in</strong> <strong>in</strong>sect antennae.<br />

Insect Pheromone Research: New Directions. R. T. Carde and A. K. M<strong>in</strong>ks.<br />

New York, Chapman and Hall.<br />

Briand, L., Nespoulous, C., Huet, J.-C., Takahashi, M. and Pernollet, J.-C. (2001).<br />

Ligand b<strong>in</strong>d<strong>in</strong>g and physico-chemical properties <strong>of</strong> ASP2, a recomb<strong>in</strong>ant<br />

odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> from honeybee (Apis mellifera L.). European Journal<br />

<strong>of</strong> Biochemistry 268(3): 752-760.<br />

Briand, L., Nespoulous, C., Perez, V., Rémy, J.-J., Huet, J.-C. and Pernollet, J.-C.<br />

(2000). Ligand-b<strong>in</strong>d<strong>in</strong>g properties and structural characterization <strong>of</strong> a novel rat<br />

odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> variant. European Journal <strong>of</strong> Biochemistry 267(10):<br />

3079-3089.<br />

Bruce, T. J. A., Wadhams, L. J. and Woodcock, C. M. (2005). Insect host location: a<br />

volatile situation. Trends Plant Sci 10: 269–274.<br />

Butenandt, A., Beckmann, R. and Hecker, E. (1961). On the sex attractant <strong>of</strong> silk-<br />

moths. I. The biological test and the isolation <strong>of</strong> the pure sex-attractant<br />

bombykol. Hoppe Seylers Zeitschrift fur Physiologische Chemie 324: 71-83.


References 149<br />

Butenandt, A., Beckmann, R., Stamm, D. and Hecker, E. (1959a). Über den<br />

sexuallockst<strong>of</strong>f des seidensp<strong>in</strong>ners Bombyx mori. Re<strong>in</strong>darstellung und<br />

konstitution. Zeitschrift fur Naturforschung B 14: 283-284.<br />

Butenandt, A., Groschel, U., Karlson, P. and Zillig, W. (1959b). N-acetyl tyram<strong>in</strong>e, its<br />

isolation from Bombyx cocoons & its chemical & biological properties.<br />

Archives <strong>of</strong> Biochemistry and Biophysics 83: 76-83.<br />

Callahan, F. E., Vogt, R. G., Tucker, M. L., Dickens, J. C. and Mattoo, A. K. (2000).<br />

High level expression <strong>of</strong> "male specific" pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s (PBPs)<br />

<strong>in</strong> the antennae <strong>of</strong> female noctuiid moths. Insect Biochemistry and Molecular<br />

Biology 30(6): 507-514.<br />

Campion, D. G., Critchley, B. R. and McVeigh, L. J. (1989). Evaluation and use <strong>of</strong><br />

behaviour-modify<strong>in</strong>g chemicals. Insect Pheromones <strong>in</strong> Plant Protection. A. R.<br />

Jutsum and R. F. S. Gordon, Wiley-Interscience.<br />

Claros, M. G. and von Heijne, G. (1994). TopPred II: An improved s<strong>of</strong>tware for<br />

membrane prote<strong>in</strong> structure predictions. CABIOS 10: 685–686.<br />

Clyne, P. J., Warr, C. G., Freeman, M. R., Less<strong>in</strong>g, D., Kim, J. and Carlson, J. R.<br />

(1999). A Novel Family <strong>of</strong> Divergent Seven-Transmembrane Prote<strong>in</strong>s:<br />

Candidate Odorant Receptors <strong>in</strong> Drosophila. Neuron 22(2): 327-338.<br />

Couto, A., Alenius, M. and Dickson, B. J. (2005). Molecular, Anatomical, and<br />

Functional Organization <strong>of</strong> the Drosophila Olfactory System. Current Biology<br />

15(17): 1535-1547.<br />

Cserzo, M., Wall<strong>in</strong>, E., Simon, I., von Heijne, G. and El<strong>of</strong>sson, A. (1997). Prediction<br />

<strong>of</strong> transmembrane alpha-helices <strong>in</strong> prokaryotic membrane prote<strong>in</strong>s: the dense<br />

alignment surface method. Prote<strong>in</strong> Eng<strong>in</strong>eer<strong>in</strong>g 10(6): 673–676.


References 150<br />

Danthanarayana, W. (1975). The bionomics, distribution and host range <strong>of</strong> the light<br />

brown apple moth, Epiphyas postvittana (Walk.) (Tortricidae). Australian<br />

Journal <strong>of</strong> Zoology 23: 419-437.<br />

Das, P. D., Ra<strong>in</strong>a, R., Prasad, A. R. and Sen, A. (2007). Electroantennogram<br />

responses <strong>of</strong> the potato tuber moth, Phthorimaea operculella (Lepidoptera:<br />

Gelichiidae) to plant volatiles. J Biosci 32: 339–349.<br />

Dess, D. B. and Mart<strong>in</strong>, J. C. (1983). Readily accessible 12-I-5 oxidant for the<br />

conversion <strong>of</strong> primary and secondary alcohols to aldehydes and ketones.<br />

Journal <strong>of</strong> Organic Chemistry 48: 4155–4156.<br />

Dobritsa, A. A., van der Goes van Naters, W., Warr, C. G., Ste<strong>in</strong>brecht, R. A. and<br />

Carlson, J. R. (2003). Integrat<strong>in</strong>g the Molecular and Cellular Basis <strong>of</strong> Odor<br />

Cod<strong>in</strong>g <strong>in</strong> the Drosophila Antenna. Neuron 37(5): 827-841.<br />

Dohm, J. C., Lottaz, C., Borod<strong>in</strong>a, T. and Himmelbauer, H. (2007). SHARCGS, a fast<br />

and highly accurate short-read assembly algorithm for de novo genomic<br />

sequenc<strong>in</strong>g. Genome Res. 17: 1697-1706.<br />

Dressman, D., Yan, H., Traverso, G., K<strong>in</strong>zler, K. W. and Vogelste<strong>in</strong>, B. (2003).<br />

Transform<strong>in</strong>g s<strong>in</strong>gle DNA molecules <strong>in</strong>to fluorescent magnetic particles for<br />

detection and enumeration <strong>of</strong> genetic variations. Proc. Natl. Acad. Sci. USA<br />

100: 8817-8822.<br />

Ebb<strong>in</strong>ghaus, D., Losel, P. M., L<strong>in</strong>demann, M., Scherkenbeck, J. and Zebitz, C. P. W.<br />

(1998). Detection <strong>of</strong> major and m<strong>in</strong>or sex pheromone components by the male<br />

codl<strong>in</strong>g moth Cydia pomonella (Lepidoptera : Tortricidae). Journal <strong>of</strong> Insect<br />

Physiology 44: 49-58.<br />

El-Sayed, A. M. (2008). The Pherobase: database <strong>of</strong> <strong>in</strong>sect pheromones and<br />

semiochemicals. www.pherobase.com.


References 151<br />

Fedurco, M., Romieu, A., Williams, S., Lawrence, I. and Turcatti, G. (2006). BTA, a<br />

novel reagent for DNA attachment on glass and efficient generation <strong>of</strong> solid-<br />

phase amplified DNA colonies. Nucleic Acids Res. 34: e22.<br />

Felsenste<strong>in</strong>, J. (2005). PHYLIP (Phylogeny Inference Package) version 3.65.<br />

http://evolution.gs.wash<strong>in</strong>gton.edu/phylip.html.<br />

Feng, L. and Prestwich, G. D. (1997). Expression and characterization <strong>of</strong> a<br />

lepidopteran general odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>. Insect Biochemistry and<br />

Molecular Biology 27(5): 405-412.<br />

Fishilevich, E. and Vosshall, L. (2005). Genetic and functional subdivision <strong>of</strong> the<br />

Drosophila antennal lobe. Curr. Biol 15: 1548–1553.<br />

Forstner, M., Breer, H. and Krieger, J. (2009). A receptor and b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

<strong>in</strong>terplay <strong>in</strong> the detection <strong>of</strong> a dist<strong>in</strong>ct pheromone component <strong>in</strong> the silkmoth<br />

Antheraea polyphemus. Int J Biol Sci 5: 745–757.<br />

Foster, S. P. (1993). Neural <strong>in</strong>activation <strong>of</strong> sex pheromone production <strong>in</strong> mated<br />

lightbrown apple moths, Epiphyas postvittana (Walker). Journal <strong>of</strong> Insect<br />

Physiology 39(3): 267-273.<br />

Foster, S. P. and Dugdale, J. S. (1988). A comparison <strong>of</strong> morphological and sex<br />

pheromone differences <strong>in</strong> some New Zealand Tortric<strong>in</strong>ae moths. Biochemical<br />

Systematics and Ecology 16(2): 227-232.<br />

Foster, S. P. and Greenwood, D. R. (1997). Change <strong>in</strong> reductase activity is responsible<br />

for senescent decl<strong>in</strong>e <strong>in</strong> sex pheromone titre <strong>in</strong> the Lightbrown Apple Moth,<br />

Epiphyas postvittana (Walker). J. Insect Physiol. 43(12): 1093 - 1100.<br />

Foster, S. P. and Roel<strong>of</strong>s, W. L. (1987). Sex pheromone differences <strong>in</strong> populations <strong>of</strong><br />

the brownheaded leafroller, Ctenopseustis obliquana. Journal <strong>of</strong> Chemical<br />

Ecology 13(3): 623-629.


References 152<br />

Foster, S. P. and Roel<strong>of</strong>s, W. L. (1990). Biosynthesis <strong>of</strong> a monoene and a conjugated<br />

diene sex pheromone component <strong>of</strong> the lightbrown apple moth by Δ11<br />

desaturation. Cellular and Molecular Life Sciences 46(3): 269-273.<br />

Fraser, A. M., Mechaber, W. L. and Hildebrand, J. G. (2003). Electroantennographic<br />

and behavioural responses <strong>of</strong> the sph<strong>in</strong>x moth Manduca sexta to host plant<br />

headspace volatiles J Chem Ecol 29: 1813–1833.<br />

Gao, Q. and Chess, A. (1999). Identification <strong>of</strong> candidate Drosophila olfactory<br />

receptors from genomic DNA sequence. Genomics 60: 31–39.<br />

Geier, P. W. and Briese, D. T. (1980). The light-brown apple moth, Epiphyas<br />

postvittana (Walker): 4. studies on population dynamics and <strong>in</strong>juriousness to<br />

apples <strong>in</strong> the Australian Capital Territory. Australian Journal <strong>of</strong> Ecology 5(1):<br />

63-93.<br />

Gilbert, W. (1978). Why genes <strong>in</strong> pieces? Nature 271(5645): 501-501.<br />

Gildemeister, E. U. and H<strong>of</strong>fman, F. (1966). Die Atherischen Ole. Akademie Verlag,<br />

Berl<strong>in</strong>. IIId: 258.<br />

Gochnauer, T. A., Boch, R. and Margetts, V. J. (1979). Inhibition <strong>of</strong> Ascosphaera<br />

apis by citral and geraniol. Journal <strong>of</strong> Invertebrate Pathology 34(1): 57-61.<br />

Green, P. (1997). Aga<strong>in</strong>st a Whole-Genome Shotgun. Genome Research 7(5): 410-<br />

417.<br />

Große-Wilde, E., Gohl, T., Bouché, E., Breer, H. and Krieger, J. (2007). Candidate<br />

pheromone receptors provide the basis for the response <strong>of</strong> dist<strong>in</strong>ct antennal<br />

neurons to pheromonal compounds. European Journal <strong>of</strong> Neuroscience 25(8):<br />

2364-2373.


References 153<br />

Große-Wilde, E., Stieber, R., Forstner, M., Krieger, J., Wicher, D. and Hansson, B. S.<br />

(2010). Sex-specific odorant receptors <strong>of</strong> the Tobacco Hornworm Manduca<br />

sexta. Frontiers <strong>in</strong> Cellular Neuroscience 4: 12.<br />

Große-Wilde, E., Svatos, A. and Krieger, J. (2006). A Pheromone-B<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong><br />

Mediates the Bombykol-Induced Activation <strong>of</strong> a Pheromone Receptor In<br />

Vitro. Chem. Senses 31(6): 547-555.<br />

Guo, Y.-L. Isolation <strong>of</strong> nuclear DNA from plants. Retrieved August, 2009.<br />

http://my.jgi.doe.gov/general/protocols/DNA_Isolation_<strong>of</strong>_Nuclear_DNA_<strong>of</strong>_<br />

Arabidopsis_lyrata.pdf.<br />

Hallem, E. A. and Carlson, J. R. (2006). Cod<strong>in</strong>g <strong>of</strong> Odors by a Receptor Repertoire.<br />

Cell 125(1): 143-160.<br />

Hallem, E. A., Ho, M. G. and Carlson, J. R. (2004). The Molecular Basis <strong>of</strong> Odor<br />

Cod<strong>in</strong>g <strong>in</strong> the Drosophila Antenna. Cell 117(7): 965-979.<br />

Hamiaux, C., Stanley, D., Greenwood, D. R., Baker, E. N. and Newcomb , R. D.<br />

(2009). Crystal Structure <strong>of</strong> Epiphyas postvittana Takeout 1 with Bound<br />

Ubiqu<strong>in</strong>one Supports a Role as Ligand Carriers for Takeout Prote<strong>in</strong>s <strong>in</strong><br />

<strong>Insects</strong>. Journal <strong>of</strong> Biological Chemistry 284: 3496-3503.<br />

Harris, T. D. (2008). S<strong>in</strong>gle-molecule DNA sequenc<strong>in</strong>g <strong>of</strong> a viral genome. Science<br />

320: 106-109.<br />

Hartlieb, E. and Anderson, P. (1999). Olfactory-Released Behaviours. Insect<br />

<strong>Olfaction</strong>. B. Hansson. Heidelberg, Spr<strong>in</strong>ger-Verlag.<br />

Haynes, K. F., Li, W. G. and Baker., T. C. (1986). Control <strong>of</strong> p<strong>in</strong>k bollworm moth<br />

(Lepidoptera: Gelechiidae) with <strong>in</strong>secticides and pheromones (attracticide):<br />

lethal and sublethal effects. J. Econ. Entomol 79: 1466-1471.


References 154<br />

He, X., Tzotzos, G., Woodcock, C., Pickett, J., Hooper, T., Field, L. and Zhou, J.-J.<br />

(2010). B<strong>in</strong>d<strong>in</strong>g <strong>of</strong> the General Odorant B<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong> <strong>of</strong> Bombyx mori<br />

BmorGOBP2 to the Moth Sex Pheromone Components. Journal <strong>of</strong> Chemical<br />

Ecology: 1-13.<br />

He<strong>in</strong>bockel, T. and Kaissl<strong>in</strong>g, K.-E. (1996). Variability <strong>of</strong> olfactory receptor neuron<br />

responses <strong>of</strong> female silkmoths (Bombyx mori L.) to benzoic acid and (±)-<br />

l<strong>in</strong>alool. Journal <strong>of</strong> Insect Physiology 42(6): 565-578.<br />

Henn<strong>in</strong>gsen, R., Gale, B. L., Straub, K. M. and DeNagel, D. C. (2002). Application <strong>of</strong><br />

zwitterionic detergents to the solubilization <strong>of</strong> <strong>in</strong>tegral membrane prote<strong>in</strong>s for<br />

two dimensional gel electrophoresis and mass spectrometry. Proteomics 2:<br />

1479–1488.<br />

Hildebrand, J. G. and Shepherd, G. M. (1997). <strong>Mechanisms</strong> <strong>of</strong> olfactory<br />

discrim<strong>in</strong>ation: converg<strong>in</strong>g evidence for common pr<strong>in</strong>ciples across phyla.<br />

Annual Review <strong>of</strong> Neuroscience 20: 595-631.<br />

Hill, C. A., Fox, A. N., Pitts, R. J., Kent, L. B., Tan, P. L., Chrystal, M. A., Cravchik,<br />

A., Coll<strong>in</strong>s, F. H., Robertson, H. M. and Zwiebel, L. J. (2002). G Prote<strong>in</strong>-<br />

Coupled Receptors <strong>in</strong> Anopheles gambiae. Science 298(5591): 176-178.<br />

H<strong>of</strong>mann, K. and St<strong>of</strong>fel, W. (1993). Mbase - A database <strong>of</strong> membrane spann<strong>in</strong>g<br />

prote<strong>in</strong>s segments. Biological Chemistry Hoppe-Seyler 374: 166–170.<br />

Honda, K. (1995). Chemical basis <strong>of</strong> differential oviposition by Lepidopterous<br />

<strong>in</strong>sects. Arch Insect Biochem Physiol 30: 1-23.<br />

Howse, P. E. (1998). Pheromones and Behaviour. Insect Pheromones and their Use <strong>in</strong><br />

Pest Management. P. E. Howse, I. D. R. Stevens and O. T. Jones. London,<br />

Chapman & Hall.


References 155<br />

Hrdy, I., F. Kuldova, F. Kocourek, J. Berankova and Pultar, O. (1996). Insect viruses<br />

and juvenoids dissem<strong>in</strong>ated by means <strong>of</strong> pheromone traps: a potential tool <strong>of</strong><br />

<strong>in</strong>tegrated pest management <strong>in</strong> orchards. XX International Congress <strong>of</strong><br />

Entomology,. Firenze.<br />

Illum<strong>in</strong>a. (2010). Mate pair sequenc<strong>in</strong>g assay. Retrieved 8 November 2010.<br />

http://www.illum<strong>in</strong>a.com/technology/mate_pair_sequenc<strong>in</strong>g_assay.ilmn.<br />

Jackson, D. M., Brown, G. C., Nord<strong>in</strong>, G. L. and Johnson, D. W. (1992).<br />

Autodissem<strong>in</strong>ation <strong>of</strong> a baculovirus for management <strong>of</strong> tobacco budworms<br />

(Lepidoptera: Noctuidea) on tobacco. J. Econ. Entomol 85: 710-719.<br />

James, D. G. and Price, T. S. (2004). Field-Test<strong>in</strong>g <strong>of</strong> Methyl Salicylate for<br />

Recruitment and Retention <strong>of</strong> Beneficial <strong>Insects</strong> <strong>in</strong> Grapes and Hops. Journal<br />

<strong>of</strong> Chemical Ecology 30(8): 1613-1628.<br />

Janssen, B. J., Thodey, K., Schaffer, R. J., Alba, R., Balakrishnan, L., Bishop, R.,<br />

Bowen, J. H., Crowhurst, R. N., Gleave, A. P., Ledger, S., McArtney, S.,<br />

Pichler, F. B., Snowden, K. C. and Ward, S. (2008). Global gene expression<br />

analysis <strong>of</strong> apple fruit development from the floral bud to ripe fruit. BMC<br />

Plant Biol.: 8-16.<br />

Jeck, W. R. (2007). L<strong>in</strong>ks extend<strong>in</strong>g assembly <strong>of</strong> short DNA sequences to handle<br />

error. Bio<strong>in</strong>formatics 23: 2942-2944.<br />

Jones, O. T. (1998). Practical Applications <strong>of</strong> Pheromones and Other Semiochemicals.<br />

Insect Pheromones and their Use <strong>in</strong> Pest Management. P. E. Howse, I. D. R.<br />

Stevens and O. T. Jones. London, Chapman & Hall.<br />

Jönsson, M. and Anderson, P. (1999). Electrophysiological response to herbivore-<br />

<strong>in</strong>duced host plant volatiles <strong>in</strong> the moth Spodoptera littoralis. 24(4): 377-385.


References 156<br />

Jordan, M. D., Anderson, A., Begum, D., Carraher, C., Authier, A., Marshall, S. D.<br />

G., Kiely, A., Gatehouse, L. N., Greenwood, D. R., Christie, D. L., Kralicek,<br />

A. V., Trowell, S. C. and Newcomb, R. D. (2009). Odorant Receptors from the<br />

Light brown Apple Moth (Epiphyas postvittana) Recognize Important Volatile<br />

Compounds Produced by Plants. Chem. Senses 34(5): 383-394.<br />

Jordan, M. D., Stanley, D., Marshall, S. D. G., Silva, D. D., Crowhurst, R. N., Gleave,<br />

A. P., Greenwood, D. R. and Newcomb, R. D. (2008). Expressed sequence<br />

tags and proteomics <strong>of</strong> antennae from the tortricid moth, Epiphyas postvittana.<br />

Insect Molecular Biology 17(4): 361-373.<br />

Jordon, M. D. (2006). An <strong>in</strong>vestigation <strong>in</strong>to the ultrastructure and olfactory prote<strong>in</strong>s <strong>of</strong><br />

antennae from the light brown apple moth (Epiphyas postvittana). School <strong>of</strong><br />

Biological Sciences. Auckland, University <strong>of</strong> Auckland. Doctor <strong>of</strong><br />

Philosophy.<br />

Jurenka, R. (2003). Biochemistry <strong>of</strong> female moth sex pheromones. Insect Pheromone<br />

Biochemistry and Molecular Biology. G. J. Blomquist and R. G. Vogt.<br />

London, Elsevier.<br />

Juretic, D., Lee, B., Tr<strong>in</strong>aistic, N. and Williams, R. W. (1993). Conformational<br />

preference functions for predict<strong>in</strong>g helices <strong>in</strong> membrane prote<strong>in</strong>s. Biopolymers<br />

33: 255–273.<br />

Kahsay, R., Liao, L. and Gao, G. (2005). An Improved Hidden Markov Model for<br />

Transmembrane Prote<strong>in</strong> Topology Prediction and Its Applications to Complete<br />

Genomes. Bio<strong>in</strong>formatics 21: 853-1858.<br />

Kaissl<strong>in</strong>g, K.-E. (2009). Olfactory perireceptor and receptor events <strong>in</strong> moths: a k<strong>in</strong>etic<br />

model revised. Journal <strong>of</strong> Comparative Physiology A: Neuroethology,<br />

Sensory, Neural, and Behavioral Physiology 195(10): 895-922.<br />

Kaissl<strong>in</strong>g, K. E. (1971). Insect olfaction. Handbook <strong>of</strong> Sensory Physiology: Chemical<br />

Senses 1 <strong>Olfaction</strong>. . L. M. Beidler. Berl<strong>in</strong>, Spr<strong>in</strong>ger-Verlag.


References 157<br />

Kaissl<strong>in</strong>g, K. E. (1974). Sensory transduction <strong>in</strong> <strong>in</strong>sect olfactory receptors.<br />

Biochemistry <strong>of</strong> Sensory Functions. L. Jaenicke, Spr<strong>in</strong>ger-Verlag.<br />

Kaissl<strong>in</strong>g, K. E. (1998). Pheromone deactivation catalyzed by receptor molecules: a<br />

quantitative k<strong>in</strong>etic model. Chem. Senses 23: 385-395.<br />

Kaissl<strong>in</strong>g, K. E. and Kasang, G. (1978). New pheromone <strong>of</strong> silkworm moth Bombyx<br />

mori – sensory pathway and behavioral effect. Naturwissenschaften 65: 382-<br />

384.<br />

Karg, G., Suckl<strong>in</strong>g, D. M., Rumbo, E. R. and Popay, A. J. (1992). Measurement <strong>of</strong> sex<br />

pheromone <strong>in</strong> orchard air us<strong>in</strong>g electroantennogram apparatus. Proceed<strong>in</strong>gs<br />

<strong>of</strong> the Forty Fifth New Zealand Plant Protection Conference, Well<strong>in</strong>gton, New<br />

Zealand, New Zealand Plant Protection Society; Rotorua; New Zealand.<br />

Kasang, G. (1971). Bombykol reception and metabolism on the antennae <strong>of</strong> the<br />

silkmoth Bombyx mori. Gustation and olfaction. Ohl<strong>of</strong>f G and T. AF. London,<br />

Academic Press.<br />

Kasang, G. (1973). Physikochemische Vorga¨nge beim Riechen des Seidensp<strong>in</strong>ners.<br />

Naturwissenschaften 60: 95–101.<br />

Kasang, G. and Kaissl<strong>in</strong>g, K. E. (1972). Specificity <strong>of</strong> primary and secondary<br />

olfactory processes <strong>in</strong> Bombyx antennae. <strong>Olfaction</strong> and Taste IV: Proceed<strong>in</strong>gs<br />

<strong>of</strong> the Fourth International Symposium, Wissenschaftliche Verlagsgesellschaft<br />

mbH, Stuttgart.<br />

Kasang, G., Nicholls, M., Keil, T. and Kanaujia, S. (1989a). Enzymatic conversion <strong>of</strong><br />

sex pheromones <strong>in</strong> olfactory hairs <strong>of</strong> the male silkworm moth Antheraea<br />

polyphemus. Z Naturforsch 44c: 920– 926.<br />

Kasang, G., Nicholls, M. and von Pr<strong>of</strong>f, L. (1989b). Sex pheromone conversion and<br />

degradation <strong>in</strong> antennae <strong>of</strong> the male silkworm moth Bombyx mori L.<br />

Experientia 45: 81-87.


References 158<br />

Kasang, G., von Pr<strong>of</strong>f, L. and Nicholls, M. (1988). Enzymatic conversion and<br />

degradation <strong>of</strong> sex pheromones <strong>in</strong> antennae <strong>of</strong> the male silkworm moth<br />

Antheraea polyphemus. Z Naturforsch 43c: 275–284.<br />

Keil, T. (1999). Morphology and development <strong>of</strong> the peripheral olfactory organs.<br />

Insect <strong>Olfaction</strong>. B. Hansson, Spr<strong>in</strong>ger.<br />

Kiely, A. (2008). Functional and Structural Analyses <strong>of</strong> an Olfactory Receptor from<br />

Drosophila melanogaster. School <strong>of</strong> Biological Sciences. Auckland,<br />

University <strong>of</strong> Auckland. Doctor <strong>of</strong> Philosophy.<br />

Kiely, A., Authier, A., Kralicek, A. V., Warr, C. G. and Newcomb, R. D. (2007).<br />

Functional analysis <strong>of</strong> a Drosophila melanogaster olfactory receptor expressed<br />

<strong>in</strong> Sf9 cells. Journal <strong>of</strong> Neuroscience Methods 159(2): 189-194.<br />

Kim, M. S., Repp, A. and Smith, D. P. (1998). LUSH odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

mediates chemosensory responses to alcohols <strong>in</strong> Drosophila melanogaster.<br />

Genetics 150: 711-721.<br />

Kim, U. J., Birren, B. W., Slepak, T., Manc<strong>in</strong>o, V., Boysen, C., Kang, H. L., Simon,<br />

M. I. and Shizuya, H. (1996). Construction and characterization <strong>of</strong> a human<br />

bacterial artificial chromosome library. Genomics 34(2): 213-218.<br />

Kle<strong>in</strong>, U. (1987). Sensillum-lymph prote<strong>in</strong>s from antennal olfactory hairs <strong>of</strong> the moth<br />

Antheraea polyphemus (Saturniidae). Insect Biochemistry 17: 1193-1204.<br />

Krieger, J., Gänle, H., Ram<strong>in</strong>g, K. and Breer, H. (1993). Odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>of</strong><br />

Heliothis virescens. Insect Biochemistry and Molecular Biology 23(4): 449-<br />

456.<br />

Krieger, J., Gondesen, I., Forstner, M., Gohl, T., Dewer, Y. and Breer, H. (2009).<br />

HR11 and HR13 receptor-express<strong>in</strong>g neurons are housed together <strong>in</strong><br />

pheromone-responsive sensilla trichodea <strong>of</strong> male Heliothis virescens.<br />

Chemical Senses 34: 469-477.


References 159<br />

Krieger, J., Große-Wilde, E., Gohl, T. and Breer, H. (2005). Candidate pheromone<br />

receptors <strong>of</strong> the silkmoth Bombyx mori. European Journal <strong>of</strong> Neuroscience<br />

21(8): 2167-2176.<br />

Krieger, J., Große-Wilde, E., Gohl, T., Dewer, Y. M. E., Ram<strong>in</strong>g, K. and Breer, H.<br />

(2004). Genes encod<strong>in</strong>g candidate pheromone receptors <strong>in</strong> a moth (Heliothis<br />

virescens). Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United<br />

States <strong>of</strong> America 101(32): 11845-11850.<br />

Krieger, J., Kl<strong>in</strong>k, O., Mohl, C., Ram<strong>in</strong>g, K. and Breer, H. (2003). A candidate<br />

olfactory receptor subtype highly conserved across different <strong>in</strong>sect orders.<br />

Journal <strong>of</strong> Comparative Physiology A: Neuroethology, Sensory, Neural, and<br />

Behavioral Physiology 189(7): 519-526.<br />

Krieger, J., Ram<strong>in</strong>g, K., Dewer, Y. M. E., Bette, S., Conzelmann, S. and Breer, H.<br />

(2002). A divergent gene family encod<strong>in</strong>g candidate olfactory receptors <strong>of</strong> the<br />

moth Heliothis virescens. European Journal <strong>of</strong> Neuroscience 16: 619-628.<br />

Krieger, J., von Nickisch-Rosenegk, E., Mameli, M., Pelosi, P. and Breer, H. (1996).<br />

B<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s from the antennae <strong>of</strong> Bombyx mori. Insect Biochemistry and<br />

Molecular Biology 26(3): 297-307.<br />

Krogh, A., Larsson, B., von Heijne, G. and Sonhammer, E. L. L. (2001). Predict<strong>in</strong>g<br />

transmembrane prote<strong>in</strong> topology with a hidden Markov model: application to<br />

complete genomes. ournal <strong>of</strong> Molecular Biology 305: 567–580.<br />

Kurtovic, A., Widmer, A. and Dickson, B. J. (2007). A s<strong>in</strong>gle class <strong>of</strong> olfactory<br />

neurons mediates behavioural responses to a Drosophila sex pheromone.<br />

Nature 446(7135): 542-546.<br />

Kuulasmaa, T. (2000). Oligo Explorer. http://www.genel<strong>in</strong>k.com/tools/gl-oe.asp.


References 160<br />

Langmead, B., Trapnell, C., Pop, M. and Salzberg, S. L. (2009). Ultrafast and<br />

memory-efficient alignment <strong>of</strong> short DNA sequences to the human genome.<br />

Genome Biology 10(3).<br />

Larsson, M. C., Dom<strong>in</strong>gos, A. I., Jones, W. D., Chiappe, M. E., Amre<strong>in</strong>, H. and<br />

Vosshall, L. B. (2004). Or83b Encodes a Broadly Expressed Odorant Receptor<br />

Essential for Drosophila <strong>Olfaction</strong>. Neuron 43(5): 703-714.<br />

Laughl<strong>in</strong>, J. D., Ha, T. S., Jones, D. N. M. and Smith, D. P. (2008). Activation <strong>of</strong><br />

Pheromone-Sensitive Neurons Is Mediated by Conformational Activation <strong>of</strong><br />

Pheromone-B<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>. Cell 133(7): 1255-1265.<br />

Lazar, J. (2001). Elephant sex pheromone transport and recognition. Department <strong>of</strong><br />

Medic<strong>in</strong>al Chemistry. Salt Lake Country, Utah, The University <strong>of</strong> Utah.<br />

Doctor <strong>of</strong> Philosophy.<br />

Leal, W. S., Chen, A. M., Ishida, Y., Chiang, V. P., Erickson, M. L., Morgan, T. I.<br />

and Tsuruda, J. M. (2005). K<strong>in</strong>etics and molecular properties <strong>of</strong> pheromone<br />

b<strong>in</strong>d<strong>in</strong>g and release. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the<br />

United States <strong>of</strong> America 102(15): 5386-5391.<br />

Li, R., Li, Y., Kristiansen, K. and Wang, J. (2008). SOAP: short oligonucleotide<br />

alignment program. Bio<strong>in</strong>formatics 24: 713-714.<br />

Li, R., Yu, C., Li, Y., Lam, T.-W., Yiu, S.-M., Kristiansen, K. and Wang, J. (2009).<br />

SOAP2: an improved ultrafast tool for short read alignment. Bio<strong>in</strong>formatics<br />

25(15): 1966-1967.<br />

Lo, P. L., Suckl<strong>in</strong>g, D. M., Bradley, S. J., Walker, J. T. S., Shaw, P. W. and Burnip,<br />

G. M. (2000). Factors affect<strong>in</strong>g feed<strong>in</strong>g site preferences <strong>of</strong> lightbrown apple<br />

moth, Epiphyas postvittana (Lepidoptera: Tortricidae) on apple trees <strong>in</strong> New<br />

Zealand. New Zealand Journal <strong>of</strong> Crop and Horticultural Science 28: 235-<br />

243.


References 161<br />

Lopez-Gomez, R. and Gomez-Lim., M. A. (1992). A method for extraction <strong>in</strong>tact<br />

RNA from fruits rich <strong>in</strong> polysaccharides us<strong>in</strong>g ripe mango mesocarp.<br />

HortScience 27: 440–442.<br />

Lu, T., Qiu, Y. T., Wang, G., Kwon, Jae Y., Rutzler, M., Kwon, H.-W., Pitts, R. J.,<br />

van Loon, J. J. A., Takken, W., Carlson, J. R. and Zwiebel, L. J. (2007). Odor<br />

Cod<strong>in</strong>g <strong>in</strong> the Maxillary Palp <strong>of</strong> the Malaria Vector Mosquito Anopheles<br />

gambiae. Current Biology 17(18): 1533-1544.<br />

Lund<strong>in</strong>, C., Käll, L., Kreher, S. A., Kapp, K., Sonnhammer, E. L., Carlson, J. R., von<br />

Heijne, G. and Nilsson, I. (2007). Membrane topology <strong>of</strong> the Drosophila<br />

OR83b odorant receptor. FEBS Letters 581(29): 5601-5604.<br />

Maibeche-Coisne, M., Sobrio, F., Delaunay, T., Lettere, M., Dubroca, J., Jacqu<strong>in</strong>-<br />

Joly, E. and Meillour, P. N.-L. (1997). Pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>of</strong> the<br />

moth Mamestra brassicae: Specificity <strong>of</strong> ligand b<strong>in</strong>d<strong>in</strong>g. Insect Biochemistry<br />

and Molecular Biology 27(3): 213-221.<br />

Maida, R., Mameli, M., Müller, B., Krieger, J. and Ste<strong>in</strong>brecht, R. (2005). The<br />

expression pattern <strong>of</strong> four odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>in</strong> male and female silk<br />

moths, &lt;i&gt;Bombyx mori&lt;/i&gt. Journal <strong>of</strong> Neurocytology 34(1): 149-<br />

163.<br />

Malnic, B., Godfrey, P. A. and Buck, L. B. (2004). The human olfactory receptor<br />

gene family. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United<br />

States <strong>of</strong> America 101: 2584-2589.<br />

Malnic, B., Hirono, J., Sato, T. and Buck, L. B. (1999). Comb<strong>in</strong>atorial receptor codes<br />

for odors. Cell 96: 713-723.<br />

Margulies, M. (2005). Genome sequenc<strong>in</strong>g <strong>in</strong> micr<strong>of</strong>abricated high-density picolitre<br />

reactors. Nature 437: 376-380.


References 162<br />

Markwick, N., Graves, S., Fairbairn, F. M., Docherty, L. C., Poulton, J. and Ward, V.<br />

K. (2002). Infectivity <strong>of</strong> Epiphyas postvittana nucleopolyhedrovirus for New<br />

Zealand leafrollers. Biological Control 25(2): 207-214.<br />

Masante-Roca, I., Gadenne, C. and Anton, S. (2005). Three-dimensional antennal<br />

lobe atlas <strong>of</strong> male and female moths, Lobesia botrana (Lepidoptera:<br />

Tortricidae) and glomerular representation <strong>of</strong> plant volatiles <strong>in</strong> females. J Exp<br />

Biol 208(6): 1147-1159.<br />

McKenna, M. P., Hekmat-Scafe, D. S., Ga<strong>in</strong>es, P. and Carlson, J. R. (1994). Putative<br />

Drosophila pheromone-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s expressed <strong>in</strong> a subregion <strong>of</strong> the<br />

olfactory system. Journal <strong>of</strong> Biological Chemistry 269: 16340-16347.<br />

McKernan, K., Blanchard, A., Kotler, L. and Costa, G. (2006). Reagents, methods,<br />

and libraries for bead-based sequenc<strong>in</strong>g, Nature Publish<strong>in</strong>g Group.<br />

McLaren, G. F., Fraser, J. A. and Suckl<strong>in</strong>g, D. M. (1998). Mat<strong>in</strong>g disruption for the<br />

control <strong>of</strong> leafrollers on apricots. New Zealand Journal <strong>of</strong> Crop and<br />

Horticultural Science 26: 259-268.<br />

Melo, A. C., Rutzler, M., Pitts, R. J. and Zwiebel, L. J. (2004). Identification <strong>of</strong> a<br />

chemosensory receptor from the yellow fever mosquito, Aedes aegypti, that is<br />

highly conserved and expressed <strong>in</strong> olfactory and gustatory organs. Chemical<br />

Senses 29: 403-410.<br />

Meng, L. Z., Wu, C. H., Wickle<strong>in</strong>, M., Kaissl<strong>in</strong>g, K. E. and Bestmann, H. J. (1989).<br />

Number and sensitivity <strong>of</strong> three types <strong>of</strong> pheromone receptor cells <strong>in</strong><br />

Antheraea pernyi and A. polyphemus. J Comp Physiol A 165: 139–146.<br />

Mitsuno, H., Sakurai, T., Murai, M., Yasuda, T., Kugimiya, S., Ozawa, R., Toyohara,<br />

H., Takabayashi, J., Miyoshi, H. and Nishioka, T. (2008). Identification <strong>of</strong><br />

receptors <strong>of</strong> ma<strong>in</strong> sex-pheromone components <strong>of</strong> three Lepidopteran species.<br />

European Journal <strong>of</strong> Neuroscience 28(5): 893-902.


References 163<br />

Miura, N., Nakagawa, T., Tatsuki, S., Touhara, K. and Ishikawa, Y. (2009). A male-<br />

specific odorant receptor conserved through the evolution <strong>of</strong> sex pheromones<br />

<strong>in</strong> Ostr<strong>in</strong>ia moth species. Int J Biol Sci 5(4): 319 - 330.<br />

Mohanty, S., Zubkov, S. and Campos-Olivas, R. (2003). Letter to the Editor: 1H, 13C<br />

and 15N backbone assignments <strong>of</strong> the pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> from the<br />

silk moth Antheraea polyphemus (ApolPBP). Journal <strong>of</strong> Biomolecular NMR<br />

27(4): 393-394.<br />

Mohanty, S., Zubkov, S. and Gronenborn, A. M. (2004). The Solution NMR Structure<br />

<strong>of</strong> Antheraea polyphemus PBP Provides New Insight <strong>in</strong>to Pheromone<br />

Recognition by Pheromone-b<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>s. 337(2): 443-451.<br />

Motulsky, H. and Christopoulos, A. (2003). Fitt<strong>in</strong>g Models to Biological Data us<strong>in</strong>g<br />

L<strong>in</strong>ear and Nonl<strong>in</strong>ear Regression. A practical guide to curve fitt<strong>in</strong>g. San Diego<br />

CA, GraphPad S<strong>of</strong>tware Inc.<br />

Nagy, B. A. L. and George, J. A. (1981). Differences <strong>in</strong> the numbers <strong>of</strong> sensilla<br />

trichodea between reared and wild adults <strong>of</strong> the Oriental fruit moth, Grapholitha<br />

molesta (Lepidoptera: Tortricidae). Proc. Entomol. Soc. Ont. 112: 67-72.<br />

Nakagawa, T., Sakurai, T., Nishioka, T. and Touhara, K. (2005). Insect Sex-<br />

Pheromone Signals Mediated by Specific Comb<strong>in</strong>ations <strong>of</strong> Olfactory<br />

Receptors. Science 307(5715): 1638-1642.<br />

Neuhaus, E., Gisselmann, G., Zhang, W., Dooley, R., Stortkuhl, K. and Hatt, H.<br />

(2004). Odorant receptor heterodimerization <strong>in</strong> the olfactory system <strong>of</strong><br />

Drosophila melanogaster. Nat Neurosci 8: 15 - 17.


References 164<br />

Neuh<strong>of</strong>f, V., Arold, N., Taube, D. and Ehrhardt, W. (1988). Improved sta<strong>in</strong><strong>in</strong>g <strong>of</strong><br />

prote<strong>in</strong>s <strong>in</strong> polyacrylamide gels <strong>in</strong>clud<strong>in</strong>g isoelectric focus<strong>in</strong>g gels with clear<br />

background at nanogram sensitivity us<strong>in</strong>g Coomassie brilliant blue G-250 and<br />

R-250. Electrophoresis 9: 255–262.<br />

Newcomb, R. D., Sirey, T. M., Rassam, M. and Greenwood, D. R. (2002). Pheromone<br />

b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>of</strong> Epiphyas postvittana (Lepidoptera: Tortricidae) are<br />

encoded at a s<strong>in</strong>gle locus. Insect Biochemistry and Molecular Biology 32(11):<br />

1543-1554.<br />

O'Neil, M. J. (2006). The Merck <strong>in</strong>dex : an encyclopedia <strong>of</strong> chemicals, drugs, and<br />

biologicals. M. J. O‟Neil, P. E. Heckelman, C. B. Koch, K. J. Roman, C. M.<br />

Kenny and M. R. D'Arecca, Whitehouse Station, NJ : Merck.<br />

O‟Callaghan, M. and Jackson., T. A. (1993). Adult grass grub dispersal <strong>of</strong> Serratia<br />

entomophila. Proc. 46th NZ Plant Protection Conf: 235-236.<br />

Oldham, N. J., Krieger, J., Breer, H., Fischedick, A., Hoskovec, M. and Svatoš, A.<br />

(2000). Analysis <strong>of</strong> the Silkworm Moth Pheromone B<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>–<br />

Pheromone Complex by Electrospray-Ionization Mass Spectrometry.<br />

Angewandte Chemie International Edition 39(23): 4341-4343.<br />

Oldham, N. J., Krieger, J., Breer, H. and Svatos, A. (2001). Detection and Removal <strong>of</strong><br />

an Artefact Fatty Acid from the B<strong>in</strong>d<strong>in</strong>g Site <strong>of</strong> Recomb<strong>in</strong>ant Bombyx mori<br />

Pheromone-b<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>. Chem. Senses 26(5): 529-531.<br />

Olson, M., Hood, L., Cantor, C. and Botste<strong>in</strong>, D. (1989). A common language for<br />

physical mapp<strong>in</strong>g <strong>of</strong> the human genome. Science 245: 1434–1435.<br />

Patch, H. M., Velarde, R. A., Walden, K. K. O. and Robertson, H. M. (2009). A<br />

Candidate Pheromone Receptor and Two Odorant Receptors <strong>of</strong> the Hawkmoth<br />

Manduca sexta. Chem. Senses 34(4): 305-316.


References 165<br />

Pell, J. K., Macaulay, E. D. M. and Wild<strong>in</strong>g, N. (1993). A pheromone trap for<br />

dispersal <strong>of</strong> the pathogen Zoophthora radicans Brefeld (Entomophthorales)<br />

amongst populations <strong>of</strong> the diamondback moth Plutella xylostella L.<br />

(Lepidoptera: Yponomeutidae). Biocontrol Sci. Tech. 3: 315-320.<br />

Pfaffl, M. W. (2001). A new mathematical model for relative quantification <strong>in</strong> real-<br />

time RT-PCR. Nucleic Acids Research 29: e45.<br />

Pikielny, C. W., Hasan, G., Rouyer, F. and Rosbach, M. (1994). Members <strong>of</strong> a family<br />

<strong>of</strong> Drosophila putative odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s are expressed <strong>in</strong> different<br />

subsets <strong>of</strong> olfactory hairs. Neuron 12(35-49).<br />

Poly, F., Read, T., Tribble, D. R., Baqar, S., Lorenzo, M. and Guerry, P. (2007).<br />

Genome Sequence <strong>of</strong> a Cl<strong>in</strong>ical Isolate <strong>of</strong> Campylobacter jejuni from<br />

Thailand. Infect. Immun. 75(7): 3425-3433.<br />

Pop, M. and Salzberg, S. L. (2008). Bio<strong>in</strong>formatics challenges <strong>of</strong> new sequenc<strong>in</strong>g<br />

technology. Trends <strong>in</strong> Genetics 24(3): 142-149.<br />

Poph<strong>of</strong>, B. (1997). Olfactory responses recorded from sensilla coeloconica <strong>of</strong> the<br />

silkmoth Bombyx mori. Physiological Entomology 22: 239-248.<br />

Poph<strong>of</strong>, B. (2004). Pheromone-b<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>s Contribute to the Activation <strong>of</strong><br />

Olfactory Receptor Neurons <strong>in</strong> the Silkmoths Antheraea polyphemus and<br />

Bombyx mori. Chemical Senses 29(2): 117-125.<br />

Poph<strong>of</strong>, B., Stange, G. and Abrell, L. (2005). Volatile organic compounds as signals<br />

<strong>in</strong> a plant-herbivore system: electrophysiological responses <strong>in</strong> olfactory<br />

sensilla <strong>of</strong> the moth Cactoblastis cactorum. Chemical Senses 30: 51-68.<br />

Prestwich, G. D., Du, G. and LaForest, S. (1995). How is pheromone specificity<br />

encoded <strong>in</strong> prote<strong>in</strong>s? Chemical Senses 20: 461-469.


References 166<br />

Prestwich, G. D., Graham, S. M., Handley, M., Latli, B., Stre<strong>in</strong>z, L. and Tasayco, M.<br />

L. J. (1989). Enzymatic process<strong>in</strong>g <strong>of</strong> pheromones and pheromone analogs.<br />

Cellular and Molecular Life Sciences 45(3): 263-270.<br />

Ramakers, C., Ruijter, J. M., Deprez, R. H. L. and Moorman, A. F. M. (2003).<br />

Assumption-free analysis <strong>of</strong> quantitative real-time polymerase cha<strong>in</strong> reaction<br />

(PCR) data. Neuroscience Letters 339: 62–66.<br />

Roel<strong>of</strong>s, W. L., Tette, J. P., Taschenberg, E. F. and Comeau, A. (1971). Sex<br />

pheromone <strong>of</strong> the grape berry moth: Identification by classical and<br />

electroantennogram methods, and field tests. Journal <strong>of</strong> Insect Physiology<br />

17(11): 2235-2243.<br />

Rogers, M. E., Krieger, J. and Vogt, R. G. (2001a). Antennal SNMPs (sensory neuron<br />

membrane prote<strong>in</strong>s) <strong>of</strong> Lepidoptera def<strong>in</strong>e a unique family <strong>of</strong> <strong>in</strong>vertebrate<br />

CD36-like prote<strong>in</strong>s. Journal <strong>of</strong> Neurobiology 49: 47-61.<br />

Rogers, M. E., Ste<strong>in</strong>brecht, R. A. and Vogt, R. G. (2001). Expression <strong>of</strong> SNMP-1 <strong>in</strong><br />

olfactory neurons and sensilla <strong>of</strong> male and female antennae <strong>of</strong> the silkmoth<br />

Antheraea polyphemus. Cell and Tissue Research 303(3): 433-446.<br />

Rogers, M. E., Sun, M., Lerner, M. R. and Vogt, R. G. (1997). Snmp-1, a novel<br />

membrane prote<strong>in</strong> <strong>of</strong> olfactory neurons <strong>of</strong> the silk moth Antheraea<br />

polyphemus with homology to the CD36 family <strong>of</strong> membrane prote<strong>in</strong>s.<br />

Journal <strong>of</strong> Biological Chemistry 272: 14792-14799.<br />

Ronaghi, M., Karamohamed, S., Pettersson, B., Uhlen, M. and Nyren, P. (1996).<br />

Real-time DNA sequenc<strong>in</strong>g us<strong>in</strong>g detection <strong>of</strong> pyrophosphate release. Anal.<br />

Biochem. 242: 84-89.


References 167<br />

Røstelien, T., Stranden, M., Borg-Karlson, A.-K. and Mustaparta, H. (2005).<br />

Olfactory Receptor Neurons <strong>in</strong> Two Helioth<strong>in</strong>e Moth Species Respond<strong>in</strong>g<br />

Selectively to Aliphatic Green Leaf Volatiles, Aromatic Compounds,<br />

Monoterpenes and Sesquiterpenes <strong>of</strong> Plant Orig<strong>in</strong>. Chemical Senses 30(5):<br />

443-461.<br />

Rumbo, E. R. (1981). Study <strong>of</strong> s<strong>in</strong>gle sensillum responses to pheromone <strong>in</strong> the light-<br />

brown apple moth, Epiphyas postvittana, us<strong>in</strong>g an averag<strong>in</strong>g technique.<br />

Physiological Entomology 6: 87-98.<br />

Rumbo, E. R. (1983). Differences between s<strong>in</strong>gle cell responses to different<br />

components <strong>of</strong> the sex pheromone <strong>in</strong> males <strong>of</strong> the lightbrown apple moth<br />

(Epiphyas postvittana) [Lepidoptera, Tortricidae]. Physiological Entomology<br />

8: 195-201.<br />

Sakurai, T., Nakagawa, T., Mitsuno, H., Mori, H., Endo, Y., Tanoue, S., Yasukochi,<br />

Y., Touhara, K. and Nishioka, T. (2004). Identification and functional<br />

characterization <strong>of</strong> a sex pheromone receptor <strong>in</strong> the silkmoth Bombyx mori.<br />

Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United States <strong>of</strong><br />

America 101(47): 16653-16658.<br />

Sambrook, J., Fritsch, E. F. and Maniatis, T. (1989). Molecular Clon<strong>in</strong>g: a Laboratory<br />

Manual. C. S. Harbor. NY, Cold Spr<strong>in</strong>g Harbor Laboratory.<br />

Sato, K., Pellegr<strong>in</strong>o, M., Nakagawa, T., Nakagawa, T., Vosshall, L. B. and Touhara,<br />

K. (2008). Insect olfactory receptors are heteromeric ligand-gated ion<br />

channels. Nature 452(7190): 1002-1006.<br />

Schneider, D. (1964). Insect Antennae. Annual Review <strong>of</strong> Entomology 9: 103-122.<br />

Seabrook, W. D. (1978). Neurobiological Contributions to Understand<strong>in</strong>g Insect<br />

Pheromone Systems. Annual Review <strong>of</strong> Entomology 23: 471-485.


References 168<br />

Shendure, J. (2005). Accurate multiplex polony sequenc<strong>in</strong>g <strong>of</strong> an evolved bacterial<br />

genome. Science 309: 1728-1732.<br />

Shields, V. D. C. and Hildebrand, J. G. (2001). Recent advances <strong>in</strong> <strong>in</strong>sect olfaction,<br />

specifically regard<strong>in</strong>g the morphology and sensory physiology <strong>of</strong> antennal<br />

sensilla <strong>of</strong> the female sph<strong>in</strong>x moth Manduca sexta. Microscopy Research and<br />

Technique 55: 307-329.<br />

Smart, R., Kiely, A., Beale, M., Vargas, E., Carraher, C., Kralicek, A. V., Christie, D.<br />

L., Chen, C., Newcomb, R. D. and Warr, C. G. (2008). Drosophila odorant<br />

receptors are novel seven transmembrane doma<strong>in</strong> prote<strong>in</strong>s that can signal<br />

<strong>in</strong>dependently <strong>of</strong> heterotrimeric G prote<strong>in</strong>s. 38(8): 770-780.<br />

Ste<strong>in</strong>brecht, R. A. (1996). Are odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>in</strong>volved <strong>in</strong> odorant<br />

discrim<strong>in</strong>ation? . Chem. Senses 21: 718-725.<br />

Ste<strong>in</strong>brecht, R. A. (1999). Olfactory receptors. Atlas <strong>of</strong> Anthropod Sensory Receptors.<br />

E. Eguchi and Y. Tom<strong>in</strong>aga. Tokyo, Spr<strong>in</strong>ger-Verlag.<br />

Ste<strong>in</strong>brecht, R. A., Laue, M. and Ziegelberger, G. (1995). Immunolocalization <strong>of</strong><br />

pheromone-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> and general odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> <strong>in</strong> olfactory<br />

sensilla <strong>of</strong> the silk moths Antheraea and Bombyx. Cell and Tissue Research<br />

282: 203-217.<br />

Ste<strong>in</strong>brecht, R. A., Ozaki, M. and Ziegelberger, G. (1992). Immunocytochemical<br />

localization <strong>of</strong> pheromone-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> <strong>in</strong> moth antennae. Cell and Tissue<br />

Research 270: 287-302.<br />

Sterky, F. and Lundeberg, J. (2000). Sequence analysis <strong>of</strong> genes and genomes.<br />

Journal <strong>of</strong> Biotechnology 76(1): 1-31.<br />

Stortkuhl, K. F. and Kettler, R. (2001). Functional analysis <strong>of</strong> an olfactory receptor <strong>in</strong><br />

Drosophila melanogaster. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences<br />

<strong>of</strong> the United States <strong>of</strong> America 98(16): 9381-9385.


References 169<br />

Stowers, L. and Logan, D. W. (2008). LUSH Shapes Up for a Starr<strong>in</strong>g Role <strong>in</strong><br />

<strong>Olfaction</strong>. Cell 133(7): 1137-1139.<br />

Suckl<strong>in</strong>g, D. M. and Brockerh<strong>of</strong>f, E. G. (2010). Invasion Biology, Ecology, and<br />

Management <strong>of</strong> the Light Brown Apple Moth (Tortricidae). Annual Review <strong>of</strong><br />

Entomology 55(1): 285-306.<br />

Suckl<strong>in</strong>g, D. M., Burnip, G. M., Clearwater, J. R., Shaw, P. W. M., McLaren, G. F.,<br />

Thomson, D. and Wear<strong>in</strong>g, C. H. (1994). Mat<strong>in</strong>g disruption for caterpillar<br />

control <strong>in</strong> organic pipfruit orchards <strong>in</strong> New Zealand, The Horticultural and<br />

Food Research Intitute <strong>of</strong> New Zealand.<br />

Suckl<strong>in</strong>g, D. M., Burnip, G. M., Walker, J. T. S., Shaw, P. W., McLaren, G. F.,<br />

Howard, C. R., Lo, P., White, V., and Fraser, J. (1998). Abundance <strong>of</strong><br />

leafrollers and their parasitoids on selected host plants <strong>in</strong> New Zealand. New<br />

Zealand Journal <strong>of</strong> Crop and Horticultural Science 26: 193 - 203.<br />

Suckl<strong>in</strong>g, D. M. and Karg, G. (1999). Pheromones and Other Semiochemicals.<br />

Biological and Biotechnological Control <strong>of</strong> Insect Pests. J. E. Rechcigl and N.<br />

A. Rechcigl, Lewis Publishers.<br />

Suckl<strong>in</strong>g, D. M., Karg, G. and Bradley, S. J. (1996). Apple foliage enhances mat<strong>in</strong>g<br />

disruption <strong>of</strong> light-brown apple moth. Journal <strong>of</strong> Chemical Ecology 22(2):<br />

325-341.<br />

Suckl<strong>in</strong>g, D. M., Karg, G., Gibb, A. R. and Bradley, S. J. (1996). Electroantennogram<br />

and oviposition responses <strong>of</strong> Epiphyas postvittana (Lepidoptera: Tortricidae)<br />

to plant volatiles. New Zealand Journal <strong>of</strong> Crop and Horticultural Science 24:<br />

323 - 333.<br />

Suckl<strong>in</strong>g, D. M. and Shaw, P. W. (1990). Prelim<strong>in</strong>ary trials <strong>of</strong> mat<strong>in</strong>g disruption <strong>of</strong><br />

lightbrown apple moth <strong>in</strong> Nelson Proceed<strong>in</strong>gs <strong>of</strong> the Forty Third New Zealand<br />

Weed and Pest Control Conference. Palmerston North; New Zealand, New<br />

Zealand Weed and Pest Control Society Inc: 311 316; 316 ref.


References 170<br />

Suckl<strong>in</strong>g, D. M. and Shaw, P. W. (1995). Large-scale trials <strong>of</strong> mat<strong>in</strong>g disruption <strong>of</strong><br />

lightbrown apple moth <strong>in</strong> Nelson, New Zealand. New Zealand Journal <strong>of</strong> Crop<br />

and Horticultural Science 23: 127-137.<br />

Suckl<strong>in</strong>g, D. M., Walker, J. T. S. and Wear<strong>in</strong>g, C. H. (1999). Ecological impact <strong>of</strong><br />

three pest management systems <strong>in</strong> New Zealand apple orchards. Agriculture,<br />

Ecosystems & Environment 73(2): 129-140.<br />

Syed, Z., Ishida, Y., Taylor, K., Kimbrell, D. A. and Leal, W. S. (2006). Pheromone<br />

reception <strong>in</strong> fruit flies express<strong>in</strong>g a moth's odorant receptor. Proceed<strong>in</strong>gs <strong>of</strong> the<br />

National Academy <strong>of</strong> Sciences 103(44): 16538-16543.<br />

Syed, Z., Kopp, A., Kimbrell, D. A. and Leal, W. S. (2010). Bombykol receptors <strong>in</strong><br />

the silkworm moth and the fruit fly. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong><br />

Sciences 107(20): 9436-9439.<br />

Tamaki, Y. (1985). Sex Pheromones. Comprehensive Insect Physiology Biochemistry<br />

and Pharmacology. G. A. Kerkut and L. I. Gilbert, Pergamon Press Ltd.<br />

Tanaka, K., Uda, Y., Ono, Y., Nakagawa, T., Suwa, M., Yamaoka, R. and Touhara,<br />

K. (2009). Highly Selective Tun<strong>in</strong>g <strong>of</strong> a Silkworm Olfactory Receptor to a<br />

Key Mulberry Leaf Volatile. Current Biology 19(11): 881-890.<br />

Touhara, K., Sengoku, S., Inaki, K., Tsuboi, A., Hirono, J., Sato, T., Sakano, H. and<br />

Haga, T. (1999). Functional identification and reconstitution <strong>of</strong> an odorant<br />

receptor <strong>in</strong> s<strong>in</strong>gle olfactory neurons. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong><br />

Sciences <strong>of</strong> the United States <strong>of</strong> America 96: 4040–4045.<br />

Turcatti, G., Romieu, A., Fedurco, M. and Tairi, A. P. (2008). A new class <strong>of</strong><br />

cleavable fluorescent nucleotides: synthesis and optimization as reversible<br />

term<strong>in</strong>ators for DNA sequenc<strong>in</strong>g by synthesis. Nucleic Acids Res. 36: e25.


References 171<br />

Turner, C. T., Davy, M. W., MacDiarmid, R. M., Plummer, K. M., Birch, N. P. and<br />

Newcomb, R. D. (2006). RNA <strong>in</strong>terference <strong>in</strong> the light brown apple moth,<br />

Epiphyas postvittana (Walker) <strong>in</strong>duced by double-stranded RNA feed<strong>in</strong>g.<br />

Insect Molecular Biology 15(3): 383-391.<br />

Tusnady, G. E. and Simon, I. (2001). The HMMTOP transmembrane topology<br />

prediction server. Bio<strong>in</strong>formatics 17(9): 849–850.<br />

Ulland, S., Ian, E., Mozuraitis, R., Borg-Karlson, A. K., Meadow, R. and Mustaparta,<br />

H. (2008). Methyl Salicylate, Identified as Primary Odorant <strong>of</strong> a Specific<br />

Receptor Neuron Type, Inhibits Oviposition by the Moth Mamestra Brassicae<br />

L. (Lepidoptera, Noctuidae). Chem. Senses 33(1): 35-46.<br />

van den Berg, M. J. and Ziegelberger, G. (1991). On the function <strong>of</strong> the pheromone<br />

b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> <strong>in</strong> the olfactory hairs <strong>of</strong> Antheraea polyphemus. Journal <strong>of</strong><br />

Insect Physiology 37(1): 79-85.<br />

Vandesompele, J., De Preter, K., Pattyn, F., Poppe, B., Roy, N., De Paepe, A. and<br />

Speleman, F. (2002). Accurate normalization <strong>of</strong> real-time quantitative RT-<br />

PCR data by geometric averag<strong>in</strong>g <strong>of</strong> multiple <strong>in</strong>ternal control genes. Genome<br />

Biology 3.<br />

Varela, N., Couton, L., Gemeno, C., Avilla, J., Rospars, J.-P. and Anton, S. (2009).<br />

Three-dimensional antennal lobe atlas <strong>of</strong> the oriental fruit moth, Cydia molesta<br />

(Busck) (Lepidoptera: Tortricidae): comparison <strong>of</strong> male and female<br />

glomerular organization. Cell and Tissue Research 337(3): 513-526.<br />

Vásquez, G. M., Fischer, P., Groz<strong>in</strong>ger, C. M. and Gould, F. (2010). Differential<br />

expression <strong>of</strong> odorant receptor genes <strong>in</strong>volved <strong>in</strong> the sexual isolation <strong>of</strong> two<br />

Heliothis moths. Insect Molecular Biology: no-no.


References 172<br />

Vetter, R. S. and Baker, T. C. (1983). Behavioral responses <strong>of</strong> male&lt;i&gt;Heliothis<br />

virescens&lt;/i&gt; <strong>in</strong> a susta<strong>in</strong>ed-flight tunnel to comb<strong>in</strong>ations <strong>of</strong> seven<br />

compounds identified from female sex pheromone glands. Journal <strong>of</strong><br />

Chemical Ecology 9(6): 747-759.<br />

Vogt, R. G. (1987). The molecular basis <strong>of</strong> pheromone reception: Its <strong>in</strong>fluence on<br />

behavior. Pheromone Biochemistry. G. D. Prestwich and G. J. Blomquist.<br />

Orlando, Fla, Academic.<br />

Vogt, R. G., Gary, B. and Richard, V. (2003). Biochemical diversity <strong>of</strong> odor<br />

detection: OBPs, ODEs and SNMPs. Insect Pheromone Biochemistry and<br />

Molecular Biology. San Diego, Academic Press: 391-445.<br />

Vogt, R. G., Kohne, A. C., Dubnau, J. T. and Prestwich, G. D. (1989). Expression <strong>of</strong><br />

pheromone b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s dur<strong>in</strong>g antennal development <strong>in</strong> the gypsy moth<br />

Lymantria dispar. J. Neurosci. 9(9): 3332-3346.<br />

Vogt, R. G., Lawrence, I. G., Kostas, I. and Sarjeet, S. G. (2005). Molecular Basis <strong>of</strong><br />

Pheromone Detection <strong>in</strong> <strong>Insects</strong>. Comprehensive Molecular Insect Science.<br />

Amsterdam, Elsevier: 753-803.<br />

Vogt, R. G., Prestwich, G. D. and Lerner, M. R. (1991). Odorant-b<strong>in</strong>d<strong>in</strong>g-prote<strong>in</strong><br />

subfamilies associate with dist<strong>in</strong>ct classes <strong>of</strong> olfactory receptor neurons <strong>in</strong><br />

<strong>in</strong>sects. Journal <strong>of</strong> Neurobiology 22(1): 74-84.<br />

Vogt, R. G., Prestwich, G. D. and Riddiford, L. M. (1988). Sex pheromone receptor<br />

prote<strong>in</strong>s. Visualization us<strong>in</strong>g a radiolabeled photoaff<strong>in</strong>ity analog. Journal <strong>of</strong><br />

Biological Chemistry 263: 3952-3959.<br />

Vogt, R. G. and Riddiford, L. M. (1981). Pheromone b<strong>in</strong>d<strong>in</strong>g and <strong>in</strong>activation by<br />

moth antennae. Nature 293(5828): 161-163.<br />

Vogt, R. G. and Riddiford, L. M. (1986). Scale esterase. Journal <strong>of</strong> Chemical Ecology<br />

12(2): 469-482.


References 173<br />

Vogt, R. G., Riddiford, L. M. and Prestwich, G. D. (1985). K<strong>in</strong>etic properties <strong>of</strong> a sex<br />

pheromone-degrad<strong>in</strong>g enzyme: the sensillar esterase <strong>of</strong> Antheraea<br />

polyphemus. Proceed<strong>in</strong>gs <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United<br />

States <strong>of</strong> America 82(24): 8827-8831.<br />

Vogt, R. G., Rogers, M. E., Franco, M. D. and Sun, M. (2002). A comparative study<br />

<strong>of</strong> odorant b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> genes: differential expression <strong>of</strong> the PBP1-GOBP2<br />

gene cluster <strong>in</strong> Manduca sexta (Lepidoptera) and the organization <strong>of</strong> OBP<br />

genes <strong>in</strong> Drosophila melanogaster (Diptera). Journal <strong>of</strong> Experimental biology<br />

205: 719–744.<br />

Vosshall, L. B., Amre<strong>in</strong>, H., Morozov, P. S., Rzhetsky, A. and Axel, R. (1999). A<br />

Spatial Map <strong>of</strong> Olfactory Receptor Expression <strong>in</strong> the Drosophila Antenna.<br />

Cell 96(5): 725-736.<br />

Vosshall, L. B., Wong, A. M. and Axel, R. (2000). An Olfactory Sensory Map <strong>in</strong> the<br />

Fly Bra<strong>in</strong>. Cell 102(2): 147-159.<br />

Wall, C. (1989). Evaluation and use <strong>of</strong> behaviour-modify<strong>in</strong>g chemicals. Insect<br />

Pheromones <strong>in</strong> Plant Protection. A. R. Jutsum and R. F. S. Gordon, Wiley-<br />

Interscience.<br />

Wang, G., Vásquez, G. M., Schal, C., Zwiebel, L. J. and Gould, F. (2010). Functional<br />

characterization <strong>of</strong> pheromone receptors <strong>in</strong> the tobacco budworm Heliothis<br />

virescens. Insect Molecular Biology.<br />

Wang, G. R., Wu, K. M. and Guo, Y. Y. (2003). Clon<strong>in</strong>g, expression and<br />

immunocytochemical localization <strong>of</strong> a general odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> gene<br />

from Helicoverpa armigera (Hübner). Insect Biochemistry and Molecular<br />

Biology 33(1): 115-124.


References 174<br />

Wanner, K. W., Anderson, A. R., Trowell, S. C., Theilmann, D. A., Robertson, H. M.<br />

and Newcomb, R. D. (2007). Female-biased expression <strong>of</strong> odourant receptor<br />

genes <strong>in</strong> the adult antennae <strong>of</strong> the silkworm, Bombyx mori. Insect Molecular<br />

Biology 16(1): 107-119.<br />

Wanner, K. W., Nichols, A. S., Allen, J. E., Bunger, P. L., Garczynski, S. F., L<strong>in</strong>n, C.<br />

E., Jr., Robertson, H. M. and Luetje, C. W. (2010). Sex Pheromone Receptor<br />

Specificity <strong>in</strong> the European Corn Borer Moth, Ostr<strong>in</strong>ia nubilalis. PLoS ONE<br />

5(1): e8685.<br />

Warren, R. L., Sutton, G. G., Jones, S. J. and Holt, R. A. (2007). Assembl<strong>in</strong>g millions<br />

<strong>of</strong> short DNA sequences us<strong>in</strong>g SSAKE. Bio<strong>in</strong>formatics 23: 500-501.<br />

Wear<strong>in</strong>g, C. H. (1993). IPM and the Consumer: The Challenge and Opportunity <strong>of</strong> the<br />

1990s. Pest Control & Susta<strong>in</strong>able Agriculture. S. Corey, D. Dall and W.<br />

Milne, CSIRO.<br />

Wear<strong>in</strong>g, C. H., Thomas, W. P., Dugdale, J. S. and Danthanarayana, W. (1991).<br />

Totricid Pests <strong>of</strong> Pome and Stone Fruits, Australian and New Zealand Species.<br />

Tortricid Pests Their Biology, Natural Enemies and Control. L. P. S. van der<br />

Geest and H. H. Evenhuis, Elsevier Science Publishers.<br />

Weber, J. L. and Myers, E. W. (1997). Human Whole-Genome Shotgun Sequenc<strong>in</strong>g.<br />

Genome Research 7(5): 401-409.<br />

Wetzel, C. H., Behrendt, H.-J., Gisselmann, G., Störtkuhl, K. F., Hovemann, B. and<br />

Hatt, H. (2001). Functional expression and characterization <strong>of</strong> a Drosophila<br />

odorant receptor <strong>in</strong> a heterologous cell system. Proceed<strong>in</strong>gs <strong>of</strong> the National<br />

Academy <strong>of</strong> Sciences <strong>of</strong> the United States <strong>of</strong> America 98(16): 9377-9380.<br />

Wicher, D., Schafer, R., Bauernfe<strong>in</strong>d, R., Stensmyr, M. C., Heller, R., He<strong>in</strong>emann, S.<br />

H. and Hansson, B. S. (2008). Drosophila odorant receptors are both ligand-<br />

gated and cyclic-nucleotide-activated cation channels. Nature 452(7190):<br />

1007-1011.


References 175<br />

Widmayer, P., Heifetz, Y. and Breer, H. (2009). Expression <strong>of</strong> a pheromone receptor<br />

<strong>in</strong> ovipositor sensilla <strong>of</strong> the female moth (Heliothis virescens). Insect<br />

Molecular Biology 18(4): 541-547.<br />

Xiu, W., Dong, S. and Miao, H. (2004). Molecular clon<strong>in</strong>g <strong>of</strong> Spodoptera exigua<br />

putative chemosensory receptor 2. Unpublished sequence Accession number<br />

AY862142.<br />

Xu, P., Atk<strong>in</strong>son, R., Jones, D. N. M. and Smith, D. P. (2005). Drosophila OBP<br />

LUSH Is Required for Activity <strong>of</strong> Pheromone-Sensitive Neurons. Neuron<br />

45(2): 193-200.<br />

Zhang, D.-D., Zhu, K. Y. and Wang, C.-Z. (2010). Sequenc<strong>in</strong>g and characterization <strong>of</strong><br />

six cDNAs putatively encod<strong>in</strong>g three pairs <strong>of</strong> pheromone receptors <strong>in</strong> two<br />

sibl<strong>in</strong>g species, Helicoverpa armigera and Helicoverpa assulta. Journal <strong>of</strong><br />

Insect Physiology 56(6): 586-593.<br />

Zhang, S., Maida, R. and Ste<strong>in</strong>brecht, R. A. (2001). Immunolocalization <strong>of</strong> odorant-<br />

b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong>s <strong>in</strong> noctuid moths (Insecta, Lepidoptera). Chemical Senses 26:<br />

885-896.<br />

Zhou, J.-J., Huang, W., Zhang, G.-A., Pickett, J. A. and Field, L. M. (2004). "Plus-C"<br />

odorant-b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> genes <strong>in</strong> two Drosophila species and the malaria<br />

mosquito Anopheles gambiae. Gene 327(1): 117-129.<br />

Zhou, J.-J., Robertson, G., He, X., Dufour, S., Hooper, A. M., Pickett, J. A., Keep, N.<br />

H. and Field, L. M. (2009). Characterisation <strong>of</strong> Bombyx mori Odorant-b<strong>in</strong>d<strong>in</strong>g<br />

Prote<strong>in</strong>s Reveals that a General Odorant-b<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong> Discrim<strong>in</strong>ates<br />

Between Sex Pheromone Components. Journal <strong>of</strong> Molecular Biology 389(3):<br />

529-545.


References 176<br />

Zhulidov, P., Bogdanova, E., Shcheglov, A., Shag<strong>in</strong>a, I., Vagner, L., Khazpekov, G.,<br />

Kozhemiako, V., Luk'ianov, S. and Shag<strong>in</strong>, D. (2005). A method for the<br />

preparation <strong>of</strong> normalized cDNA libraries enriched with full-length sequences.<br />

Bioorg Khim 31(12): 186-194.<br />

Zhulidov, P., Bogdanova, E., Shcheglov, A., Vagner, L., Khaspekov, G.,<br />

Kozhemyako, V., Matz, M., Meleshkevitch, E., Moroz, L., Lukyanov, S. and<br />

DA., S. (2004). Simple cDNA normalization us<strong>in</strong>g kamchatka crab duplex-<br />

specific nuclease. Nucleic Acids Res 32(3).<br />

Zubkov, S., Gronenborn, A. M., Byeon, I.-J. L. and Mohanty, S. (2005). Structural<br />

Consequences <strong>of</strong> the pH-<strong>in</strong>duced Conformational Switch <strong>in</strong> A. polyphemus<br />

Pheromone-b<strong>in</strong>d<strong>in</strong>g Prote<strong>in</strong>: <strong>Mechanisms</strong> <strong>of</strong> Ligand Release. Journal <strong>of</strong><br />

Molecular Biology 354(5): 1081-1090.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!