Characterisation of a novel antinematode agent produced by the marine epiphytic bacterium tunicata and its impact on Caenorhabditis elegans

Nor Hawani Salikin

A thesis in fulfilment of the requirements for the degree of

Doctor of Philosophy

School of Biological, Earth and Environmental Sciences

Faculty of Science

August 2020

Thesis/Dissertation Sheet

Surname/Family Name : Salikin Given Name/s : Nor Hawani Abbreviation for degree as give in the University : Ph.D. calendar Faculty : UNSW Faculty of Science School : School of Biological, Earth and Environmental Sciences Characterisation of a novel antinematode agent produced Thesis Title : by the marine epiphytic bacterium Pseudoalteromonas tunicata and its impact on Caenorhabditis elegans

Abstract 350 words maximum: (PLEASE TYPE)

Drug resistance among parasitic has resulted in an urgent need for the development of new therapies. However, the high re-discovery rate of antinematode compounds from terrestrial environments necessitates a new repository for future drug research. Marine epiphytic are hypothesised to produce nematicidal compounds as a defence against bacterivorous predators, thus representing a promising, yet underexplored source for antinematode drug discovery. The marine epiphytic bacterium Pseudoalteromonas tunicata is known to produce a number of bioactive compounds. Screening genomic libraries of P. tunicata against the Caenorhabditis elegans identified a clone (HG8) showing fast-killing activity. However, the molecular, chemical and biological properties of HG8 remain undetermined. A novel Nematode killing protein-1 (Nkp-1) encoded by an uncharacterised gene of HG8 annotated as hp1 was successfully discovered through this project. The Nkp-1 toxicity appears to be nematode-specific, with the protein being highly toxic to nematode larvae but having no impact on nematode eggs. A putative carbohydrate binding module was identified at the N-terminus of Nkp-1 protein sequence which is suggested to bind to a yet unknown nematode glycoconjugate receptor. This study also provides the first insights into the mode of action of Nkp-1 and the nematode response towards its toxicity. The Nkp-1 expressing clones; HG8 and HP1 (expressing respectively either the original 13.8 kb genomic insert of HG8 or the hp1 gene only) colonised C. elegans intestine, in addition exposure to both strains and protein extracts resulted in multiple physical damages and necrosis. As a defence, C. elegans utilised its innate and associative learned avoidance behaviour to prevent contact with the Nkp-1 strains. Further I found evidence for the involvement of daf-2/daf-16 ILR and sek-1 p38_MAPK immune pathways in response to Nkp-1 exposure and the subsequent expression of genes involved in lysozyme, superoxide dismutase production and dar (deformed anal region) formation. Moreover, this study revealed the impact of different gut microbiota has on nematode survival and the resulting physical damages upon exposure to the Nkp-1. The outcome of these studies not only kickstart the development of Nkp-1 as a future antinematode drug but has re-affirmed the potential of marine epiphytic bacteria as a new source of novel antinematode drugs.

Declaration relating to disposition of project thesis/dissertation

I hereby grant to the University of New South Wales or its agents a non-exclusive licence to archive and to make available (including to members of the public) my thesis or dissertation in whole or in part in the University libraries in all forms of media, now or here after known. I acknowledge that I retain all intellectual property rights which subsist in my thesis or dissertation, such as copyright and patent rights, subject to applicable law. I also retain the right to use all or part of my thesis or dissertation in future works (such as articles or books).

13 August 2020 ………………………………………………… ……….……………………...…….… Signature Date

The University recognises that there may be exceptional circumstances requiring restrictions on copying or conditions on use. Requests for restriction for a period of up to 2 years can be made when submitting the final copies of your thesis to the UNSW Library. Requests for a longer period of restriction may be considered in exceptional circumstances and require the approval of the Dean of Graduate Research.

i

ORIGINALITY STATEMENT

‘I hereby declare that this submission is my own work and to the best of my knowledge it contains no materials previously published or written by another person, or substantial proportions of material which have been accepted for the award of any other degree or diploma at UNSW or any other educational institution, except where due acknowledgement is made in the thesis. Any contribution made to the research by others, with whom I have worked at UNSW or elsewhere, is explicitly acknowledged in the thesis. I also declare that the intellectual content of this thesis is the product of my own work, except to the extent that assistance from others in the project's design and conception or in style, presentation and linguistic expression is acknowledged.’

Signed ……………………………………………......

13 August 2020 Date ……………………………………………......

ii

INCLUSION OF PUBLICATIONS STATEMENT

UNSW is supportive of candidates publishing their research results during their candidature as detailed in the UNSW Thesis Examination Procedure.

Publications can be used in their thesis in lieu of a Chapter if: • The candidate contributed greater than 50% of the content in the publication and is the “primary author”, ie. the candidate was responsible primarily for the planning, execution and preparation of the work for publication • The candidate has approval to include the publication in their thesis in lieu of a Chapter from their supervisor and Postgraduate Coordinator. • The publication is not subject to any obligations or contractual agreements with a third party that would constrain its inclusion in the thesis

Please indicate whether this thesis contains published material or not:

This thesis contains no publications, either published or submitted for ☒ publication

Some of the work described in this thesis has been published and it ☐ has been documented in the relevant Chapters with acknowledgement

This thesis has publications (either published or submitted for publication) incorporated into it in lieu of a chapter and the details are ☐ presented below

CANDIDATE’S DECLARATION I declare that: • I have complied with the UNSW Thesis Examination Procedure • where I have used a publication in lieu of a Chapter, the listed publication(s) below meet(s) the requirements to be included in the thesis. Candidate’s Name Signature Date (dd/mm/yy)

Nor Hawani Salikin 13 August 2020

iii

COPYRIGHT STATEMENT

‘I hereby grant the University of New South Wales or its agents a non-exclusive licence to archive and to make available (including to members of the public) my thesis or dissertation in whole or part in the University libraries in all forms of media, now or here after known. I acknowledge that I retain all intellectual property rights which subsist in my thesis or dissertation, such as copyright and patent rights, subject to applicable law. I also retain the right to use all or part of my thesis or dissertation in future works (such as articles or books).’

‘For any substantial portions of copyright material used in this thesis, written permission for use has been obtained, or the copyright material is removed from the final public version of the thesis.’

Signed ……………………………………………......

13 August 2020 Date ……………………………………………......

AUTHENTICITY STATEMENT ‘I certify that the Library deposit digital copy is a direct equivalent of the final officially approved version of my thesis.’

Signed ……………………………………………......

13 August 2020 Date ………………………………………………………………….

iv

ABSTRACT

Drug resistance among parasitic nematodes has resulted in an urgent need for the development of new therapies. However, the high re-discovery rate of antinematode compounds from terrestrial environments necessitates a new repository for future drug research. Marine epiphytic bacteria are hypothesised to produce nematicidal compounds as a defence against bacterivorous predators, thus representing a promising, yet underexplored source for antinematode drug discovery. The marine epiphytic bacterium Pseudoalteromonas tunicata is known to produce a number of bioactive compounds. Screening genomic libraries of P. tunicata against the nematode Caenorhabditis elegans identified a clone (HG8) showing fast-killing activity. However, the molecular, chemical and biological properties of HG8 remain undetermined. A novel Nematode killing protein-1 (Nkp-1) encoded by an uncharacterised gene of HG8 annotated as hp1 was successfully discovered through this project. The Nkp-1 toxicity appears to be nematode- specific, with the protein being highly toxic to nematode larvae but having no impact on nematode eggs. A putative carbohydrate binding module was identified at the N-terminus of Nkp-1 protein sequence which is suggested to bind to a yet unknown nematode glycoconjugate receptor. This study also provides the first insights into the mode of action of Nkp-1 and the nematode response towards its toxicity. The Nkp-1 expressing clones; HG8 and HP1 (expressing respectively either the original 13.8 kb genomic insert of HG8 or the hp1 gene only) colonised C. elegans intestine, in addition exposure to both strains and protein extracts resulted in multiple physical damages and necrosis. As a defence, C. elegans utilised its innate and associative learned avoidance behaviour to prevent contact with the Nkp-1 strains. Further I found evidence for the involvement of daf-2/daf-16 ILR and sek-1 p38_MAPK immune pathways in response to Nkp-1 exposure and the subsequent expression of genes involved in lysozyme, superoxide dismutase production and dar (deformed anal region) formation. Moreover, this study revealed the impact of different gut microbiota has on nematode survival and the resulting physical damages upon exposure to the Nkp-1. The outcome of these studies not only kickstart the development of Nkp-1 as a future antinematode drug but has re-affirmed the potential of marine epiphytic bacteria as a new source of novel antinematode drugs.

v

TABLE OF CONTENT

THESIS / DISSERTATION SHEET……………………………………………………..i ORIGINALITY STATEMENT …………………………………………………………ii INCLUSION OF PUBLIC STATEMENT ……………………………………………..iii COPYRIGHT & AUTHENTICITY STATEMENT ……………………………………iv ABSTRACT …………………………………………………….……...………………..v TABLE OF CONTENT ……...………………………………………………………....vi ACKNOWLEDGEMENT ………...…………………………………………………..xiii LIST OF PUBLICATION AND PRESENTATIONS …..…………………………...…xv LIST OF ABREVIATIONS AND SYMBOLS.……………………………………….xvi LIST OF FIGURES ……………………………………………………………...... xviii LIST OF TABLES ...…………………………………….…………………………...xxiv

CHAPTER 1. General Introduction and thesis aims ……………………….…....….1 1.1 What is nematode? ……………………..………………………………….…….…2 1.2 Nematodes as human parasites ……………………………..…………..…………..3 1.3 Parasitic nematodes; a significant threat to agriculture, fisheries and livestock industries ………………………………………………………………...…………4 1.4 The increasing trend of drug resistance among the parasitic nematodes; an urgent need for new antinematode treatments ………………………………………….….6 1.5 Caenorhabditis elegans as a model organism for antinematode drug discovery ……8 1.5.1 Characterising the mode of action of antinematode compound using C. elegans……………………………………………………………………11 1.5.2 Understanding nematode behavioural response and innate immunity regulation towards antinematode compounds ………..…………………. 13 1.6 C. elegans gut microbiota ………………………..………………………………. 18 1.7 Antinematode drug discovery; transition from terrestrial to marine-derived microbial compounds …..………………………………………………………... 19 1.8 Surface associated marine bacteria; a reservoir of novel antimicrobial and antinematode drugs discovery …………………………………………………… 23

vi

1.9 Pseudoalteromonas tunicata; a model of marine surface associated bacterium …………………………………………………...……………………. 30 1.10 Aims of study………………………………………..………………….………...36 1.11 REFERENCES…...……………………………………………………………….38

CHAPTER 2. Characterisation of the fast nematode-killing activity of the marine epiphytic bacterium Pseudoalteromonas tunicata D2 ……………………………….59

2.1 INTRODUCTION …………………...……………………………………………60 2.2 MATERIALS AND METHODS…………………………………………...... 62 2.2.1 Bacterial strains and culture condition ……………………...……………62 2.2.2 Maintenance and synchronisation of nematode C. elegans……………….64 2.2.3 DNA sequence analysis and annotation …………………………...……..65 2.2.4 Generation of HG8 mutant libraries ……………………...………………65 2.2.4.1 Transposon library screening ………………………………….66 2.2.5 Sequencing analysis ……………………………………...………………67 2.2.6 Cloning of P. tunicata gene inserts into the pBAD24 vector………….…67 2.2.6.1 Transformation of pBAD24 recombinant vector into EPI300 and 7C8 E. coli strains and screening against C. elegans ………….68 2.2.7 Protein extraction ……………………………...…………………………71 2.2.8 SDS PAGE, mass spectrometry and protein modelling …………………..71 2.2.9 Nematode killing assay …………………...……………………………...72 2.2.10 Preparation of heat-killed bacterial clones …………………...…………..73 2.2.11 Nematode egg hatching assay………………………...…………………..73 2.2.12 Nematode brood size assay……………………………...………………..74 2.2.13 Nematode killing assay using 24-well microtiter plates……………..…...74 2.2.14 Antibacterial and antifungal assay…………...…………………………...75 2.3 RESULTS………………...………………………………………………………. 76 2.3.1 Annotation of the 13.8 kb P. tunicata D2 DNA fragment encoded in the HG8 fosmid……………………………...…………………………………….. 76 2.3.2 Generation of transposon mutants in HG8 with attenuated nematode killing activity…………………………...………………………………………..86 2.3.3 Attenuated activity of the 7C8 mutant is restored upon complementation with hp1………………...…………………………………………………88 vii

2.3.4 Expression of hp1 gene in E. coli results in reduced C. elegans survival …89 2.3.5 Exposure to E. coli HP1 and HG8 does not affect C. elegans egg hatching efficiency but decreases the brood size ……...……………………………91 2.3.6 Heat-killing treatment significantly reduces the toxic activity of E. coli HP1 and HG8…………………………………………………………………...93 2.3.7 E. coli HP1 and HG8 bioactivities are nematode specific………..……….94 2.3.8 The antinematode compound produced by E. coli HP1 and HG8 is not secreted extracellularly……………………….………………………….. 97 2.3.9 E. coli HP1 and HG8 protein extracts are toxic against C. elegans……….98 2.3.10 SDS PAGE, mass spectrometry analysis and 3D protein modelling of the Nematode killing protein-1 (Nkp-1)………...…………………………….99 2.4 DISCUSSION……...……………………………………………………………. 103 2.4.1 The gene hp1 is responsible for the expression of Nkp-1; a novel nematode killling protein with a putative carbohydrate binding module……………103 2.4.2 Different expression system and heterologous recombinant cell background effects Nkp-1 expression level in E. coli HG8 and HP1 strains………….107 2.4.3 The products of other genes in HG8 might have an indirect effect on Nkp-1 toxicity against C. elegans………...……………………………………..108 2.4.4 E. coli HP1 and HG8 strains appear to be specifically toxic to nematodes………………………………………………………………..112 2.4.5 High temperature diminished E. coli HP1 and HG8 toxic activity against C. elegans…………………...……………………………………………... 112 2.5 CONCLUSION……………………...………………………………………...... 113 2.6 REFERENCES …………..………………………...……………………………. 114 SUPPLEMENTARY MATERIALS ………………………………………...………. 124

CHAPTER 3. Investigation of the putative mode of action and nematode response to Nkp-1……………………………………………..………………………………..132

3.1 INTRODUCTION ………………………………………………...……………..133 3.2 MATERIALS AND METHODS……………..…………………………………..135 3.2.1 Bacterial strains and culture condition ……..…………………………...135 3.2.2 Cultivation and manipulation of Caenorhabditis elegans………..……..137 3.2.3 C. elegans exposure to the liquid culture of Nkp-1 expressing clones…..137 viii

3.2.4 Tagging E. coli strains with GFP-expressing plasmid…………..……….137 3.2.5 Bacterial colonisation assay……………………………………….....…. 137 3.2.6 Assessment of C. elegans damage due to the Nkp-1 exposure…………..138 3.2.7 Determination of E. coli clones HP1 and HG8 killing mechanism against C. elegans ……………………...…………………………………………...139 3.2.7.1. Enzymatic assays…………………...……………………...….139 3.2.7.2. Necrosis assay…………………………………...…………….140 3.2.8 Microscopy imaging…………………………………...………………...140 3.2.9 C. elegans avoidance behaviour and aversive olfactory learning against the Nkp-1 and HG8 strains………………………………...………………...141 3.2.10 C. elegans RNA isolation and cDNA synthesis………………...………..143 3.2.11 Quantitative Polymerase Chain Reaction (qPCR)……………...………..144 3.3 RESULTS ………………………………………………...……………………...148 3.3.1 E. coli strains expressing Nkp-1 are able to colonise and persist in the gastrointestinal lumen of C. elegans………………………………..…...148 3.3.2 Exposure to E. coli strains HP1::GFP and HG8::GFP resulted in morphological changes in C. elegans…………………………………….151 3.3.3 Total protein fraction of E. coli strains expressing Nkp-1 resulted in morphological changes on C. elegans……………………………………156 3.3.4 E. coli HP1 and HG8 bacterial strains and their protein extracts do not show any evidence of hydrolytic enzyme activity……………………………..161 3.3.5 Exposure to E. coli strains expressing Nkp-1 results in loss of C. elegans cell membrane integrity…………………………………...………………… 162 3.3.6 C. elegans has an innate and associative learned avoidance behaviour against E. coli strains HP1 and HG8……………………………………………...165 3.3.7 Exposure to E. coli HP1 resulted in differential expression of innate immunity genes in C. elegans…………………...……………………….167 3.3.8 Exposure to E. coli HP1 downregulates the sod-3 and upregulates the lys-8 gene expression in C. elegans……………………………...…………….169 3.4 DISCUSSION……………………………………………...……………………. 170 3.4.1 Nkp-1 expressing E. coli clones kill C. elegans via a step by step mode of action (MOA)……………………………………………………………170

ix

3.4.1.1. Stage 1: Ingestion and digestion of Nkp-1 expressing bacteria by C. elegans…………………..………………………………… 170 3.4.1.2. Stage 2: Nkp-1 triggers cellular level (necrosis) damage in C. elegans………………………...……………………………… 171 3.4.1.3. Stage 3: Nkp-1 expressing bacteria cause a range of physical damages in C. elegans…………………………….……………172 3.4.1.4. Stage 4: Enhanced Nkp-1 expressing bacterial colonisation…..173 3.4.2 C. elegans response against Nkp-1 expressing bacteria……..…………..176 3.4.2.1. C. elegans avoids Nkp-1 expressing clones and shows associative learning behaviour………...…………………………………...176 3.4.2.2. C. elegans internal hatching and dar formation………………...177 3.4.2.3. The daf-2/daf-16 genes control the downstream genes sod-3 and lys-8 maybe essential for C. elegans immune defence against the Nkp-1 expressing strains………………………………………………………..... 178 3.4.2.4. sek-1 maybe important for the dar phenotype in C. elegans exposed to E. coli HP1…………………………………………………..180 3.5 CONCLUSION…………………………...……………………………………... 182 3.6 REFERENCES………...…………...…………………………………………….183 SUPPLEMENTARY MATERIALS…………………………………...…………….. 191

CHAPTER 4. The presence of a defined gut microbiota in Caenorhabditis elegans alleviates toxicity of the antinematode protein Nkp- 1………………………………………………………………………...…….……….206

4.1 INTRODUCTION……………………...………………………………………...207 4.2 MATERIALS AND METHODS……………………………...………………… 209 4.2.1 Bacterial strains and culture conditions………………………………….209 4.2.2 Maintenance and manipulation of C. elegans …………………………...211 4.2.3 Enriched soil preparation ……………………...………………………..211 4.2.4 Establishment of C. elegans N2 with an undefined gut microbiota from enriched soil …………………………………………………………….212

x

4.2.5 Isolation and identification of the cultivable C. elegans N2_UGM gut microbiota …………………………………..…………………………..214 4.2.6 Establishment of C. elegans N2 with a defined gut microbiota …………215 4.2.7 Nematode killing assay ………………………………………..………..216 4.2.8 E. coli HP1::GFP and HG8::GFP bacterial colonisation assay and microscopy imaging …………………………………………………….216 4.2.9 16S rRNA gene amplicon sequencing analysis of enriched soil and C. elegans samples ………………………………………………………...217 4.2.9.1. DNA isolation ……………………………………………...….217 4.2.9.2. 16S rRNA gene amplification and sequencing ………………..217 4.2.10 16S rRNA gene sequencing analysis ……………………………………218 4.2.11 Microbial community analysis …………………………………...……..219 4.2.11.1. Detection of sequences corresponding to individual gut bacteria and the differentially abundant bacterial communities in the N2_DGM gut microbiota ……………………………………220 4.3 RESULTS ……………………………………………………………...………...221 4.3.1 Establishment of an undefined gut microbiota (UGM) in C. elegans N2..221 4.3.2 C. elegans with an undefined gut microbiota (N2_UGM) showed a reduced survival compared to the N2 monoxenic nematodes ……………………225 4.3.3 Isolation of bacterial cultures from C. elegans N2 UGM and establishment of C. elegans N2 with a defined gut microbiota (DGM)…………………227 4.3.4 A defined gut microbiota provides resistance to E. coli HP1 compared to monoxenic nematodes …………………………………………………..232 4.3.5 Reduction of Nkp-1 expressing E. coli colonisation in C. elegans with a defined gut microbiota compared to the N2 monoxenic nematodes …….234 4.3.6 Morphology of C. elegans N2_DGM with a defined gut microbiota was less affected by E. coli HP1::GFP and HG8::GFP compared to the N2 monoxenic nematodes …………………………………………………..236 4.4 DISCUSSION…………………...………………………………………………..242 4.4.1 Establishment of potentially detrimental gut microbiota reduce C. elegans survival ………………...………………………………………………. 242 4.4.2 The high relative abundance of potentially beneficial gut bacteria may improve C. elegans survival against the toxic Nkp-1 expressing clones...243

xi

4.4.3 An increased proportion of potentially beneficial gut microbiota diminish physical damages caused by the Nkp-1 expressing strains………………245 4.5 CONCLUSION………………………………………………………………...... 247 4.6 REFERENCES...... ……………………………………………………………….248 SUPPLEMENTARY MATERIALS ………………………………………………….254

CHAPTER 5. General Discussion…………………………………………………...282

5.1 Climate change; an inevitable factor that escalates parasitic nematode infection…………………………………………………………………………..283 5.2 Blue gold from the ; a new antinematode compound from the surface associated marine bacterium Pseudoalteromonas tunicta D2 ………………………………..284 5.3 Nematode behaviour and innate immunity regulation against Nkp-1…………….287 5.4 Native nematode gut microbiota; a neglected factor in antinematode drug development ……………………………………………………………………...289 5.5 Conclusion, research limitation and future perspective ……………………………………………………………………….293 5.6 REFERENCES….………………………...……………………………………...295 APPENDIX I ………………………………………………………………………….302 APPENDIX II ………………………………………………………………………...314 APPENDIX III ……………………………………………………………………...... 318 APPENDIX IV…………………………………………………………...…………... 319

xii

ACKNOWLEDGEMENT

Completing a PhD study is an incredible adventure but yet, a strong everlasting friendship and memories have been inspired from this journey. So, as it happens in life, there were times when it was challenging. In those moments, few inspiring and wonderful people help me to believe that everything is possible. I am blessed to have them always by my side during my laugh and tears.

My first special person was my Supervisor Associate Professor Suhelen Egan. Su, you were always beside me during my good and bad. You believed in me, and were always being patient to guide me in this journey. You are a great Mentor, an incredible Woman and Scientist that I admire and will always consider as an example to follow. Thank you so much for being an extraordinary Supervisor and friend to me. I also would like to convey my highest appreciation to my Co-Supervisor Associate Professor Belinda Ferrari. Thank you so much for your comforting words and kind motivation that you gave me. Also, appreciation goes to Professor Torsten Thomas. Torsten, I always remember the time when you introduced me to Su. Thank you so much for the opportunity and your help during my PhD journey. I will never forget that definitely!

I also would like to convey a special appreciation to Helen and Associate Professor Christopher Marquis from the UNSW Recombinant Products Facility for providing precious guidance and equipment for my study. A hearty thanks also goes to the BABS and BEES people and also for Anneli from the GRS who were always being nice and supportive. Thank you so much for your kind help and I will never forget you!

Utmost appreciation is also extended to these wonderful people; Jadi, Marwan, Jen and Priscila. Guys, you were always there next to me, always supported me and helped me during my PhD journey. I am blessed to have you as my friends and I hope that this friendship will be everlasting. To Mika, Willis, Shan, David, Jess, Syukur, Irene, Eve, Ale, Alex, Hung, Vicky, Mansura, Tahsin, Zillur, Jerome, Miera, Rabeya, Kate, Sally and

xiii

Manue, thank you so much for the friendship that we have and you guys make me feel that I am having a big family in Australia. Thank you so much!

Highest appreciation also goes to my sponsors University of Science Malaysia (USM) and Ministry of Higher Education Malaysia. Thank you so much for sponsoring my study and allow me to have the opportunity to learn from great scientists and wonderful people in UNSW. I also would like to thank the teachers and staffs of the Eastlakes Public School and SMOOSH who were always being nice and supportive to me and my children. We will always miss all of you.

To my beloved family, my mum and my late father, thank you so much for your prayers, unconditional love, support and sacrifice. To my beloved husband Hasrul Hafizan and my sweet brilliant children; Iman, Fawwaz and Muaz, thank you so much for your understanding, your supports, patience, sacrifice and love. This PhD is yours!

THANK YOU

xiv

LIST OF PUBLICATION AND PRESENTATION

PUBLICATION

Wu, Z., Palanimuthu, D., Braidy, N., Salikin, N. H., Egan, S., Huang, M. L., & Richardson, D. R. (2020). Novel multifunctional iron chelators of the aroyl nicotinoyl hydrazone class that markedly enhance cellular NAD+/NADH ratios. British Journal of Pharmacology, 177(9), 1967-1987.

In preparation; *Salikin, N. H., Nappi, J., Majzoub, E. M. and Egan, S. Marine microbial biofilms; A reservoir for anti-parasites drug discovery. (Review article)

*Salikin, N. H., Daim, M. F. and Egan, S. Novel Nematode Killing Protein-1 from a marine bacterium Pseudoalteromonas tunicata and the resulting Caenorhabditis elegans responses

*Salikin, N. H., Majzoub, E. M. and Egan, S. Caenorhabditis elegans native gut microbiota alleviates the toxicity of Nematode Killing Protein-1 produced by marine bacterium Pseudoalteromonas tunicata

PRESENTATION

*Salikin, N. H., Daim, M. F. and Egan, S. Genetic analysis of antinematode activity in a marine epiphytic bacterium; Pseudoalteromonas tunicata. Marine Biotechnology Convention. Tauranga, New Zealand, 2017. (Oral presentation)

*Salikin, N. H. and Egan, S. Characterisation of antinematode activity by a marine epiphytic bacterium Pseudoalteromonas tunicata. 3rd Australia New Zealand Marine Biotechnology Society Conference. Sydney, Australia, 2019. (Oral presentation)

*Salikin, N. H. and Egan, S. Nematode killing activity by a marine epiphytic bacterium Pseudoalteromonas tunicata against Caenorhabditis elegans. 8th Congress of European Microbiologists (FEMS), 2019. Glasgow, Scotland. (Poster presentation)

*Publications or presentations relevant to this thesis.

xv

LIST OF ABBREVIATIONS AND SYMBOLS

% percent oC degree Celcius ± plus minus µ micro (10-6) aa amino acid Amp ampicillin ANOVA analysis of variance BLAST Basic Local Alignment Search Tool bp basepair(s) CBM6 Carbohydrate binding module 6 CBM35 Carbohydrate binding module 35 CFU colony forming units cm centimetre(s) Cm chloramphenicol df degree of freedom CMSI Centre for Marine Scince and Innovation DGM Defined gut microbiota DNA deoxyribonucleic acid EDTA ethylene diamine tetra acetic acid, trisodium salt F F-test Fw forward primer g gram GFP green fluorescent protein GLM generalized linear model h hour(s) hp hypothetical protein IMG Integrated Microbial Genomes Kan Kanamycin kb kilobase(s) kDa kilodalton(s) L litre(s) L4 C. elegans larval stage 4 LB10 Lysogeny Broth (medium) m Milli (10-3) M molar MA DifcoTM Marine Broth 2216 solidified with 1.5% agar MB DifcoTM Marine Broth 2216 min minute(s) mL millilitre xvi mol Mole n number of replicates N2_DGM C. elegans N2 with a defined gut microbiota N2_UGM C. elegans N2 with an undefined gut microbiota NGM Nematode growth media nMDS non-metric multidimensional scaling OD Optical density ORF Open reading frame zOTU zero-radius operational taxonomic unit p Pico (10-12) PBS Phosphate Buffer Saline PCR Polymerase Chain Reaction PERMANOVA Permutational Multivariate Analysis of Variance PFT Pore forming q q-value rpm revolutions per minute Rv reverse primer s second(s) SDS PAGE Sodium dodecyl sulfate–polyacrylamide gel electrophoresis sp. spp. several species TAE Tris-acetate-EDTA buffer Tn transposon UGM undefined gut microbiota V Volt v/v volume per volume vs versus wt wild type strain w/v weight per volume x g gravitational force

xvii

LIST OF FIGURES

Figure 1.1 C. elegans lifecycle and body anatomy.………………………….….…...9

Figure 1.2 Schematic diagram representing C. elegans functional role as a model nematode for the development of antinematode drugs……...………………………….12

Figure 1.3 Schematic diagram representing the daf-2/daf-16 ILR signalling and p38- MAPK pathways in C. elegans.…………………………………………………………17

Figure 1.4 Surface associated marine bacteria live on the nutrient-rich marine surfaces for example macroalgae or cnidarian in a form of biofilms.………………….26

Figure 1.5 Bioactive compounds produced by the surface associated marine bacterium Pseudoalteromonas tunicata D2.………………….…………………...……31

Figure 2.1 Schematic diagram of the gene cloning protocol…………….…………70

Figure 2.2 Schematic diagram of 13.8 kb P. tunicata D2 insert encoded in the HG8 fosmid.……………………………………………………………...…………………..78

Figure 2.3 HG8 transposon mutant generation ………………………….…………87

Figure 2.4. C. elegans survival on the 7C8 mutants complemented with the genes encoding for the uncharacterised hypothetical protein of P. tunicata D2 (hp1, hp2 or hp1/hp2) expressed in trans.…………………………………………………………….89

Figure 2.5 C. elegans survival on E. coli clones HP1, HP2 and HP1/HP2………………………………………………………..………………………90

Figure 2.6 Effect of E. coli clones HP1, HG8 and the negative control BD24 exposure on (A) C. elegans egg hatching efficiency and (B) C. elegans brood size.……………………………………………………………………………………..92

Figure 2.7 C. elegans survival on live and heat-killed E. coli clones HP1 and HG8 and the negative control E. coli BD24 bacterial strains……..…………………………………………………………………………….93

xviii

Figure 2.8 C. elegans survival on E. coli HP1, HG8, BD24 and their corresponding cell-free supernatants.……………….……………….…………………………………97

Figure 2.9 C. elegans survival in the soluble and insoluble protein fractions of E. coli clones HP1, HG8 and BD24 using the 24-well microtiter plates……………………………………………………………………………………98

Figure 2.10 SDS Polyacrylamide gel (12%) showing soluble and insoluble protein fraction of E. coli clones HP1, HG8 and the control strains; BD24 and EPI300…………………………………………….….……………………………….100

Figure 2.11 Multiple sequence alignment of Nkp-1 protein sequence and closely related protein sequences………………………………...……………………………101

Figure 2.12 Schematic diagrams representing Nkp-1 protein sequence and the 3D protein models generated by different software……….………………………………102

Figure 2.13 Characterised protein templates showing resemblance to Nkp-1 3D protein models.………………………………….….………………………………….106

Figure 2.14 Schematic diagram representing the proposed function of key genes encoded on HG8 that are involved in involved in antinematode activity.……………………………………………….………………………….……111

Figure S2.1 Nkp-1 protein bands from E. coli HP1 soluble protein extracts induced with different L-(+)-Arabinose concentration; 0.10% (w/v), 0.20% (w/v) and 1.0% (w/v) for 0, 2, 4, 8 and 18 hours at 25°C and 200 rpm of culturing condition……………………………………………………...……………………….130

Figure S2.2 Nkp-1 protein bands expressed by E. coli BL21 (λDE3) transformed with the pET28b(+)hp1 vector.……………………..………………………………………131

Figure 3.1 Schematic diagram of C. elegans training and the assessment of innate aversion and induced associative learning avoidance behaviour between the trained and naïve nematodes.……………………………………...……………………………….143

Figure 3.2 E. coli HP1::GFP and HG8::GFP bacterial colonisation within C. elegans gastrointestinal system.………………………………………….…………………….149

xix

Figure 3.3 Relative fluorescence representing the colonisation of E. coli strains HP1::GFP, HG8::GFP and the negative control BD24::GFP within C. elegans body.…………………………………………………………..………………………150

Figure 3.4 Bacterial colonisation assay of E. coli HP1::GFP, HG8::GFP and BD24::GFP against C. elegans.………………………..………………………………150

Figure 3.5 Examples of C. elegans body structure after exposure to the Nkp-1 expressing clones (E. coli HP1::GFP and HG8::GFP) and the negative control strain BD24::GFP.………………………………………………..………………………….152

Figure 3.6 Proportion of C. elegans with evidence of morphological changes after exposure to different GFP-tagged E. coli clones.………...…………………………… 155

Figure 3.7 Examples of physical appearances observed on C. elegans treated with the total insoluble protein fractions of E. coli HP1, HG8 and the negative control BD24 strains.…………………………………………………………………………………157

Figure 3.8 Proportion of C. elegans with morphological changes after exposure to the protein fractions of strains HP1 and HG8.…………...………………………………...159

Figure 3.9 Propidium iodide (PI) staining on C. elegans exposed to E. coli strains HP1, HG8 and BD24 strains.………………………………….……………………….163

Figure 3.10 Proportion of C. elegans with loss of cell membrane integrity after 48 hours of exposure to E. coli HP1 and HG8 bacteria.………………………………….164

Figure 3.11 Assessment for C. elegans innate and associative learned avoidance behaviour against the toxic E. coli strains HP1 and HG8 and the non-toxic E. coli BD24 and A1A strains.……………………………………………………………………….166

Figure 3.12 C. elegans genes expression of different immunity cascades including the (A) daf-2/daf-16 ILR signalling, (B) p38-MAPK pathway and (C) ERK-MAPK pathway………………………………………………...………….…………………168

Figure 3.13 The expression level of several downstream genes that are important for C. elegans protective response against bacterial toxicity………….………………….169

xx

Figure 3.14 Schematic diagram representing the proposed mode of action (MOA) of Nkp-1 expressing bacteria (E. coli HP1 and HG8) against C. elegans.…………………………………………………………………………….….175

Figure 3.15 Schematic diagram representing the proposed C. elegans response to the toxicity of Nkp-1 expressing clones and their protein extracts.…………………………………………………………………………...…...181

Figure S3.1 Example of size reduction observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones…………………….199

Figure S3.2 Example of internal organ damage observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones………………….200

Figure S3.3 Example of pharynx distortion observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones……………………..201

Figure S3.4 Example of vacuole formation observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones…………………….202

Figure S3.5 Example of internal hatching observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones……………………..203

Figure S3.6 Melt curves of the amplified qPCR products of C. elegans innate immunity genes.…………………………………………………………...………………...…...204

Figure 4.1 Schematic diagram representing the preparation of enriched soils and establishment of C. elegans undefined gut microbiota.……………………………………………...………………………...…...213

Figure 4.2 Schematic diagram representing the establishment of C. elegans N2_DGM with a defined gut microbiota.…………………………………………………….………………………215

Figure 4.3 Comparison of Shannon diversity and richness (zOTU counts) of bacterial communities between SOIL and C. elegans N2_UGM microbial samples……………………...…………………..…………………………………….221

Figure 4.4 Taxonomic profiles of bacterial communities shown at the level of all SOIL and N2_UGM samples.…………………..………………………………….224

xxi

Figure 4.5 C. elegans N2 monoxenic and N2_UGM survival against E. coli HP1, HG8 and the non-toxic control BD24 bacterial strains.……………………………………………………………...………………….226

Figure 4.6 Morphology of C. elegans N2 after the establishment of a defined gut microbiota (DGM).………………………………...………………………………….228

Figure 4.7 The average of bacterial relative abundances associated with (A) the starting communities and (B) the resulting gut microbiota of C. elegans N2_DGM at the genus level…………………………………………………………………………….231

Figure 4.8 Survival of C. elegans N2_DGM and N2 monoxenic when exposed to E. coli HP1, HG8 and the non-toxic control BD24 bacterial strains………………………233

Figure 4.9 Relative fluorescence representing the colonisation of E. coli HP1::GFP, HG8::GFP and the non-toxic control E. coli BD24::GFP in C. elegans N2 monoxenic and N2_DGM.………………………………………….………………………………….235

Figure 4.10 Proportion of C. elegans N2 monoxenic and N2_DGM with pharynx distortion………………………………………...…………………………………….237

Figure 4.11 Proportion of C. elegans N2 monoxenic and N2_DGM with internal hatching.………………………………………………………………………………238

Figure 4.12 Proportion of C. elegans N2 monoxenic and N2_DGM with internal organ (intestine or gonad) damage.……………………….….………………………………239

Figure 4.13 Proportion of C. elegans N2 monoxenic and N2_DGM with deformed anal region (dar).……………………………...…...……………………………………….240

Figure S4.1 Rarefaction curve (A) before and (B) after normalisation………………………………………….………………………………274

Figure S4.2 NMDS ordination of the SOIL and C. elegans N2_UGM bacterial communities using Bray-Curtis dissimilarity of transformed rarefied zOTU data (stress: 9.096333e-05).………………………………………………………………….….….275 Figure S4.3 Taxonomic profiles of bacterial communities shown at the class level of all SOIL and N2_UGM samples.…………...…………………………………………276

xxii

Figure S4.4 Taxonomic profiles of bacterial communities shown at the family level of all SOIL and N2_UGM samples.……...………………………………………………277

Figure S4.5 Microscopic examination on C. elegans N2_UGM after the establishment of an undefined gut microbiota from the enriched SOIL.…………………………….278

Figure S4.6 Morphology of C. elegans N2 after the establishment of an undefined gut microbiota (UGM).……………………………...…………………………………….279

Figure S4.7 E. coli HP1::GFP, HG8::GFP and non-toxic control BD24::GFP bacterial colonisation in C. elegans N2.…………………………………………………………280

Figure S4.8 E. coli HP1::GFP, HG8::GFP and non-toxic control BD24::GFP bacterial colonisation in in C. elegans N2_DGM………………………….……………………281

Figure 5.1 Schematic diagram representing the antinematode drug discovery pipeline from the marine environment………………………………………………………….292

xxiii

LIST OF TABLES

Table 1A (in APPENDIX I) Example of parasitic nematode infection in human and clinical symptoms ………………………………………………………………….….302

Table 2A (in APPENDIX I) Example of parasitic nematodes, the associated host and clinical symptoms…………………………………………………………………304

Table 3A (in APPENDIX I) Example of fish parasitic nematodes, associated fish host and clinical symptoms ………………………………………….………………….….306

Table 4A (in APPENDIX I) Example of livestock parasitic nematodes, associated host and clinical symptoms ……………………………………………………307

Table 1.1 Nematicidal compounds produced by bacteria isolated from terrestrial environments and their mode of actions (MOAs) against the target nematodes ……….21

Table 1.2 Examples of anthelmintic or nematicidal bioactivities isolated from marine bacteria …………………………………………………………………………27

Table 1.3 Examples of anthelmintic or nematicidal bioactivities isolated from marine eukaryotic organisms …………………………………………………………..29

Table 2.1 Bacterial strains, vectors and P. tunicata ORFs used in this study……..62

Table 2.2 List of primer pairs used in this study…………………………………………………………………………………….69

Table 2.3 Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata HG8 insert ………...………………………………….…………………79

Table 2.4 Antimicrobial assay of E. coli clones HP1, HG8 and BD24 on human and marine bacterial strains……………………...………………………………………….95

Table 2.5 Antifungal assay of E. coli clones HP1, HG8 and BD24 on marine fungal and yeast isolates………….….………………………...……………………………….96

Table S2.1 C. elegans survival on 7C8 mutants complemented with P. tunicata D2 uncharacterised hypothetical proteins (hp) wild type (wt) ……………….…...……….124

xxiv

Table S2.2 C. elegans survival on E. coli clones expressing individual P. tunicata D2 hypothetical proteins……………………………………………………………….….125

Table S2.3 C. elegans eggs hatching competency and brood size on E. coli clones expressing individual P. tunicata D2 hypothetical proteins ………………………….126

Table S2.4 C. elegans survival on live and heat-killed E. coli clones expressing individual P. tunicata D2 hypothetical proteins ……...……………………………….127

Table S2.5 C. elegans survival on E. coli HP1, HG8 and BD24 cells and cell-free supernatants…………………………………………………………………….……..128

Table S2.6 C. elegans survival on E. coli HP1, HG8 and BD24 soluble and insoluble protein fractions.………………………………………………………………………129

Table 3.1 Bacterial strains and vectors used in this study...….……………….…. 136

Table 3.2 List of C. elegans genes and primer pairs used in qPCR reaction ….….146

Table 3.3 Summary of morphological changes for C. elegans resulting from exposure to different E. coli bacterial cells, their soluble or insoluble protein fractions.………………………………………………………………...………….…160

Table 3.4 Assay on potential hydrolysis enzymatic activity of E. coli strains HP1 and HG8 bacterial culture and their protein extracts against the different substrates.…………………………….……………………………………………….162

Table S3.1 qPCR efficiency and R2 value of C. elegans genes tested in this study…………………………………………………………………………………...191

Table S3.2 Relative fluorescence quantification indicating E. coli HP1::GFP, HG8::GFP and BD24::GFP colonisation in C. elegans gastrointestinal system…….………………………………………………………………………..….192

Table S3.3 Bacterial colonisation assay of E. coli HP1::GFP, HG8::GFP and BD24::GFP colonisation in C. elegans gastrointestinal system ………………………………………………………………………….……..193

Table S3.4 Morphological changes on C. elegans resulted from E. coli HP1::GFP and HG8::GFP bacterial colonisation ……………………………………………………..194

xxv

Table S3.5 Morphological changes on C. elegans resulted from E. coli HP1 and HG8 total protein fraction exposure …………………………………………………….….196 Table S3.6 Proportion of C. elegans with membrane loss of integrity (necrosis) resulted from E. coli HP1 and HG8 bacterial exposure ………………………………………………...…………………………….197

Table S3.7 Innate and associative learned avoidance behaviour of C. elegans against E. coli HP1 and HG8 ………………………………………………………………….198

Table S3.8 Differential expression of innate immunity genes in C. elegans resulted from E. coli HP1 exposure …………………………………………………………….197

Table 4.1 Bacterial strains and vectors used in this study …………………….….210

Table 4.2 List of isolated gut bacteria from C. elegans N2_UGM previously raised in the enriched SOIL that were detected in the N2_DGM nematodes ………………...229

Table 4.3 Summary of morphological changes observed on C. elegans N2 monoxenic and N2_DGM exposed to the toxic Nkp-1 expressing strains (E. coli HP1::GFP and HG8::GFP) and the non-toxic control BD24::GFP……………………………………………………………………………241

Table S4.1 Assessment on the bacterial community structure in SOIL and nematodes samples; C. elegans N2_UGM and N2_DGM …………………………………….….254

Table S4.2 Comparison of diversity and richness of bacterial communities in SOIL and N2_UGM microbial DNA samples.……………………………………….……...254

Table S4.3 PERMANOVA based on Bray-Curtis (BC) dissimilarity measure for square-root transformed abundances of bacterial communities in SOIL and C. elegans N2_UGM.……...………………………………………….….……………………….255

Table S4.4 Relative abundance of zOTUs found to be significantly affected by the sample types; SOIL or N2_UGM microbial samples (ANOVA with p adjusted < 0.05).……………………………………………………………...…...………………256

Table S4.5 C. elegans N2 monoxenic and N2_UGM survival on E. coli strains HP1, HG8 and the non-toxic control BD24 …………………………………………………269

xxvi

Table S4.6 The difference of bacterial abundance between the starting bacterial communities and the defined gut microbiota of C. elegans N2_DGM at the genus level….……….…………………………………………………………..…….270

Table S4.7 Survival of C. elegans N2 with a defined gut microbiota (N2_DGM) upon exposure to the Nkp-1 expressing E. coli clones ………………………………………271

Table S4.8 Comparison of fluorescence intensity of GFP-tagged E. coli clones colonising C. elegans monoxenic N2 and N2_DGM with defined gut microbiota ……272

Table S4.9 Proportion of morphological changes observed on N2 monoxenic and N2_DGM nematodes following exposure to GFP-tagged E. coli strains …………….273

xxvii

CHAPTER 1

General introduction and thesis aims

1

1.1 What are nematodes?

Nematodes are moulting roundworms classified under the phylum Nematoda (Kingdom Metazoa) with the body size ranging from 0.2 mm to 6 m (Blaxter and Koutsovoulos, 2015). Nematodes are referred to as ‘ with a tube within a tube’ representing the gastrointestinal tube extending from mouth (at the anterior end) to anus (at the posterior region) which is located in a long-round shaped body cavity with fluid-filled pseudocoelom (Basyoni and Rizk, 2016). The nematode body is built on approximately 1000 somatic cells and equipped with digestive, reproductive, nervous and excretory systems, however without a distinct respiratory or circulatory function (Basyoni and Rizk, 2016). The animals also lack a well-defined head, eyes and appendages and rely on their chemosensory and mechanosensory neurons embedded in the thick cuticle of the outer covering of the body to respond to various environmental stimuli (Kaplan and Horvitz, 1993; Goodman and Sengupta, 2019).

Nematodes are ubiquitous in nature and can be found in terrestrial, freshwater and marine environments (Bik et al., 2010). Approximately, more than 14 000 nematode species are distributed on the earth, representing more than 80% of metazoans diversity in the soil (Hodda et al., 2009; Kergunteuil et al., 2016). While in the deep-sea, more than 90% of metazoan abundance on the ocean floor consists of nematodes (Dell’Anno et al., 2015). However, the true nematode species level diversity is predicted to be at least 1 million (Lambshead, 1993; Blaxter and Koutsovoulos, 2015).

From an ecological perspective, nematodes play an important role in balancing the food web (Ferris, 2010) due to their role in nutrient mineralisation resulting in increasing biomass and productivity (Gebremikael et al., 2016). However, some nematodes are also considered facultative or obligate parasites (reviewed in (Blaxter and Koutsovoulos, 2015)), causing infectious diseases in humans and major economic constrain in agricultural and livestock sectors (Tang et al., 2014; Rashid et al., 2019).

2

1.2 Nematodes as human parasites

Humans are vulnerable to infectious diseases caused by parasitic helminths (nematodes) resulting in morbidity and mortality within the population (Cox, 2004; Bañuls et al., 2013; Viney, 2017). There are almost 300 nematodes associated with zoonotic diseases that are able to infect humans, including some of the most devastating parasites such as Ascaris lumbricoides (roundworm), Ancylostoma duodenale (hookworm), Gnathostoma spinigerum, Halicephalobus gingivalis and Trichinella spiralis (Trichina worm) (Table 1, Appendix I) (Ashford and Crewe, 2003; Cox, 2004). While some parasitic nematodes (e.g. Ancylostoma duodenale, Strongyloides stercoralis and Halicephalobus gingivalis) can penetrate human skin or invade through the existing skin lesion (Papadi et al., 2013; Bryant et al., 2018), several parasites infect humans via ingestion of food products contaminated with the embryonated eggs (e.g. Ascaris lumbricoides and Trichuris trichiura) (Jourdan et al., 2018; Bundy et al., 2020) or from eating raw or undercooked freshwater fish, birds, frogs or reptiles contaminated with the parasitic nematode larvae (e.g. Gnathostoma spinigerum, Dracunculus medinensis, Eustrongylides sp. and Trichinella spiralis) (Magnino et al., 2009; Eberhard et al., 2016; Frean, 2020). Some parasites e.g. Wuchereria bancrofti, Brugia malayi, Brugia timori and Onchocerca volvulus infect humans via vector transmission involving mosquitoes (Culex quinquefasciatus) (Rai et al., 2019) or black fly (Simulium damnosum), resulting in debilitating neglected tropical diseases including lymphatic filariasis or river blindness (Hendy et al., 2018; Rai et al., 2019).

Despite more than a century of effort to control parasitic infections, disease remains a major and growing public health and economic threat particularly for developing and low- income countries. Approximately 24% of the global human population, corresponding to 1.5 billion people are suffering from parasitic helminth infections (WHO, 2020) (Garcia- Bustos et al., 2019). Apart from causing diseases (see Table 1A, Appendix I), high nematode burden also reduces human fecundity (Persson et al., 2019) and affects children through malnutrition, stunted development and cognitive delay (Ezeamama et al., 2005). Moreover, infections also impose a major socioeconomic burden, leaving people in a poverty trap, low rate of employment, poor education and unsanitary conditions (Kiani et al., 2016; Gadisa and Jote, 2019)

3

1.3 Parasitic nematodes; a significant threat to agriculture, fisheries and livestock industries

Supplying food for the growing global population is one of societies major challenges, with over 2 billion people already living with insufficient food supply (Hassan et al., 2013). Unfortunately, diseases caused by parasitic nematodes pose a substantial threat to the food industry and the situation is predicted to worsen with an estimated 35% increase in the human population by 2050 (Shaheerah Alsounosi Alshareef and Mohamed, 2019; Ramos, 2020) (World Bank Report, 2008).

There are approximately 4 100 plant-parasitic nematodes (PPNs) (Table 2A, Appendix I) that cause disease (Abad et al., 2008; Bernard et al., 2017) either through damage to the root system, retarding the plant development or by exposing plants to secondary bacterial, fungal or viral infections (Villate et al., 2012; Jones et al., 2013; Liu and Park, 2018; Kassie et al., 2020). It is estimated that damages caused by PPNs result in >12% loss in global crop productivity and an average annual loss of ~ US$ 215 billion (Askary and Martinelli, 2015). In Australia for example, root-lesion PPNs e.g. Pratylenchus thornei is estimated to be responsible for losses of 85% and 25% in the expected wheat and chickpea productivity respectively (Thompson et al., 2000; Smiley, 2010; Zwart et al., 2019). Until now, eradicating the PPNs infestation remains a major challenge given their microscopic size and the often long evolutionary history that they share with the plants (Dutta et al., 2019).

Parasitic nematodes also impede the productivity of fisheries and aquaculture industries (Table 3A, Appendix I), resulting in worldwide economic losses and health hazards to the consumers (Shinn et al., 2015; Mehrdana and Buchmann, 2017). For example, global financial loss in the finfish industry due to parasitic infection is estimated to be as large as US$ 134 million per annum (Shinn et al., 2015). Major fish products such as Atlantic mackerel, herring, European hake, Atlantic cod and anchovy are commonly associated with parasitic nematodes e.g. Anisakis sp. and Pseudoterranova sp. (Levsen et al., 2018) and transmitted to human via ingestion of undercooked or raw fish (Aibinu et al., 2019; Shamsi, 2019). While causing zoonotic diseases, parasitic nematodes also can cause 4 allergies to hypersensitive individuals due to heat-resistant allergenic peptides produced by the nematodes (Moneo et al., 2005; Mehrdana and Buchmann, 2017), resulting in 56 million of reported cases related to allergic response to parasite nematode-contaminated fish products (WHO, 2012).

Lungworm and/or gastrointestinal parasitic nematodes of livestock are devastating, causing different pathophysiological symptoms, reduced meat quality and animal mortality (Table 4A, Appendix I) (Mushonga et al., 2018; Elseadawy et al., 2019). In Australia, economic loss due to parasite infection (e.g. Ostertagia sp. and Trichostrongyfus sp.) and the cost of control management is estimated to be ~ AU$ 1 billion (McLeod, 1995; Sackett et al., 2006; Roeber et al., 2013) whereas in Kenya, South Africa and India, ~ US$ 26, 46 and 103 million are spent just to control the Haemonchus contortus nematode infection of livestock (McLeod, 2004; Maqbool et al., 2017). Apart from causing diseases and reduced animal fecundity nematode parasites such as Ostertagia ostertagi (brown stomach worm) and Dictyocaulus viviparus (lungworm) reduce dairy production levels (~ 1.25 L/day/animal or 1.6 L/day/cow respectively) (May et al., 2018; Rashid et al., 2019). Furthermore, parasitic nematodes impose food safety concern due to the potential of transmitted zoonotic diseases to consumers (Lopes et al., 2015). Unfortunately, the economic loss attributed to nematode infection and the prevalence of antinematode drug resistance are predicted to escalate due to increasing temperature, extended parasite nematode transmission period and increasing animal host susceptibility to parasite diseases as a result of climate change. (GLOWORM Project Report; https://cordis.europa.eu/project/id/288975/reporting) (Fox et al., 2015; Verschave et al., 2016; Morgan et al., 2019).

5

1.4 The increasing trend of drug resistance among the parasitic nematodes; an urgent need for new antinematode treatments

Currently, infection control strategies for parasitic nematodes rely predominantly on water, sanitation and hygiene intervention (WASH) (Haldeman et al., 2020). However, WASH improvement is often a challenge in profoundly affected low-income countries (Hutton and Bartram, 2008; Nery et al., 2015). Vaccine development is ongoing, however current research focuses only on a narrow spectrum of target organisms and long-term protection from a vaccine is still under debate (Hawdon, 2014; Diemert et al., 2017; Haldeman et al., 2020). Thus, arguably the most promising longer-term control strategy is dependent on pharmaceutically-derived chemotherapeutic treatments to kill parasitic nematodes and/or mitigate the spread of infection (Abongwa et al., 2017; Zajíčková et al., 2020).

For the community antinematode treatment, single-dose albendazole or mebendazole drug tablet is recommended by the World Health Organisation (WHO) to reduce parasitic nematode burden in high-risk individuals i.e. children and pregnant women (WHO, 2017). Antinematode drugs are also widely used in aquaculture and livestock industries for deworming activities of the animals (Buchmann and Bjerregaard, 1990; Enejoh and Suleiman, 2017; Orobets et al., 2019). This is performed either by oral administration, injection or drenching the infected animals with different regimes of antinematodal drug treatments (Beuvry et al., 2003; Melville et al., 2016; Orobets et al., 2019). Unfortunately, long-term single drug consumption and the application of sub-optimal dosages are resulting in increasing prevalence of drug resistance among parasitic nematodes (Geerts and Gryseels, 2000).

Increasing cases of antinematode drug resistance has been reported across the global livestock sector (Kaplan and Vidyashankar, 2012). For example, in year 2000, about 40% of sheep farms in Western Australia (WA) were diagnosed with avermectin-resistance involving nematode parasite Teladorsagia circumcincta (Besier and Love, 2003). However, resistance increased to 60% after five years and it is estimated that more than 80% of WA sheep farms had cases of avermectin-resistant T. circumcincta in 2012 (Love,

6

2007; Kaplan and Vidyashankar, 2012). The current prevalence of drug resistance is not only limited to the older classes of antinematode drugs but also those introduced in recent years. For example, monepantel resistance in T. circumcincta, Trichostrongylus colubriformis and H. contortus has been reported in New Zealand and the Netherlands within just five years of launching the drug (Scott et al., 2013; Van den Brom et al., 2015). Even worse, multiple drug resistance cases have been reported in Brazil when treatments using ivermectin, albendazole, levamisole, trichlorfon, moxidectin, closantel and a combination of levamisole, ivermectin and albendazole against sheep infected with parasites Haemonchus sp., Trichostrongylus sp., Strongyloides sp., sp. and Cooperia sp were ineffective (Sczesny-Moraes et al., 2010). Multiple antinematode drug resistance against the combinations of derquantel and abamectin or abamectin, albendazole, levamisole and closantel against H. contortus was also detected in sheep farms in New South Wales (NSW), Australia (Lamb et al., 2017). An increase of 47% resistance cases from 2012 to 2018 was also reported in domestic dogs in the US (Drake and Carey, 2019). The increasing prevalence of resistance cases in livestock or domestic animals is intimidating since those parasites can also be transmitted to humans, resulting in devastating zoonotic diseases (see examples in Table 4A, Appendix I) (Geerts and Gryseels, 2000; Ngcamphalala et al., 2020). Alarmingly, there are also reports of drugs resistance in parasitic nematodes that cause human neglected tropical diseases (NTD) i.e. lymphatic filariasis and onchocerciasis (river blindness) (Cobo, 2016; Akinsolu et al., 2019).

Drug resistance is often linked to genetic mutation in those parasites. Such mutations can include those that result in a reduction of the nematode drug-target receptor, affinity of the drug with the receptor or the absence of enzymes for the drug activation (Gilleard, 2006). Almost all of the currently available anthelmintic drugs and nematicides classes including Piperazine, Benzimidazoles, Levamisole, (pyrantel and morantel), Paraherquamide, Ivermectin (macrocylic lactones and milbemycins), Emodepside (cyclodepsipeptides, PF1022A) and Nitaz0oxanide have been reported with antinematode drug cases (Holden-Dye and Walker, 2014). Unfortunately, drugs from similar classes or possessing the same mode of action are prone to have side or cross parasitic resistance (Gilleard, 2006). Given the impacts of resistant parasitic nematodes in human and economic growth, a novel antinematode chemotherapeutic agent is urgently needed as a

7 preventive control against parasite infestation (Elfawal et al., 2019; Garcia-Bustos et al., 2019).

1.5 Caenorhabditis elegans as a model organism for antinematode drug discovery

Antinematode drug research is impeded by several challenges. These include (i) similarity of biochemical reaction between parasitic nematodes and the infected host, (ii) complex parasite life cycles that involve infections in multiple hosts, (iii) different parasite geographical location and (iv) rapid development of resistant phenotypes. Therefore, an easily maintained nematode model with the capability for rapid screening of potential antinematode compounds represents a solution to some of these challenges (Mathew et al., 2016; Blasco-Costa and Poulin, 2017).

Sydney Brenner and colleagues first introduced the free-living soil nematode C. elegans in 1965 as an animal model for research including anti-infective and antinematode drug studies (Riddle et al., 1997; Leung et al., 2008; Burns et al., 2015; Kong et al., 2016; Wang et al., 2018; Queirós et al., 2019). Owing to its small size (1-1.5 mm-adult length, 80 µm-diameter), transparency, rapid life cycle (~ 3 days) and diet based on a simple bacterial culture E. coli OP50, C. elegans has emerged as a valuable animal model (Riddle et al., 1997; Artal‐Sanz et al., 2006; Frézal and Félix, 2015). C. elegans undergoes two simple life cycles, depending on the surrounding condition. In an environment with sufficient food, the nematode eggs hatch and pass through the L1, L2, L3 and L4 stage larvae to become an adult (Figure 1.1). However, in an adverse environment, for example with inadequate food, heat stress or overwhelmed population, the L1 larvae will enter different developmental route into a predauer larvae (L2d), followed by a dauer stage of which the larvae stop feeding and become stress-resistant (Frézal and Félix, 2015). Upon exposure to favourable condition (e.g. sufficient food supply), the dauer larvae develop to the L4 stage and become an adult nematode (Karp, 2018) (Figure 1.1).

8

A

B Eggs Vulva Pharyngeal grinder Pharynx Gonad Tail

Intestinal lumen Terminal bulb Intestinal cell Anus

Figure 1.1 C. elegans lifecycle and body anatomy. (A) C. elegans undergoes two different lifecycles depending on the environment. The eggs are developed in the uterus and excreted through the vulva. Under favourable conditions, the nematode eggs hatch and pass through the L1, L2, L3 and L4 stage larvae to become an adult. However, in stressed conditions i.e. inadequate food, excessive heat or overwhelmed population, the L1 larvae will enter different developmental route to a predauer larvae (L2d), followed by a dauer stage. When food is sufficient, the dauer larvae develop to the L4 stage and become adult nematodes. (B) Visualisation of C. elegans body under the microscope using the DIC filter at 10x magnification. A healthy gravid C. elegans adult hermaphrodite demonstrates an intact pharynx, body structure and internal organs i.e. intestine and gonad as indicated in (B). Scale bar indicates 100 µm. Figure A was designed using Biorender at https://biorender.com/

9

C. elegans possess several features which make the organism an efficient and low-cost surrogate organism for antinematode drug discovery. Unlike parasitic nematodes which require hosts for reproduction and maintenance (Holden-Dye and Walker, 2014; Lok and Unnasch, 2018), synchronised C. elegans can be easily propagated at the desired life stage on Nematode Growth Media (NGM) ready to be used for antinematode drug studies (Stiernagle, 2006). The earliest antinematode drug testing using synchronised C. elegans individuals was performed by exposing the animals to the nematicidal agents incorporated into the agar media (Brenner, 1974). After a few years, the screening protocol evolved rapidly with the development of high throughput screening methods employing micro-fluidic systems or high-content screening (HCS) technologies, allowing for fast and large-scale drug testing (~ 14 000 to 360 000 of compounds) using C. elegans as the model organism (Figure 1.2) (Fraietta and Gasparri, 2016; Midkiff and San-Miguel, 2019). As a result, important drug discoveries such as Benzimidazoles (Driscoll et al., 1989), ivermectin and its analogues; moxidectin, milbemycin oxime, doramectin, selamectin, abamectin, eprinomectin (Haber et al., 1991) and the nematicidal activity of crystal protein insecticide Cry5B, Cry21A from (Marroquin et al., 2000; Wei et al., 2003) can be attributed to the use of C. elegans as an effective animal model (Holden-Dye and Walker, 2014).

C. elegans shares many conserved genes and protein functions with parasitic nematodes. Analysis of the intestinal parasite Strongyloides stercoralis genetic sequences showed 85% of protein homologs to C. elegans genes. The infective stage of S. stercoralis (L3i/dauer) also shows an increased proportion of protein homologs to C. elegans dauer larvae (Mitreva et al., 2004). Genetic manipulation and RNA interference (RNAi) studies have been widely performed on C. elegans to provide a better understanding of the nematode response against the nematicidal drugs at the molecular level (Figure 1.2). Quantitative polymerase chain reaction (qPCR) and ‘omic’ technologies, i.e. transcriptomic profiling and proteomics are also being used to provide a global snapshot of the molecular response of C. elegans to drug exposure (Kumarasingha et al., 2019; Mir and Krishnaswamy, 2019). Given the conserved gene homologs and protein function among the members of phylum Nematoda, these studies provide insight into the possible mechanisms used by parasitic nematodes against similar drug exposure (Holden-Dye and Walker, 2014).

10

1.5.1 Characterising the mode of action of antinematode compound using C. elegans

The majority of drugs used to treat parasitic nematode infection target proteins that regulate neuromuscular activity including neurotransmitter receptors and ion channels (Holden-Dye and Walker, 2014). Utilisation of C. elegans as a model organism confers a better understanding of the mode of action (MOA) of potential nematicidal drugs (Figure 1.2). Studies have shown that C. elegans’ neuromuscular system, its major neurotransmitters GABA (4-aminobutyric acid) and glutamate and its enzyme choline acetyltransferase responsible for the synthesis of the neurotransmitter acetylcholine display strong similarities to parasitic roundworms Ascaris suum and Ascaris lumbricoides (Johnson and Stretton, 1985; Johnson and Stretton, 1987; Angstadt et al., 1989; Davis, 1998; Holden-Dye and Walker, 2014). Therefore, exposure of C. elegans to antinematode compounds which target the neurotransmitter receptor and ion channels has enabled the discovery of the target binding molecule and the resulting toxicity to parasitic nematodes (Weeks et al., 2018). For example, observations of body muscle contraction and spastic paralysis in C. elegans exposed to levamisole led Lewis and colleagues (Lewis et al., 1980) to determine that binding to the muscle acetylcholine receptors was key to its activity. The initiation of amino-acetonitrile derivatives (AAD) toxic activity against nematode by binding to a nicotinic acetylcholine receptor was also revealed via forwards genetic screening using C. elegans mutants (Kaminsky et al., 2008). More recently, the MOA of paraherquamide; a broad-spectrum nicotinic nematicidal alkaloid isolated from Penicillium paraherquei (Yamazaki et al., 1981) and antinematode plant-derived compounds (Hernando et al., 2019) were also elucidated using C. elegans (Schaeffer et al., 1992; Ruiz-Lancheros et al., 2011; Hernando et al., 2019).

In addition, while the insecticidal activity of B. thuringiensis Cry toxin was known (Vadlamudi et al., 1995; Zhao et al., 2020), the use of C. elegans as a model allowed the nematicidal properties and MOA of these to be revealed (Marroquin et al., 2000; Griffitts et al., 2005). Cry toxins for example Cry5B binds to the glycolipid and cadherin receptors of C. elegans’ intestinal cell prior to oligomerisation and insertion into the cell membrane to form pore leading to necrosis and nematode mortality (Griffitts et al., 2005; Peng et al., 2018).

11

Figure 1.2 Schematic diagram representing C. elegans functional role as a model nematode for the development of antinematode drugs. The propagation and handling methods of the human, animal or plant-parasitic nematodes in the laboratory are always challenging. Given the conserved genetic sequence and protein function to those parasites, C. elegans is used as a surrogate nematode for the initial screening against the potential compounds or microorganisms with nematotoxic properties either via the conservative agar plate method or through high throughput screening technologies. C. elegans-based research offers several advantages including nematode genetic manipulation, determination of drug MOA, evaluation of the resulting nematode responses and a large-scale initial drug screening due to easy nematode maintenance and propagation in the laboratory. Figure was designed using Biorender at https://biorender.com

12

Some microbial products are also toxic to C. elegans due to their lytic degradative properties towards nematode body structures that are predominantly composed of chitin, collagen and lipids (Page and Johnstone, 2007). For example, Pseudomonas aeruginosa produce chitinase which can hydrolyse C. elegans cuticle and eggshell (Chen et al., 2015). Pseudomonas aeruginosa and some other Rhizobacteria i.e. Paenibacillus polymyxa, Bacillus subtilis, , Lysinibacillus sphaericus and xylosoxidans secrete chitinase and protease resulting in the degradation of plant parasitic nematode Meloidogyne incognita’s eggs and death (Soliman et al., 2019). There is also evidence that shows the nematicidal activity of Chryseobacterium nematophagum against C. elegans and other important veterinary nematodes; Haemonchus contortus and Ostertagia ostertagi is due to the bacterial collagenase activity (Page et al., 2019). Further analysis of the Chryseobacterium nematophagum genome led to the identification of collagenase and chitinase expressing genes that may potentially act as novel virulence factors (Page et al., 2019).

1.5.2 Understanding nematode behavioural response and innate immunity regulation towards antinematode compounds

Owing to the conserved sensory anatomy, C. elegans also serves as a powerful model to study the behaviour of parasitic nematodes (Figure 1.2) (Rengarajan and Hallem, 2016). C. elegans’ avoidance behaviour is controlled by chemosensory neurons with ciliated projections exposed to the environment to detect different molecular cues (Meisel and Kim, 2014). C. elegans are predominantly attracted to bacteria, however some bacteria can repel the nematode via the production of molecules that results in the nematode avoidance behaviour (Meisel and Kim, 2014). For example, C. elegans avoid the pathogenic bacterium Serratia marcescens due to its bacterial surfactant serrowettin W2 (Pradel et al., 2007). Interestingly, nematode avoidance behaviour can be learnt upon subsequent bacterial encounters. The response is known as induced associative learning avoidance behaviour (Zhang et al., 2005) and involves a neuronal circuit that is different from the initial innate avoidance behaviour. Taking advantage of this innate protective mechanism, the behaviour of nematodes against nematicidal compounds/drugs could be

13 determined via food choice or lawn-leaving assays on agar media (Zhang et al., 2005; Meisel and Kim, 2014).

There is evidence that the daf-2/daf-16 ILR (insulin like receptor) signalling pathway controls both the innate avoidance and associative aversive learning behaviour in C. elegans (Vellai et al., 2006; Hasshoff et al., 2007). The daf-2/daf-16 ILR signalling is a general C. elegans stress response towards heat shock (Singh and Aballay, 2006), oxidative stress (Chávez et al., 2007) and toxic microorganisms such as Pseudomonas aeruginosa, Bacillus thuringiensis, Salmonella typhimurium, Enterococcus faecalis and Coxiella burnetii (Garsin et al., 2003; Evans et al., 2008; Jia et al., 2009; Wang et al., 2012; Battisti et al., 2017). This signalling pathway is also important for C. elegans longevity by regulating the arrested development of L1 stage larvae to dauer during starvation (Lakowski and Hekimi, 1998; Hu, 2018). This pathway comprises a series of highly conserved proteins including the insulin-like receptor DAF-2 (a transmembrane tyrosine kinase) and the transcription factor DAF-16 (a FOXO/forkhead family transcription factor that is negatively regulated by DAF-2) (Figure 1.3). In the presence of antagonist ligand (insulin-like peptides) such as INS-1 or downregulation/mutation in daf-2, the AGE-1 (a phosphatidylinositol‐3‐OH kinase), PIP3 (phosphatidyl‐inositol trisphosphate), PDK‐1 (a 3‐phos‐phoinositide‐dependent kinase 1), AKT‐1 and AKT‐2 (protein‐Ser/Thr kinases) are not activated thus resulting in non-phosphorylation of DAF- 16. The non-phosphorylated DAF-16 is then translocated into the nucleus which eventually results in subsequent transcriptional regulation of multiple downstream genes expressing antimicrobial effectors, metabolic and stress response and toxin detoxification (Figure 1.3) (Murphy et al., 2003; McElwee et al., 2004; Halaschek-Wiener et al., 2005; Oh et al., 2006). Two lysozyme encoding genes; lys-7 and lys-8, the mitochondrial superoxide dismutase gene sod-3 and the dual oxidase 1 gene bli-3 are part of the downstream genes regulated by the DAF-16 transcription factor in response to ageing and/or microbial infection (Murphy et al., 2003; Zou et al., 2013; Bai et al., 2014). Furthermore, the detoxification gene B0213.15 encoding cytochrome P450 enzyme is activated in response to toxic molecules such as xenobiotics (e.g. nematicidal drugs) via daf-2/daf-16 signalling (Murphy et al., 2003; Laing et al., 2013). Cytochrome P450 enzyme has been linked to the drug resistance in insect parasites (Amenya et al., 2008) and was largely found in parasitic nematode Haemonchus contortus intestine suggesting

14 a possible underlying reason for antinematode drug resistance in the parasite (Laing et al., 2013).

The p38-MAPK (mitogen activated protein kinase) that involves NSY-1, SEK-1 and PMK-1 proteins signalling is another highly conserved immune pathway in plants and animals including in C. elegans (Figure 1.3). The p38-MAPK cascade is activated by TIR-1 (a Toll-Interleukin-1 receptor) upon exposure to different external stimuli for example the microbial pathogen-associated molecular patterns (PAMPs) i.e. or peptidoglycan (Liberati et al., 2004). The activation of PMK-1 in the cell nucleus results in the regulation of the downstream genes that are indispensable for stress response (Berman et al., 2001; Mertenskötter et al., 2013) and defence against pathogens for example antimicrobial or antifungal peptides i.e. caenacins, lysozymes and C-type lectins (Troemel et al., 2006; Pujol et al., 2008). Loss-of-function mutations in nsy-1, sek-1 or pmk-1 result in increased nematode susceptibility to toxic bacteria for example Pseudomonas aeruginosa, Staphylococcus aureus, Mycobacterium marinum and Salmonella enterica (Aballay et al., 2003; Sifri et al., 2003; Troemel et al., 2006; Galbadage et al., 2016). A study also showed that sek-1 expression was associated with dar (deformed anal region) and important for C. elegans survival against Coxiella burnetii as mutation in sek-1 reduced the dar phenotype as well as the nematode survival (Battisti et al., 2017). The p38-MAPK pathway is also pivotal for defence against pathogenic fungi Drechmeria coniospora through the induction of wound healing response encoded by the downstream gene nlp-29 in the nematode epidermis (Pujol et al., 2008). Studies also showed that PMK-1 of the p38-MAPK and DAF-16 of the daf-2/daf-16 ILR pathways act in parallel to contribute to enhanced C. elegans daf-2 mutant longevity upon exposure to toxic bacterium Pseudomonas aeruginosa (Troemel et al., 2006) indicating a highly complex C. elegans innate immunity system.

In addition to the aforementioned pathways, C. elegans possess several other conserved distinct immune cascades including the ERK-MAPK, TGF-β, FSHR-1, ZIP-2, Wnt/Hox, UPR and autophagy (Engelmann and Pujol, 2010; Ermolaeva and Schumacher, 2014). Despite its simplicity as a model organism, C. elegans immune system is highly specific

15 in which different microorganisms will elicit a distinct immune pathway (Irazoqui et al., 2010).

16

Figure 1.3 Schematic diagram representing the daf-2/daf-16 ILR signalling and p38- MAPK pathways in C. elegans. Upon exposure to pathogenic microorganisms or toxic compounds (for example an antinematode drug), several immunity cascades are activated including the daf-2/daf-16 ILR signalling and p38-MAPK pathways. In the presence of antagonist ligand (for example INS-1), expressed in C. elegans neurons or intestine) that binds to the DAF-2 receptor, the following AGE-1, PIP3, PDK-1 and AKT 1/AKT 2 are not activated (indicated by the red x) resulting in no phosphorylation activity of DAF-16, hence allowing the translocation of non-phosphorylated DAF-16 into the nucleus. The DAF-16 FOXO/forkhead family transcription factor subsequently coordinates the expression of downstream genes that are important for defence, stress response, detoxification and longevity. Upon detection of different external stimuli (for example microbial pathogen-associated molecular patterns (PAMPs) such as lipopolysaccharide or peptidoglycan), the toll-receptor TIR-1 and the following protein cascades in the p38- MAPK pathway (NSY-1, SEK-1 and PMK-1) are activated (indicated by tick mark). The PMK-1 is translocated into the cell nucleus, leading to the coordination of downstream genes expression encoding for stress response and the production of different antimicrobial and/or antifungal peptides.

17

1.6 C. elegans gut microbiota

In nature, C. elegans are commonly found in soil with decomposed plant materials and are always in close contact with highly diverse set of microorganisms (Schulenburg and Félix, 2017). The continuous microbial interaction enables the association and persistence of distinct soil microorganisms in the C. elegans gut environment. However, the synchronisation and maintenance of C. elegans in the laboratory has resulted in the majority of studies being performed on nematodes without their native gut microbiota (Stiernagle, 2006; Dirksen et al., 2016; Samuel et al., 2016; Zhang et al., 2017).

Extensive studies to unravel C. elegans natural gut microbiota have been implemented recently using 16S rRNA gene amplicon sequencing methods resulting in the discovery of (Enterobacteriaceae, Pseudomonaceae, Xanthomonodaceae) and Bacteroidetes (Sphingobacteriaceae, Weeksellaceae, Flavobacteriaceae) among of the main C. elegans gut bacterial taxa (Berg et al., 2016; Dirksen et al., 2016; Samuel et al., 2016; Zhang et al., 2017). The understanding of the dynamic interaction between C. elegans and its gut microbiota, including the role of these interactions on nematode fitness, is still in its infancy. However, several studies have shown that the gut microbiota confers various advantages to the nematode. This includes increased survival of C. elegans upon exposure to pathogenic microorganisms, for example, P. aeruginosa, Bacillus thuringiensis and Chryseobacterium sp. (Montalvo-Katz et al., 2013; Kissoyan et al., 2019). This increased survival may result from an enhanced host innate immunity and/or vitamin provided to the host by the gut bacteria (Montalvo-Katz et al., 2013; Samuel et al., 2016; Zimmermann et al., 2020). However, there is also evidence that colonisation with certain gut bacteria can result in the demise of C. elegans fitness and survival (Samuel et al., 2016).

Establishment of the gut microbiota is believed as pivotal for parasitic nematodes (White et al., 2018). For example, the gut microbiota of Trichuris muris support the nematode development and fitness as well as shaping the control for future parasite infection levels to maintain nematode persistence and survival in the associated host (White et al., 2018).

18

Furthermore, a study by Whittaker et al (Whittaker et al., 2016) also found that co- incubation of C. elegans or adult ascarid with members of the Enterobacteriaceae successfully increased nematode survival upon exposure to benzimidazole and albendazole respectively. Interestingly, those bacterial isolates can use the benzimidazole-drug compound as a carbon source to support growth, thus suggesting a mechanism by which gut microbiota protect nematodes from toxic compounds (Whittaker et al., 2016). Understanding the interactions between nematodes and their native microbiota may open up new avenues for therapeutics and drug resistance studies and this can be achieved via the utilisation of C. elegans as a versatile model organism (Gerbaba et al., 2017; Hogan et al., 2019).

1.7 Antinematode drug discovery; transition from terrestrial to marine-derived microbial compounds

Terrestrial plants and plant extracts have been documented as an ancient therapeutic treatment against parasitic nematodes (Monaghan and Tkacz, 1990) and today, extensive studies to isolate plant-derived nematicidal compounds are still ongoing (Spiegler et al., 2017; Giovanelli et al., 2018). However, the re-discovery rate of bioactive metabolites is high (Garcia-Bustos et al., 2019), reducing the number of novel compounds in the drug discovery pipeline (Gaudêncio and Pereira, 2015). Moreover, external factors e.g. specific planting season, environmental temperature and humidity may also affect the compounds reproducibility by the plant producers (Garcia-Bustos et al., 2019). Given those inevitable challenges, microbial-derived anthelmintic compounds offer a promising solution (Shalaby et al., 2020).

Extensive exploration of terrestrial microbial compounds for therapeutic drug development was initiated in the 20th century (Monaghan and Tkacz, 1990). Since then, more than 50 000 of beneficial bioactive metabolites had been successfully identified (Bérdy, 2005; Xiong et al., 2013) some of which having potent nematotoxicity via distinct mode of actions (MOAs) (Table 1.1). Unfortunately, after almost 50 years of drug screening, few novel compounds have been identified from terrestrial-borne microorganisms (Xiong et al., 2013) hence requiring a new source for antinematode drug 19 discovery to combat the rapidly growing nematode resistance. Here, the underexplored marine ecosystem with highly diverse unidentified macro and microorganism represent a new repository for novel nematicidal drug researches (Xiong et al., 2013). In fact, marine bioactive compounds have been acknowledged as having substantial chemical novelty compared to the terrestrial metabolites (Kong et al., 2010).

“We are not marine organisms. So, until about 1970, no one even thought of the ocean. It was left as a deep secret. It seemed ridiculous to me that the ocean – with such a vast habitat – had escaped anyone’s notice. But there are good reasons. People fear the ocean; it has been considered a very hostile, inhospitable place” (Marris, 2006).

The marine environment represents the largest biome on the earth (70% of the earth surface; ~ 361 million kilometers2 with average depth of 3730 meters) and provide a habitat for a wide diversity of life that outnumbers terrestrial environments (Fenical and Jensen, 2006; Costello et al., 2010; Kanase and Singh, 2018; Flemming and Wuertz, 2019). Microorganisms are abundant in this ecosystem (105 to 106 of cells per millilitre), reaching an average of 1028 to 1029 number of cells either in the open ocean, deep sea, sediment or on subsurface (Bar-On et al., 2018; Magnabosco et al., 2018; Flemming and Wuertz, 2019). However, the majority of marine bacteria remain unrecognised or uncultivable (Pascoal et al., 2020). Given the enormous diversity and untapped bioactive potential, marine microorganisms are likely to produce novel of novel bioactive compounds with a well-defined molecular architecture and biological function including the nematicidal activity (Blockley et al., 2017; Zhang et al., 2019).

20

Table 1.1 Nematicidal compounds produced by bacteria isolated from terrestrial environments and their mode of actions (MOAs) against the target nematodes

Affected Microbial producer Compound Mode of action Target nematode Reference nematode region Bacillus Crystal toxin Toxin binds to nematode glycoconjugate Ancylostoma Gastrointestinal (Wei et al., 2003; thuringiensis Cry5B, Cry21A receptor and disrupt the intestinal cells ceylanicum, Ascaris system Conlan et al., 2012; membrane integrity. This action causes fall suum, C. elegans Urban Jr et al., of nematode brood size and mortality 2013) Streptomyces Avermectin and Compounds exposure resulted in Haemonchus Neuromuscular (Burg et al., 1979; avermectinius Ivermectin (semi- pharyngeal paralysis and nematode death contortus, Brugia system Omura, 2008; synthetic) malayi, C. elegans Holden-Dye and Walker, 2014) Serratia Prodigiosin Compound is toxic against juveniles larvae Radopholus similis, Unknown (Rahul et al., 2014) marcescens and inhibit egg hatching competency Meloidogyne javanica Pseudomonas Phenazine toxin Phenazine-1-carboxylic shows fast killing C. elegans Neuromuscular (Cezairliyan et al., aeruginosa (phenazine-1- activity against C. elegans in acidic system, cell 2013; Ray et al., carboxylic , environment while pyocyanin is toxic in mitochondria and 2015) pyocyanin and 1- neutral or basic pH. The toxicity of 1- protein folding hydroxyphenazine) hydroxyphenazine is not depending on environmental pH. Continuous exposure to phenazine affects protein homeostasis and cause neurodegeneration

21

Table 1.1 (Continue) Nematicidal compounds produced by bacteria isolated from terrestrial environments and their mode of actions (MOAs) against the target nematodes

Affected Microbial producer Compound Mode of action Target nematode Reference nematode region Pseudomonas aeruginosa A and other Slow-killing activity against C. elegans C. elegans Gastrointestinal (Tan et al., 1999) undetermined effectors is based on infection-like process thus system resulting in accumulation of bacteria in the gut. Continuous exposure leads to ceased pharyngeal pumping, nematode immobility and death.

Pseudomonas Glycolipid Reduction of nematode development, C. elegans Unknown (Devaraj et al., plecoglossicida biosurfactant survival and fecundity 2019)

Bacillus simplex, B. Volatile organic VOCs reduce nematode motility and Panagrellus Unknown (Gu et al., 2007) subtilis, B. compound (VOC) i.e. cause death redivivus, weihenstephanensis, benzaldehyde, Bursaphelenchus Microbacterium oxydans, benzeneacetaldehyde, xylophilus Stenotrophomonas decanal, 2-nonanone, maltophilia, Streptomyces 2-undecanone, lateritius and Serratia cyclohexene and marcescens dimethyl disulfide

Pseudomonas aeruginosa Chitinase enzyme Chitinase degrades nematode cuticle, C. elegans Cuticle, eggs, (Chen et al., intestine and egg shell leading to the gastrointestinal 2015) animal death system

22

1.8 Surface associated marine bacteria; a reservoir of novel antimicrobial and antinematode drugs discovery

Marine inhabitants, particularly the microorganisms, are continuously exposed to multiple detrimental or interactions (Rao et al., 2005; Welsh et al., 2016) and different physical-chemical variables such as fluctuating temperature, pH, UV exposure, salinity, toxic compounds, and desiccation particularly in the intertidal zone (Figure 1.4) (Chiu et al., 2005; Ortega-Morales et al., 2010; de Carvalho, 2018). As a survival strategy, some of marine bacteria adhere to each other and/or surfaces and embedded in enclosed matrix to form a biofilm (Vlamakis et al., 2013; Antunes et al., 2019).

The continuous development of biofilm on marine surfaces leads to epibiosis, which involves multispecies biofilm formation (Egan et al., 2013; Katharios-Lanwermeyer et al., 2014). Macroalgal surfaces for example are a hot-spot for colonization by opportunistic epibionts such as algal spores, larvae, , fungi and other bacteria (Steinberg and De Nys, 2002; de Carvalho, 2018) mostly due to the accumulation of nutrients and macroalgal exudates composed of organic carbon and nitrogen particles (Armstrong et al., 2001; Haas and Wild, 2010; Dang and Lovell, 2016). Consequently, the competition among marine microorganisms to reserve a space within the biofilm community is tremendously intense and bacterial strains that are equipped with broad-spectrum inhibitory phenotypes are likely to be successful epibiotic colonizers (Figure 1.4) (Rao et al., 2005; Thomas et al., 2008).

In addition, predation by heterotrophic protozoa and bacterivorous nematodes also represent as another biotic stress resulting in major mortality for both planktonic and surface associated bacteria in the marine habitat (Figure 1.4) (Matz et al., 2008; de Carvalho, 2018). Protozoans for example Rhynchomonas nasuta and Cafeteria roenbergensis are among the most abundant ubiquitous species in the ocean and the major controller of the food web in the marine environment through their function as bacterial predators (Moestrup, 2000; Hisatugo et al., 2014; De Corte et al., 2019). Whereas nematodes such as Pareudiplogaster pararmatus are among the natural

23 consumers of organic biomass in benthic habitats actively grazing bacterial mats and the biofilms of biotic surfaces (Moens et al., 2006; Weitere et al., 2018).

The omnipresence of inter- and intra-species interactions supports the evolution of diverse defence strategies by surface associated marine bacteria. Such defence mechanisms include the production of diverse industrially or pharmacologically important bioactive compounds showing antibacterial, antifungal, antitumor, antifouling, antiprotozoal and antinematode activities (Table 1.2) (Penesyan et al., 2010; Adnan et al., 2018). Interestingly, the physical-chemical properties, molecular structure and functional features of those marine microbial compounds are believed to be shaped by the naturally harsh conditions of marine environment (Rocha-Martin et al., 2014). Moreover, it is speculated that bioactive metabolites originally attributed to marine such as sponges, tunicates, bryozoans and molluscs are actually produced by their associated microorganisms (Proksch et al., 2002; Thomas et al., 2010). For example, the antibiotics peptide andrimid and trisindoline isolated from the sponge Hyatella sp and Hyrtios altum are believed to be produced by symbiotic Vibrio sp. (Kobayashi et al., 1994; Oclarit et al., 1994). An antitumor cyclic peptide leucamide A isolated from the sponge Leucetta microraphis is closely related to compounds produced by cyanobacterial symbionts (König et al., 2006). In addition, a commercialised antitumor drug Didemnin B initially isolated from the tunicate Trididemnum solidum (Rinehart et al., 1981) was recently demonstrated to be produced by symbioic bacteria Tistrella mobilis and T. bauzanensis (Tsukimoto et al., 2011; Xu et al., 2012). These observations also hold true for numerous antinematode compounds that have been successfully isolated from the marine eukaryotes (Table 1.3), with production of many now attributed to host-associated microorganisms (Romano et al., 2017; Sekurova et al., 2019).

The potential for marine surface associated microorganisms to be repositories of unusual gene functions and bioactivities is further corroborated by global biodiversity studies such as the Tara (Sunagawa et al., 2015; Zhang et al., 2019) and the Global Ocean Sampling (GOS) expedition (Venter et al., 2004; Rusch et al., 2007). These and other studies continue to reveal unprecedented level of information on

24 microbial diversity, which has subsequently led to investigations on underexplored bacterial diversity in various marine ecosystems. Furthermore, abundant and novel biosynthetic gene clusters encoding rare non-ribosomal peptides (NRPS), polyketides (PKS), and NRPS-PKS hybrids and CRISPR-Cas systems were discovered from marine biofilm samples (Zhang et al., 2019), indicating the potential of marine environment as a new source of extraordinary bioactive compounds for industries and novel drug development. (Blockley et al., 2017).

25

Figure 1.4 Surface associated marine bacteria live on the nutrient-rich marine surfaces for example macroalgae or cnidarian in a form of biofilms. However, the marine biofilms are exposed to biotic (intra and/or interspecies interaction with other microorganisms or predators i.e. protozoa and nematodes) and abiotic physical-chemical stressors. The predator-prey interaction leads to the production of novel nematicidal metabolites by the surface associated marine bacteria while the harsh environmental condition enhances the chemical, molecular and functional property of the produced microbial compounds. Figure was designed using Biorender at https://biorender.com/

26

Table 1.2 Examples of anthelmintic or nematicidal bioactivities isolated from marine bacteria

Associated Marine microbial producer Compound Mode of action Responsible gene(s) Reference surface/host Microbulbifer sp. D250 Violacein Delisea Facilitate bacterial accumulation VioA-VioE (Ballestriero et al., pulchra accompanied by tissue damage 2014) and

Pseudoalteromonas tunicata Tambjamine Algae Ulva Slow-killing activity by a heat TamA-TamT (Ballestriero et al., D2 australis resistance tambjamine and 2010) substantial bacterial colonisation in the nematode gut

Pseudoalteromonas tunicata Unknown Algae Ulva Fast-killing activity by a heat Unknown (Ballestriero et al., D2 australis sensitive unknown compound 2010) through colonisation independent manner Uncultured alpha- Unknown Algae Ulva Undetermined Possibly NRPS genes (Penesyan et al., 2013) proteobacterium, JN874385 australis cluster (strain U95)

Pseudovibrio sp. Pv348, Unknown Algae Delisea Undetermined Unknown (Penesyan et al., 2013) 1413, HE818384 (strain pulchra D323) Heterologous clone jj117 Unknown Ulva australis Undetermined ATP-grasp (Nappi, 2019) (NCBI Accession number metagenomic protein/alpha-E SRX4339430) library protein/transglutamina se protein/protease

27

Table 1.2 (Continue) Examples of anthelmintic or nematicidal bioactivities isolated from marine bacteria

Responsible Marine microbial producer Compound Associated surface/host Mode of action Reference gene(s) Aequorivita sp. Unknown Antartic marine sediment Undetermined Unknown (Palma Esposito et al., 2018) Vibrio atlanticus strain S- Volatile organic Scallop Argopecten irradians Undetermined Unknown (Yu et al., 2015) 16 compounds (VOC)

Pseudoalteromonas rubra Unknown Marine organisms (copepod or Undetermined Unknown (Gram et al., 2010; Neu et fish) or environmental samples al., 2014)

Pseudoalteromonas Unknown Marine organisms (copepod or Undetermined Unknown (Gram et al., 2010; Neu et piscicida fish) or environmental samples al., 2014)

Arthrobacter davidanieli Unknown Marine environmental samples Undetermined Unknown (Wietz et al., 2012; Neu et al., 2014) Pseudoalteromonas Unknown Marine organisms (copepod or Undetermined Unknown (Gram et al., 2010; Neu et luteoviolacea fish) or environmental samples al., 2014)

Photobacterium Unknown Marine organisms (copepod or Undetermined Unknown (Gram et al., 2010; Neu et halotolerans fish) or environmental samples al., 2014)

28

Table 1.3 Examples of anthelmintic or nematicidal bioactivities isolated from marine eukaryotic organisms

Marine eukaryotic producer Compound Mode of Action Tested nematode Reference Sponge Trachycladus Onnamide F Undetermined Haemonchus contortus (Vuong et al., 2001) laevispirulifer Sponge Oceanapia sp. Thiocyanatins Undetermined Haemonchus contortus (Capon et al., 2004) Sponge Echinodictyum sp (−)-echinobetaine A Undetermined Haemonchus contortus (Capon et al., 2005)

Sponge Phoriospongia sp. Phorioadenine A Undetermined Haemonchus contortus (Farrugia et al., 2014)

Sponge Jaspis sp Jasplakinolide Undetermined Nippostrongylus brasiliensis (Crews et al., 1986) Sponge Monanchora Fromiamycalin and Inhibition of nematode motility and Haemonchus contortus (Herath et al., 2019) unguiculata larval development Sponge Haliclona sp Halaminol A Inhibition of nematode motility and Haemonchus contortus (Herath et al., 2019) larval development Algae Laurencia scoparia Terpene Undetermined Nippostrongylus brasiliensis (Davyt et al., 2006) Algae Chondria atropurpurea Chondriamide C Undetermined Nippostrongylus brasiliensis (Davyt et al., 1998) Algae Notheia anomala Tetrahydrofurans Inhibition of parasites eggs Haemonchus contortus and (Capon et al., 1998) development to the third stage of Trichostrongylus colubriformis infective larvae Bryozoan Bugula neritina Bryostatin-1 Causing defective morphological Syphacia muris (Harras et al., 2017) changes of parasite mouth, anus and cuticle

29

1.9 Pseudoalteromonas tunicata; a model of surface associated marine bacterium

P. tunicata is a dark-green-pigmented epiphytic bacterium with strong antifouling activity that was originally isolated from an adult tunicate Ciona intestinalis on western coast of Sweden (Holmstrom et al., 1998). Additional strains of P. tunicata exhibiting the same potent antifouling properties have subsequently been isolated from the surface of the green seaweed Ulva australis off the east coast of Australia (Egan et al., 2000). Furthermore, 16S rRNA gene sequences corresponding to P. tunicata have been found in sea ice samples from Antarctic and Artic (Brown and Bowman, 2001), from the Nullarbor Caves, Australia (Holmes et al., 2001) and in association with dinoflagellate blooms in a temperate Australian Estuary (Skerratt et al., 2002). Likewise, based on real-time quantitative PCR (RTQ-PCR) and genus-specific denaturing gradient gel electrophoresis (PCR-DGGE), P. tunicata has been detected colonising the surface of green alga Ulvaria fusca, Ulva lactuca as well as the tunicate C. intestinalis sampled around Aarhus, Denmark (Skovhus et al., 2004; Skovhus et al., 2007). Taxonomically related strains were also found on diverse marine surfaces in Sydney Harbour, Australia (Wilson et al., 2010). Taken together, these studies indicate that P. tunicata is likely to be ubiquitous in the marine environment and predominantly found in association with surfaces.

P. tunicata has arguably become one of the best studied marine surface associated bacterial species due to its production of bioactive compounds with broad spectrum antifouling and antimicrobial activities. These include a purple pigmented indole- based violacein (Matz et al., 2008), tambjamine YP1 (Franks et al., 2005), autolytic protein Pseudoalteromonas (AlpP) (James et al., 1996), formyl-tyrosine compounds with unknown bioactivity (Blasiak and Clardy, 2010) and uncharacterised antidiatom (Holmström et al., 1996; Kumar et al., 2011), antialgal, antilarval (Egan et al., 2001) and fast killing antinematode compounds (Ballestriero et al., 2010) (Figure 1.5).

30

Figure 1.5 Bioactive compounds produced by the surface associated marine bacterium Pseudoalteromonas tunicata D2. The uncharacterised antinematode compound (indicated by asterisks and coloured in blue box) showing fast killing activity against C. elegans is the subject of this thesis. Figure was designed using Biorender at https://biorender.com/

31

Violacein, is produced by P. tunicata and other pigmented bacterial species such as Chromobacterium violaceum (Lozano et al., 2020), Pseudoalteromonas byunsanensis, P. shioyasakiensis, P. arabiensis, P. gelatinilytica (Wu et al., 2017) and Microbulbifer sp. (Won et al., 2017). In P. tunicata, the expression of violacein corresponds to its potent antiprotozoal activity against different species of flagellates, ciliates and amoebas (Matz et al., 2008). The expression of violacein is encoded by an 8 kb biosynthetic gene cluster containing five open reading frames (ORFs) (Matz et al., 2008) showing strong sequence similarity to violacein operon vioABCDE in C. violaceum (August et al., 2000; Matz et al., 2008; Kothari et al., 2017). Antiprotozoal effect of violacein produced by P. tunicata has been studied through transposon mutation of the vioA gene which results in defective violacein biosynthesis allowing subsequent degradation of P. tunicata mutant (∆vioA) biofilm by flagellate Rynchomonas nasuta (Matz et al., 2008). Matz et al. observed that violacein accumulates in the periplasm and outer membrane of P. tunicata and proposed that “eating” P. tunicata causes apoptosis and inevitable death of the protozoa (Matz et al., 2008). Violacein production appears to be more prominent in P. tunicata biofilms relative to planktonic growth, which likely represents the underlying mechanism behind biofilm resistance to predatory protozoans (Matz et al., 2008). In addition to the antiprotozoal effect, violacein also shows other bioactivities including antiviral (Andrighetti-Fröhner et al., 2003), anti-dinoflagellates (Matz et al., 2004), antitumor (Queiroz et al., 2012), antinematode (Ballestriero et al., 2014), antibacterial (Aruldass et al., 2018) and antifungal (Sasidharan et al., 2015).

Tambjamine YP1 produced by P. tunicata is a 4-methoxypyrrole molecule possessing a 2,2’-bipyrrole ring system (Franks et al., 2005; Burke et al., 2007; Marchetti et al., 2018). This small molecule (356 Da) is responsible for the inhibitory activity against several fungal and yeast strains (Franks et al., 2006) and in part the antinematode activity of P. tunicata against C. elegans (Ballestriero et al., 2010). Genomic analyses of P. tunicata revealed a 25 kb genomic region containing 21 open reading frames (ORFs) where 19 of the ORFs were designated as core components of the Tam cluster (TamA to TamT) which is responsible for the biosynthesis of the 4-methoxy-2-2’- bipyrrole-5-carbaldehyde (MBC) structure (Burke et al., 2007). Another gene outside of the core Tam cluster; afaA (“antifungal activity A”) produces an acyl-CoA

32 synthetase, which is proposed to activate the long-chain fatty acids CoA thioester (Franks et al., 2006). Both 4-methoxy-2-2’-bipyrrole-5-carbaldehyde (MBC) structure and the long-chain fatty acids (dodec-3-en-1-amine) are condensed together by TamQ (a putative gene expressing terminal condensing enzyme) to become Tambjamine YP1 (Burke et al., 2007). In addition to its antifungal and antinematode effect, tambjamine has been reported possessing other medicinal important activities such as antitumor (Pinkerton et al., 2010) and antimalaria (Kancharla et al., 2015).

P. tunicata also produce AlpP, a novel 190 kDa protein with antibacterial and autolytic activity. This protein consists of two different subunits (60 kDa and 80 kDa) that are combined together by noncovalent bonds (James et al., 1996). A wide range of Gram- positive and Gram-negative bacteria from different habitats including marine, soil, gastrointestinal environment and even the stationary-phase of P. tunicata bacterial cells itself are sensitive to the antibacterial activity of AlpP (James et al., 1996; Mai- Prochnow et al., 2004). While the precise mode of action for AlpP is yet to be determined, the protein itself has homology to marinocine produced by the melanogenic marine bacterium; Marinomonas mediterranea (Lucas-Elío et al., 2006). Further analyses of AlpP and its homolog LodA from M. mediterranea (Lucas-Elio et al., 2005) have demonstrated that both proteins are able to catalyse the oxidation of L- lysine to produce hydrogen peroxide, which eventually causes cell death. Studies on biofilm formation and differentiation in P. tunicata have shown that the production of hydrogen peroxide by AlpP plays an important role in facilitating localised cell death (i.e. autolytic activity) and dispersal stage of the established biofilm (Mai-Prochnow et al., 2004; Mai-Prochnow et al., 2008). The AlpP mediated cell death and biofilm dispersal are thought to be an efficient strategy of P. tunicata to protect the host (i.e. U. lactuca and C. intestinalis) from unrestrained biofilm formation and fouling problem by the P. tunicata cell itself (Mai-Prochnow et al., 2004). As both macroalgae and tunicate hosts are reported to be free from fouling, the controllable biofilm formation by P. tunicata by AlpP is thought to be important in addition to the secretion of antifouling metabolites by P. tunicata, thereby preventing other microfouler settlement (Holmström et al., 1992).

33

P. tunicata cells also show potent growth inhibition against a number of marine benthic diatoms including Amphora sp. (Holmström et al., 1996) and Cylindrotheca fusiformis (Kumar et al., 2011). While the exact nature of the compound responsible for this activity remains unknown, preliminary evidence suggests the involvement of a non-ribosomal peptide (Stelzer et al., 2006; Kumar, 2008). Transposon mutagenesis of P. tunicata identified a strain (designated DM4) that was no longer able to inhibit growth and was found to be disrupted in a gene (PTD2_0056) located within a large (64 kb) putative non-ribosomal peptide synthetase (NRPS) gene cluster (Kumar, 2008). NRPSs are large multienzyme complexes with assembly-line organisations, which function independently from ribosomal pathway during protein construction. NRPS clusters are found in a range of bacterial and fungal genomes and are responsible for producing peptides with a broad range of activities including antibiotic, immunomodulating and antitumor activities (Caboche et al., 2010; Klapper et al., 2018; Rischer et al., 2018).

A small (3 to 10 kDa), extracellular and polar compound which demonstrates inhibitory activity against algal spores (green alga Ulva lactuca and red alga Polysiphonia sp.) is another bioactive metabolite produced by P. tunicata (Egan et al., 2001). This uncharacterised antialgal compound is water soluble and is believed to be extracellularly secreted by P. tunicata, hence preventing settlement and germination of any potential macroalgal spore on the host surface. A polar antilarval compound (< 500 Da in size) is also produced by P. tunicata which is toxic against barnacle larvae Balanus Amphitrite and ascidian larvae Ciona intestinalis. However, until now, both the active molecule and the genes responsible for the activities remain unknown.

Research conducted in the laboratory of Blasiak and Clardy (Blasiak and Clardy, 2010) resulted in the discovery of two new P. tunicata metabolites identified as 3- formyl-L-tyrosine-L-threonine dipeptide (C14H19N2O6) and 3-formyl-L-tyrosine, both with yet unknown functions. Both compounds are derived from the expression of biosynthetic genes cluster in P. tunicata during the search of sequence homologous to ATP-grasp enzymes in the marine bacterium. A cluster of fty (formyl tyrosine) genes were found on the genome of P. tunicata where the ATP-grasp enzyme FtyB in P. tunicata genome exhibited 30% of similarity to an ATP-grasp-type amide ligase in

34 dapdiamide biosynthesis (DdaF) found in the human pathogen Pantoe agglomerans (Delétoile et al., 2009; Hollenhorst et al., 2009). These new compounds do not possess inhibitory effect against Escherichia coli, Bacillus subtilis and Saccharomyces cerevisiae, however are speculated to have antihypertensive and appetite suppressant properties based on characteristic depicted by similar synthetic compound (Blasiak and Clardy, 2010). Until now, the gene regulation and biosynthetic pathway remain unclear. Ongoing analysis would determine the specific features of these newly isolated compounds.

Additionally, high throughput screening of genomic clone libraries against the model nematode C. elegans revealed two antinematode clones with inserts derived from the genome of P. tunicata (Ballestriero et al., 2010). The activity of one of these antinematode clones (E. coli AA1) was attributed to the yellow pigment tambjamine YP1 (as discussed above) and resulted in 70% death of C. elegans within 5 days exposure to the tambjamine producing clone. Whereas exposure of C. elegans to the second E. coli clone (designated as HG8) killed 70% of the nematodes only within 24 hours and was therefore designated the fast killing activity (Ballestriero et al., 2010). Unlike the heat-stable tambjamine compound produced by the AA11 clone, the cytotoxicity effect of antinematode activity produced by HG8 diminishes after one hour of 65o-heat-treatment, suggesting the denaturation of the unidentified antinematode activity in high environmental temperature (Ballestriero et al., 2010). In order to characterise the nematicidal activity of clone HG8, transposon mutagenesis on the HG8 fosmid was performed and successfully identified a mutant designated as H10 fosmid mutant exhibiting attenuated antinematode activity, resulting from the disruption of an uncharacterised gene (NCBI GenBank accession number ZP_01132244) (Ballestriero et al., 2010). Complementation of mutant Tn::ZP_01132244 with a plasmid carrying P. tunicata wild type gene ZP_01132244 resulted in the restoration of fast killing antinematode activity (Ballestriero et al., 2010). However, expression of ZP_01132244 gene alone did not result in antinematode activity (Ballestriero et al., unpublished data), suggesting other gene(s) located elsewhere on the HG8 fosmid were required for this antinematode activity.

35

1.10 Aims of study

The increasing prevalence of drug-resistant parasitic nematode infection in humans, animals and plants has been recognised as a global health issue. Current control strategies on parasitic nematodes predominantly rely on drug treatments. Unfortunately, genetic mutations develop in those parasites due to prolonged exposure to a single drug and inappropriate dosage treatment, rendering the parasite drug- resistant. New antinematode drugs are urgently needed, however, the re-discovery rate of bioactive compounds from the terrestrial environment is high and represents a bottleneck for novel antinematode drug development. Owing to the highly diverse macro and microorganisms, the marine environment and particularly surface associated bacteria, provide a promising habitat for antinematode compound discovery and drug research. The surface associated marine bacterium Pseudoalteromonas tunicata D2 has arguably become one of the best studied marine surface associated bacterial species due to the production of compounds with commercially and pharmaceutically-related functions (see section 1.9, Figure 1.5) including the fast killing antinematode activity (Ballestriero et al., 2010). However, this antinematode activity including its mode of action remains uncharacterised. Given the maintenance of parasitic nematodes in the laboratory is challenging and hazardous, Caenorhabditis elegans represent as a powerful model to be used for antinematode compound identification and characterisation. Furthermore, given soil-dwelling and host- associated parasitic nematodes are associated with a natural gut microbiota, utilisation of C. elegans associated with its natural gut microbiota is likely to provide greater insight into the efficacy of antinematode compounds under near-natural conditions.

Therefore, this research project was designed to provide fundamental understanding of the characteristics of the novel fast killing antinematode activity of P. tunicata D2, its mode of action and the impact of gut microbiota association on C. elegans survival upon exposure to the compound toxicity. These goals were achieved using a previously characterised E. coli clone HG8 carrying the 13.8 kb P. tunicata D2 DNA insert that showed the fast killing antinematode activity against C. elegans. The specific aims of this research project were;

36

1.0 To characterise and identify the bioactive agent responsible for antinematode activity of the clone HG8 (Chapter 2).

2.0 To elucidate the mode of action (MOA) employed by the nematicidal agent producing E. coli clones i.e. HP1 and HG8 against C. elegans and to determine any cellular damage, behavioural and/or immunity responses of the nematode as a result of the clones’ toxic activity (Chapter 3).

3.0 To establish a native gut microbiota in C. elegans N2-laboratory cultures and subsequently determine what impact this microbiota has on the sensitivity of C. elegans to the E. coli clones HP1 and HG8 (Chapter 4).

4.0 To summarise and provide comprehensive discussion of the major discoveries in this study and highlight the future perspective for subsequent research development (Chapter 5).

37

1.11 REFERENCES

Abad, P., Gouzy, J., Aury, J.-M., Castagnone-Sereno, P., Danchin, E. G. J., Deleury, E., Perfus-Barbeoch, L., Anthouard, V., Artiguenave, F., Blok, V. C., Caillaud, M.- C., Coutinho, P. M., Dasilva, C., De Luca, F., Deau, F., Esquibet, M., Flutre, T., Goldstone, J. V., Hamamouch, N., Hewezi, T., Jaillon, O., Jubin, C., Leonetti, P., Magliano, M., Maier, T. R., Markov, G. V., McVeigh, P., Pesole, G., Poulain, J., Robinson-Rechavi, M., Sallet, E., Ségurens, B., Steinbach, D., Tytgat, T., Ugarte, E., van Ghelder, C., Veronico, P., Baum, T. J., Blaxter, M., Bleve-Zacheo, T., Davis, E. L., Ewbank, J. J., Favery, B., Grenier, E., Henrissat, B., Jones, J. T., Laudet, V., Maule, A. G., Quesneville, H., Rosso, M.-N., Schiex, T., Smant, G., Weissenbach, J., & Wincker, P. (2008). Genome sequence of the metazoan plant- parasitic nematode Meloidogyne incognita. Nature Biotechnology, 26(8), 909- 915. Aballay, A., Drenkard, E., Hilbun, L. R., & Ausubel, F. M. (2003). Caenorhabditis elegans innate immune response triggered by Salmonella enterica requires intact LPS and is mediated by a MAPK signaling pathway. Current Biology, 13(1), 47- 52. Abongwa, M., Martin, R. J., & Robertson, A. P. (2017). A brief review on the mode of action of antinematodal drugs. Acta Veterinaria, 67(2), 137-152. Adnan, M., Alshammari, E., Patel, M., Ashraf, S. A., Khan, S., & Hadi, S. (2018). Significance and potential of marine microbial natural bioactive compounds against biofilms/biofouling: necessity for green chemistry. PeerJ, 6, e5049. Aibinu, I. E., Smooker, P. M., & Lopata, A. L. (2019). Anisakis nematodes in fish and shellfish- from infection to allergies. International Journal for Parasitology. Parasites and wildlife, 9, 384-393. Akinsolu, F. T., Nemieboka, P. O., Njuguna, D. W., Ahadji, M. N., Dezso, D., & Varga, O. (2019). Emerging resistance of neglected tropical diseases: a scoping review of the literature. International Journal of Environmental Research and Public Health, 16(11). Amenya, D. A., Naguran, R., Lo, T. C., Ranson, H., Spillings, B. L., Wood, O. R., Brooke, B. D., Coetzee, M., & Koekemoer, L. L. (2008). Over expression of a cytochrome P450 (CYP6P9) in a major African malaria vector, Anopheles funestus, resistant to pyrethroids. Insect Molecular Biology, 17(1), 19-25. Andrighetti-Fröhner, C., Antonio, R., Creczynski-Pasa, T., Barardi, C., & Simões, C. (2003). Cytotoxicity and potential antiviral evaluation of violacein produced by Chromobacterium violaceum. Memórias do Instituto Oswaldo Cruz, 98(6), 843- 848. Angstadt, J. D., Donmoyer, J. E., & Stretton, A. O. (1989). Retrovesicular ganglion of the nematode Ascaris. The Journal of Comparative Neurology, 284(3), 374-388. Antunes, J., Leão, P., & Vasconcelos, V. (2019). Marine biofilms: Diversity of communities and of chemical cues. Environmental Microbiology Reports, 11(3), 287-305. Armstrong, E., Yan, L., Boyd, K. G., Wright, P. C., & Burgess, J. G. (2001). The symbiotic role of marine microbes on living surfaces. Hydrobiologia, 461(1-3), 37-40.

38

Artal‐Sanz, M., de Jong, L., & Tavernarakis, N. (2006). Caenorhabditis elegans: a versatile platform for drug discovery. Biotechnology Journal: Healthcare Nutrition Technology, 1(12), 1405-1418. Aruldass, C. A., Masalamany, S. R. L., Venil, C. K., & Ahmad, W. A. (2018). Antibacterial mode of action of violacein from Chromobacterium violaceum UTM5 against Staphylococcus aureus and methicillin-resistant Staphylococcus aureus (MRSA). Environmental Science and Pollution Research, 25(6), 5164- 5180. Ashford, R., & Crewe, W. (2003). Parasites of Homo sapiens: An Annotated Checklist of the Protozoa, Helminths and Arthropods for which we are Home: CRC Press. Askary, T. H., & Martinelli, P. R. P. (2015). Biocontrol agents of phytonematodes: CABI. August, P., Grossman, T., Minor, C., Draper, M., MacNeil, I., Pemberton, J., Call, K., Holt, D., & Osburne, M. (2000). Sequence analysis and functional characterization of the violacein biosynthetic pathway from Chromobacterium violaceum. Journal of Molecular Microbiology and Biotechnology, 2(4), 513-519. Bai, Y., Zhi, D., Li, C., Liu, D., Zhang, J., Tian, J., Wang, X., Ren, H., & Li, H. (2014). Infection and immune response in the nematode Caenorhabditis elegans elicited by the phytopathogen Xanthomonas. Journal of Microbiology and Biotechnology, 24(9), 1269-1279. Ballestriero, F., Daim, M., Penesyan, A., Nappi, J., Schleheck, D., Bazzicalupo, P., Di Schiavi, E., & Egan, S. (2014). Antinematode activity of Violacein and the role of the insulin/IGF-1 pathway in controlling violacein sensitivity in Caenorhabditis elegans. PloS One, 9(10), e109201-e109201. Ballestriero, F., Thomas, T., Burke, C., Egan, S., & Kjelleberg, S. (2010). Identification of compounds with bioactivity against the nematode Caenorhabditis elegans by a screen based on the functional genomics of the marine bacterium Pseudoalteromonas tunicata D2. Applied and Environmental Microbiology, 76(17), 5710-5717. Bañuls, A.-L., Thomas, F., & Renaud, F. (2013). Of parasites and men. Infection, Genetics and Evolution, 20, 61-70. Bar-On, Y. M., Phillips, R., & Milo, R. (2018). The biomass distribution on Earth. Proceedings of the National Academy of Sciences, USA 115(25), 6506-6511. Basyoni, M. M., & Rizk, E. M. (2016). Nematodes ultrastructure: complex systems and processes. Journal of Parasitic Diseases, 40(4), 1130-1140. Battisti, J. M., Watson, L. A., Naung, M. T., Drobish, A. M., Voronina, E., & Minnick, M. F. (2017). Analysis of the Caenorhabditis elegans innate immune response to Coxiella burnetii. Innate immunity, 23(2), 111-127. Bérdy, J. (2005). Bioactive microbial metabolites. The Journal of Antibiotics, 58(1), 1- 26. Berg, M., Stenuit, B., Ho, J., Wang, A., Parke, C., Knight, M., Alvarez-Cohen, L., & Shapira, M. (2016). Assembly of the Caenorhabditis elegans gut microbiota from diverse soil microbial environments. The ISME Journal, 10(8), 1998-2009. Berman, K., McKay, J., Avery, L., & Cobb, M. (2001). Isolation and characterization of pmk-(1-3): three p38 homologs in Caenorhabditis elegans. Molecular Cell Biology Research Communications : MCBRC, 4(6), 337-344. Bernard, G. C., Egnin, M., & Bonsi, C. (2017). The impact of plant-parasitic nematodes on agriculture and methods of control. In Shah, M. M. and Mahamood, M. (Ed.), Nematology-Concepts, Diagnosis and Control, IntechOpen, doi: 10.5772/intechopen.68958.

39

Besier, R., & Love, S. (2003). Anthelmintic resistance in sheep nematodes in Australia: the need for new approaches. Australian Journal of Experimental Agriculture, 43(12), 1383-1391. Beuvry, V., Bufala, G., & Derrieu, G. (2003). Injectable anthelmintic compositions and methods for using same. U.S. Patent No. US6653288B1. Washington, DC: U.S. Patent and Trademark Office. Bik, H. M., Lambshead, P. J. D., Thomas, W. K., & Lunt, D. H. (2010). Moving towards a complete molecular framework of the Nematoda: a focus on the and early-branching clades. BMC Evolutionary Biology, 10(1), 353. Blasco-Costa, I., & Poulin, R. (2017). Parasite life-cycle studies: a plea to resurrect an old parasitological tradition. Journal of Helminthology, 91(6), 647-656. Blasiak, L. C., & Clardy, J. (2010). Discovery of 3-formyl-tyrosine metabolites from Pseudoalteromonas tunicata through heterologous expression. Journal of the American Chemical Society, 132(3), 926-927. Blaxter, M., & Koutsovoulos, G. (2015). The evolution of parasitism in Nematoda. Parasitology, 142(S1), S26-S39. Blockley, A., Elliott, D. R., Roberts, A. P., & Sweet, M. (2017). Symbiotic microbes from marine invertebrates: driving a new era of natural product drug discovery. Diversity, 9(4), 49. Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics, 77(1), 71-94. Brown, M. V., & Bowman, J. P. (2001). A molecular phylogenetic survey of sea-ice microbial communities (SIMCO). FEMS Microbiol Ecology, 35(3), 267-275. Bryant, A. S., Ruiz, F., Gang, S. S., Castelletto, M. L., Lopez, J. B., & Hallem, E. A. (2018). A critical role for thermosensation in host seeking by skin-penetrating nematodes. Current Biology, 28(14), 2338-2347.e2336. Buchmann, K., & Bjerregaard, P. (1990). Effect of ivermectin, pyrantel and morantel on the European eel and its monogenean parasites. Bulletin-European Association of Fish Pathologists, 10(5), 146-148. Bundy, D. A., de Silva, N., Appleby, L. J., & Brooker, S. J. (2020). Intestinal Nematodes: Ascariasis. In Hunter's Tropical Medicine and Emerging Infectious Diseases (pp. 840-844): Elsevier. Burg, R. W., Miller, B. M., Baker, E. E., Birnbaum, J., Currie, S. A., Hartman, R., Kong, Y. L., Monaghan, R. L., Olson, G., Putter, I., Tunac, J. B., Wallick, H., Stapley, E. O., Oiwa, R., & Omura, S. (1979). Avermectins, new family of potent anthelmintic agents: producing organism and fermentation. Antimicrobial Agents and Chemotherapy, 15(3), 361-367. Burke, C., Thomas, T., Egan, S., & Kjelleberg, S. (2007). The use of functional genomics for the identification of a gene cluster encoding for the biosynthesis of an antifungal tambjamine in the marine bacterium Pseudoalteromonas tunicata. Environmental Microbiology, 9(3), 814-818. Burns, A. R., Luciani, G. M., Musso, G., Bagg, R., Yeo, M., Zhang, Y., Rajendran, L., Glavin, J., Hunter, R., Redman, E., Stasiuk, S., Schertzberg, M., Angus McQuibban, G., Caffrey, C. R., Cutler, S. R., Tyers, M., Giaever, G., Nislow, C., Fraser, A. G., MacRae, C. A., Gilleard, J., & Roy, P. J. (2015). Caenorhabditis elegans is a useful model for anthelmintic discovery. Nature Communications, 6, 7485-7485. Caboche, S., Leclère, V., Pupin, M., Kucherov, G., & Jacques, P. (2010). Diversity of monomers in nonribosomal peptides: towards the prediction of origin and biological activity. Journal of Bacteriology, 192(19), 5143.

40

Capon, R. J., A Barrow, R., Rochfort, S., Jobling, M., Skene, C., Lacey, E., H Gill, J., Friedel, T., & Wadsworth, D. (1998). Marine nematocides: Tetrahydrofurans from a southern Australian brown alga, Notheia anomala. Tetrahedron, 54(10), 2227-2242. Capon, R. J., Skene, C., Liu, E. H.-T., Lacey, E., Gill, J. H., Heiland, K., & Friedel, T. (2004). Nematocidal thiocyanatins from a southern Australian marine sponge Oceanapia sp. Journal of Natural Products, 67(8), 1277-1282. Capon, R. J., Vuong, D., McNally, M., Peterle, T., Trotter, N., Lacey, E., & Gill, J. H. (2005). (+)-Echinobetaine B: isolation, structure elucidation, synthesis and preliminary SAR studies on a new nematocidal betaine from a southern Australian marine sponge, Echinodictyum sp. Organic & Biomolecular Chemistry, 3(1), 118- 122. Cezairliyan, B., Vinayavekhin, N., Grenfell-Lee, D., Yuen, G. J., Saghatelian, A., & Ausubel, F. M. (2013). Identification of Pseudomonas aeruginosa phenazines that kill Caenorhabditis elegans. PLoS Pathogens, 9(1), e1003101. Chávez, V., Mohri-Shiomi, A., Maadani, A., Vega, L. A., & Garsin, D. A. (2007). Oxidative stress enzymes are required for DAF-16-mediated immunity due to generation of reactive oxygen species by Caenorhabditis elegans. Genetics, 176(3), 1567-1577. Chen, L., Jiang, H., Cheng, Q., Chen, J., Wu, G., Kumar, A., Sun, M., & Liu, Z. (2015). Enhanced nematicidal potential of the chitinase pachi from Pseudomonas aeruginosa in association with Cry21Aa. Scientific Reports, 5, 14395. Chiu, J. M. Y., Thiyagarajan, V., Tsoi, M. M. Y., & Qian, P. Y. (2005). Qualitative and quantitative changes in marine biofilms as a function of temperature and salinity in summer and winter. Biofilms, 2(3), 183-195. Cobo, F. (2016). Determinants of parasite drug resistance in human lymphatic filariasis. Revista Española de Quimioterapia, 29(6), 288-295. Conlan, J. V., Khamlome, B., Vongxay, K., Elliot, A., Pallant, L., Sripa, B., Blacksell, S. D., Fenwick, S., & Thompson, R. C. (2012). Soil-transmitted helminthiasis in Laos: a community-wide cross-sectional study of humans and dogs in a mass drug administration environment. The American Journal of Tropical Medicine and Hygiene, 86(4), 624-634. Costello, M. J., Cheung, A., & De Hauwere, N. (2010). Surface area and the seabed area, volume, depth, slope, and topographic variation for the world’s seas, oceans, and countries. Environmental Science & Technology, 44(23), 8821-8828. Cox, F. E. G. (2004). History of human parasitic diseases. Infectious Disease Clinics of North America, 18(2), 171-188. doi:https://doi.org/10.1016/j.idc.2004.01.001 Crews, P., Manes, L. V., & Boehler, M. (1986). Jasplakinolide, a cyclodepsipeptide from the marine sponge, Jaspis sp. Tetrahedron Letters, 27(25), 2797-2800. Dang, H., & Lovell, C. R. (2016). Microbial surface colonization and biofilm development in marine environments. Microbiology and Molecular Biology Reviews, 80(1), 91-138. Davis, R. (1998). Neurophysiology of glutamatergic signalling and anthelmintic action in Ascaris suum: pharmacological evidence for a kainate receptor. Parasitology, 116(5), 471-486. Davyt, D., Entz, W., Fernandez, R., Mariezcurrena, R., Mombrú, A. W., Saldana, J., Domínguez, L., Coll, J., & Manta, E. (1998). A new indole derivative from the red alga Chondria atropurpurea. Isolation, structure determination, and anthelmintic activity. Journal of Natural Products, 61(12), 1560-1563.

41

Davyt, D., Fernandez, R., Suescun, L., Mombrú, A. W., Saldaña, J., Domínguez, L., Fujii, M. T., & Manta, E. (2006). Bisabolanes from the Red Alga Laurencia scoparia. Journal of Natural Products, 69(7), 1113-1116. de Carvalho, C. C. (2018). Marine biofilms: a successful microbial strategy with economic implications. Frontiers in Marine Science, 5, 126. De Corte, D., Paredes, G., Yokokawa, T., Sintes, E., & Herndl, G. J. (2019). Differential response of Cafeteria roenbergensis to different bacterial and archaeal prey characteristics. Microbial Ecology, 78(1), 1-5. Delétoile, A., Decré, D., Courant, S., Passet, V., Audo, J., Grimont, P., Arlet, G., & Brisse, S. (2009). Phylogeny and identification of Pantoea Species and typing of Pantoea agglomerans strains by multilocus gene sequencing. Journal of Clinical Microbiology, 47(2), 300. Dell’Anno, A., Carugati, L., Corinaldesi, C., Riccioni, G., & Danovaro, R. (2015). Unveiling the biodiversity of deep-sea nematodes through metabarcoding: are we ready to bypass the classical taxonomy? PloS One, 10(12), e0144928. Devaraj, S., Amirthalingam, M., Sabapathy, P. C., Govindan, S., Palanisamy, S., & Kathirvel, P. (2019). Anthelmintic efficacy of glycolipid biosurfactant produced by Pseudomonas plecoglossicida: an insight from mutant and transgenic forms of Caenorhabditis elegans. Biodegradation, 30(4), 203-214. Diemert, D. J., Freire, J., Valente, V., Fraga, C. G., Talles, F., Grahek, S., Campbell, D., Jariwala, A., Periago, M. V., & Enk, M. (2017). Safety and immunogenicity of the Na-GST-1 hookworm vaccine in Brazilian and American adults. PLoS Neglected Tropical Diseases, 11(5), e0005574. Dirksen, P., Marsh, S. A., Braker, I., Heitland, N., Wagner, S., Nakad, R., Mader, S., , C., Kowallik, V., Rosenstiel, P., Félix, M.-A., & Schulenburg, H. (2016). The native microbiome of the nematode Caenorhabditis elegans: gateway to a new host-microbiome model. BMC Biology, 14(1), 38. Drake, J., & Carey, T. (2019). Seasonality and changing prevalence of common canine gastrointestinal nematodes in the USA. Parasites & Vectors, 12(1), 430. Driscoll, M., Dean, E., Reilly, E., Bergholz, E., & Chalfie, M. (1989). Genetic and molecular analysis of a Caenorhabditis elegans beta-tubulin that conveys benzimidazole sensitivity. Journal of Cell Biology, 109(6 Pt 1), 2993-3003. Dutta, T. K., Khan, M. R., & Phani, V. (2019). Plant-parasitic nematode management via biofumigation using brassica and non-brassica plants: Current status and future prospects. Current Plant Biology, 17, 17-32. Eberhard, M. L., Cleveland, C. A., Zirimwabagabo, H., Yabsley, M. J., Ouakou, P. T., & Ruiz-Tiben, E. (2016). Guinea Worm (Dracunculus medinensis) Infection in a wild-caught frog, Chad. Emerging Infectious Diseases, 22(11), 1961-1962. Egan, S., Harder, T., Burke, C., Steinberg, P., Kjelleberg, S., & Thomas, T. (2013). The seaweed holobiont: understanding seaweed–bacteria interactions. FEMS Microbiology Reviews, 37(3), 462-476. Egan, S., James, S., Holmstrom, C., & Kjelleberg, S. (2001). Inhibition of algal spore germination by the marine bacterium Pseudoalteromonas tunicata. FEMS Microbial Ecology, 35(1), 67-73. Egan, S., Thomas, T., Holmström, C., & Kjelleberg, S. (2000). Phylogenetic relationship and antifouling activity of bacterial epiphytes from the marine alga Ulva lactuca: brief report. Environmental Microbiology, 2(3), 343-347. Elfawal, M. A., Savinov, S. N., & Aroian, R. V. (2019). Drug screening for discovery of broad-spectrum agents for soil-transmitted nematodes. Scientific Reports, 9(1), 12347. 42

Elseadawy, R., Abbas, I., Al-Araby, M., Hildreth, M. B., & Abu-Elwafa, S. (2019). First evidence of Teladorsagia circumcincta infection in sheep from Egypt. Journal of Parasitology, 105(4), 484-490. Enejoh, O., & Suleiman, M. (2017). Anthelmintics and their application in veterinary medicine. Research in Medical & Engineering Sciences, 2(3). RMES.000536. Engelmann, I., & Pujol, N. (2010). Innate immunity in C. elegans. In Invertebrate Immunity (pp. 105-121): Springer. Ermolaeva, M. A., & Schumacher, B. (2014). Insights from the worm: the C. elegans model for innate immunity. Seminars in Immunology, 26(4), 303-309. Evans, E. A., Chen, W. C., & Tan, M. W. (2008). The DAF-2 insulin-like signaling pathway independently regulates aging and immunity in C. elegans. Aging Cell, 7(6), 879-893. Ezeamama, A. E., Friedman, J. F., Acosta, L. P., Bellinger, D. C., Langdon, G. C., Manalo, D. L., Olveda, R. M., Kurtis, J. D., & McGarvey, S. T. (2005). Helminth infection and cognitive impairment among Filipino children. The American Journal of Tropical Medicine and Hygiene, 72(5), 540-548. Farrugia, M., Trotter, N., Vijayasarathy, S., Salim, A. A., Khalil, Z. G., Lacey, E., & Capon, R. J. (2014). Isolation and synthesis of N-acyladenine and adenosine alkaloids from a southern Australian marine sponge, Phoriospongia sp. Tetrahedron Letters, 55(43), 5902-5904. Fenical, W., & Jensen, P. R. (2006). Developing a new resource for drug discovery: marine actinomycete bacteria. Nature Chemical Biology, 2(12), 666-673. Ferris, H. (2010). Contribution of nematodes to the structure and function of the soil food web. Journal of Nematology, 42(1), 63-67. Flemming, H.-C., & Wuertz, S. (2019). Bacteria and archaea on Earth and their abundance in biofilms. Nature Reviews Microbiology, 17(4), 247-260. Fox, N. J., Marion, G., Davidson, R. S., White, P. C., & Hutchings, M. R. (2015). Climate- driven tipping-points could lead to sudden, high-intensity parasite outbreaks. Royal Society Open Science, 2(5), 140296. Fraietta, I., & Gasparri, F. (2016). The development of high-content screening (HCS) technology and its importance to drug discovery. Expert Opinion on Drug Discovery, 11(5), 501-514. Franks, A., Egan, S., Holmström, C., James, S., Lappin-Scott, H., & Kjelleberg, S. (2006). Inhibition of fungal colonization by Pseudoalteromonas tunicata provides a competitive advantage during surface colonization. Applied and Environmental Microbiology, 72(9), 6079-6087. Franks, A., Haywood, P., Holmstrom, C., Egan, S., Kjelleberg, S., & Kumar, N. (2005). Isolation and structure elucidation of a novel yellow pigment from the marine bacterium Pseudoalteromonas tunicata. Molecules, 10(10), 1286-1291. Frean, J. (2020). Gnathostomiasis acquired by visitors to the Okavango Delta, Botswana. Tropical Medicine and Infectious Disease, 5(1), 39. Frézal, L., & Félix, M.-A. (2015). C. elegans outside the Petri dish. eLife, 4, e05849. Gadisa, E., & Jote, K. (2019). Prevalence and factors associated with intestinal parasitic infection among under-five children in and around Haro Dumal Town, Bale Zone, Ethiopia. BMC Pediatrics, 19(1), 385. Galbadage, T., Shepherd, T. F., Cirillo, S. L. G., Gumienny, T. L., & Cirillo, J. D. (2016). The Caenorhabditis elegans p38 MAPK gene plays a key role in protection from mycobacteria. MicrobiologyOpen, 5(3), 436-452.

43

Garcia-Bustos, J. F., Sleebs, B. E., & Gasser, R. B. (2019). An appraisal of natural products active against parasitic nematodes of animals. Parasites & Vectors, 12(1), 306. Garsin, D. A., Villanueva, J. M., Begun, J., Kim, D. H., Sifri, C. D., Calderwood, S. B., Ruvkun, G., & Ausubel, F. M. (2003). Long-lived C. elegans daf-2 mutants are resistant to bacterial pathogens. Science, 300(5627), 1921. Gaudêncio, S. P., & Pereira, F. (2015). Dereplication: racing to speed up the natural products discovery process. Natural Product Reports, 32(6), 779-810. Gebremikael, M. T., Steel, H., Buchan, D., Bert, W., & De Neve, S. (2016). Nematodes enhance plant growth and nutrient uptake under C and N-rich conditions. Scientific Reports, 6, 32862-32862. Geerts, S., & Gryseels, B. (2000). Drug Resistance in Human Helminths: Current situation and lessons from livestock. Clinical Microbiology Reviews, 13(2), 207. Gerbaba, T. K., Green-Harrison, L., & Buret, A. G. (2017). Modeling host-microbiome interactions in Caenorhabditis elegans. Journal of Nematology, 49(4), 348-356. Gilleard, J. S. (2006). Understanding anthelmintic resistance: the need for genomics and genetics. International Journal for Parasitology, 36(12), 1227-1239. Giovanelli, F., Mattellini, M., Fichi, G., Flamini, G., & Perrucci, S. (2018). In vitro anthelmintic activity of four plant-derived compounds against sheep gastrointestinal nematodes. Veterinary Sciences, 5(3), 78. Goodman, M. B., & Sengupta, P. (2019). How Caenorhabditis elegans senses mechanical stress, temperature, and other physical stimuli. Genetics, 212(1), 25-51. Gram, L., Melchiorsen, J., & Bruhn, J. B. (2010). Antibacterial activity of marine culturable bacteria collected from a global sampling of ocean surface waters and surface swabs of marine organisms. Marine Biotechnology, 12(4), 439-451. Griffitts, J. S., Haslam, S. M., Yang, T., Garczynski, S. F., Mulloy, B., Morris, H., Cremer, P. S., Dell, A., Adang, M. J., & Aroian, R. V. (2005). Glycolipids as receptors for Bacillus thuringiensis crystal toxin. Science, 307(5711), 922-925. Gu, Y.-Q., Mo, M.-H., Zhou, J.-P., Zou, C.-S., & Zhang, K.-Q. (2007). Evaluation and identification of potential organic nematicidal volatiles from soil bacteria. Soil Biology and Biochemistry, 39(10), 2567-2575. Haas, A. F., & Wild, C. (2010). Composition analysis of organic matter released by cosmopolitan coral reef-associated green algae. Aquatic Biology, 10(2), 131-138. Haber, C. L., Heckaman, C. L., Li, G. P., Thompson, D. P., Whaley, H., & Wiley, V. (1991). Development of a mechanism of action-based screen for anthelmintic microbial metabolites with avermectinlike activity and isolation of milbemycin- producing Streptomyces strains. Antimicrobial Agents and Chemotherapy, 35(9), 1811-1817. Halaschek-Wiener, J., Khattra, J. S., McKay, S., Pouzyrev, A., Stott, J. M., Yang, G. S., Holt, R. A., Jones, S. J., Marra, M. A., & Brooks-Wilson, A. R. (2005). Analysis of long-lived C. elegans daf-2 mutants using serial analysis of gene expression. Genome Research, 15(5), 603-615. Haldeman, M. S., Nolan, M. S., & Ng'habi, K. R. N. (2020). Human hookworm infection: Is effective control possible? A review of hookworm control efforts and future directions. Acta Tropica, 201, 105214. Harras, S. F., Mona, M. H., Abdalla, A. G., & Salim, E. I. (2017). Effect of bryostatin-1 as an anthelmintic drug on adult Syphacia muris infection in rats. The Journal of Basic and Applied Zoology, 78(1), 8.

44

Hassan, M. A., Pham, T. H., Shi, H., & Zheng, J. (2013). Nematodes threats to global food security. Acta Agriculturae Scandinavica, Section B — Soil & Plant Science, 63(5), 420-425. Hasshoff, M., Böhnisch, C., Tonn, D., Hasert, B., & Schulenburg, H. (2007). The role of Caenorhabditis elegans insulin-like signaling in the behavioral avoidance of pathogenic Bacillus thuringiensis. The FASEB Journal, 21(8), 1801-1812. Hawdon, J. M. (2014). Controlling soil-transmitted helminths: time to think inside the box? Journal of Parasitology, 100(2), 166-188. Hendy, A., Krüger, A., Pfarr, K., De Witte, J., Kibweja, A., Mwingira, U., Dujardin, J.- C., Post, R., Colebunders, R., O’Neill, S., & Kalinga, A. (2018). The blackfly vectors and transmission of Onchocerca volvulus in Mahenge, south eastern Tanzania. Acta Tropica, 181, 50-59. Herath, H. M. P. D., Preston, S., Jabbar, A., Garcia-Bustos, J., Taki, A. C., Addison, R. S., Hayes, S., Beattie, K. D., McGee, S. L., Martin, S. D., Ekins, M. G., Hooper, J. N. A., Chang, B. C. H., Hofmann, A., Davis, R. A., & Gasser, R. B. (2019). Identification of Fromiamycalin and Halaminol A from Australian marine sponge extracts with anthelmintic activity against Haemonchus contortus. Marine Drugs, 17(11), 598. Hernando, G., Turani, O., & Bouzat, C. (2019). Caenorhabditis elegans muscle Cys-loop receptors as novel targets of terpenoids with potential anthelmintic activity. PLOS Neglected Tropical Diseases, 13(11), e0007895. Hisatugo, K. F., Mansano, A. S., & Seleghim, M. H. (2014). Protozoans bacterivory in a subtropical environment during a dry/cold and a rainy/warm season. Brazilian Journal of Microbiology, 45(1), 145-151. Hodda, M., , L., & Traunspurger, W. (2009). Nematode diversity in terrestrial, freshwater aquatic and marine systems. Nematodes as Environmental Indicators. CAB International, 45-93. Hogan, G., Walker, S., Turnbull, F., Curiao, T., Morrison, A. A., Flores, Y., Andrews, L., Claesson, M. J., Tangney, M., & Bartley, D. J. (2019). Microbiome analysis as a platform R&D tool for parasitic nematode disease management. The ISME Journal, 13(11), 2664-2680. Holden-Dye, L., & Walker, R. (2014). Anthelmintic drugs and nematocides: studies in Caenorhabditis elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_anthelminticdrugs.2/anthelminticdrug s.2.html. Hollenhorst, M. A., Clardy, J., & Walsh, C. T. (2009). The ATP-dependent amide ligases DdaG and DdaF assemble the fumaramoyl-dipeptide scaffold of the dapdiamide antibiotics. Biochemistry, 48(43), 10467-10472. Holmes, A. J., Tujula, N. A., Holley, M., Contos, A., James, J. M., Rogers, P., & Gillings, M. R. (2001). Phylogenetic structure of unusual aquatic microbial formations in Nullarbor caves, Australia. Environmental Microbiology, 3(4), 256-264. Holmström, C., James, S., Egan, S., & Kjelleberg, S. (1996). Inhibition of common fouling organisms by marine bacterial isolates ith special reference to the role of pigmented bacteria. Biofouling, 10(1-3), 251-259. Holmstrom, C., James, S., Neilan, B. A., White, D. C., & Kjelleberg, S. (1998). Pseudoalteromonas tunicata sp. nov., a bacterium that produces antifouling agents. International Journal of Systematic Bacteriology, 48(4), 1205-1212.

45

Holmström, C., Rittschof, D., & Kjelleberg, S. (1992). Inhibition of settlement by larvae of Balanus amphitrite and Ciona intestinalis by a surface-colonizing marine bacterium. Applied and Environmental Microbiology, 58(7), 2111. Hu, P. J. (2018). Dauer. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_dauer/dauer.html. Hutton, G., & Bartram, J. (2008). Global costs of attaining the Millennium Development Goal for water supply and sanitation. Bulletin of the World Health Organization, 86, 13-19. Irazoqui, J. E., Urbach, J. M., & Ausubel, F. M. (2010). Evolution of host innate defence: insights from Caenorhabditis elegans and primitive invertebrates. Nature Reviews Immunology, 10(1), 47. James, S. G., Holmström, C., & Kjelleberg, S. (1996). Purification and characterization of a novel antibacterial protein from the marine bacterium D2. Applied and Environmental Microbiology, 62(8), 2783-2788. Jia, K., Thomas, C., Akbar, M., Sun, Q., Adams-Huet, B., Gilpin, C., & Levine, B. (2009). Autophagy genes protect against Salmonella typhimurium infection and mediate insulin signaling-regulated pathogen resistance. Proceedings of the National Academy of Sciences of the USA, 106(34), 14564-14569. Johnson, C. D., & Stretton, A. O. (1985). Localization of choline acetyltransferase within identified motoneurons of the nematode Ascaris. The Journal of Neuroscience, 5(8), 1984-1992. Johnson, C. D., & Stretton, A. O. (1987). GABA-immunoreactivity in inhibitory motor neurons of the nematode Ascaris. The Journal of Neuroscience, 7(1), 223. Jones, J. T., Haegeman, A., Danchin, E. G., Gaur, H. S., Helder, J., Jones, M. G., Kikuchi, T., Manzanilla‐López, R., Palomares‐Rius, J. E., & Wesemael, W. M. (2013). Top 10 plant‐parasitic nematodes in molecular plant pathology. Molecular Plant Pathology, 14(9), 946-961. Jourdan, P. M., Lamberton, P. H. L., Fenwick, A., & Addiss, D. G. (2018). Soil- transmitted helminth infections. The Lancet, 391(10117), 252-265. Kaminsky, R., Ducray, P., Jung, M., Clover, R., Rufener, L., Bouvier, J., Weber, S. S., Wenger, A., Wieland-Berghausen, S., & Goebel, T. (2008). A new class of anthelmintics effective against drug-resistant nematodes. Nature, 452(7184), 176. Kanase, H., & Singh, K. N. (2018). Marine pharmacology: Potential, challenges, and future in India. Journal of Medical Sciences, 38(2), 49-53. Kancharla, P., Kelly, J. X., & Reynolds, K. A. (2015). Synthesis and structure–activity relationships of tambjamines and B-ring functionalized prodiginines as potent antimalarials. Journal of Medicinal Chemistry, 58(18), 7286-7309. Kaplan, J. M., & Horvitz, H. R. (1993). A dual mechanosensory and chemosensory neuron in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the USA, 90(6), 2227-2231. Kaplan, R. M., & Vidyashankar, A. N. (2012). An inconvenient truth: Global worming and anthelmintic resistance. Veterinary Parasitology, 186(1), 70-78. Karp, X. (2018). Working with dauer larvae. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at https://www.ncbi.nlm.nih.gov/books/NBK535516/. Kassie, Y. G., Yesuf, Ebrahim, A. S. & Mohamed, M. Y. (2020). Interaction effect between Meloidogyne incognita and Fusarium oxysporum f. sp. lycopersici on selected tomato (Solanum lycopersicum L.) genotypes. African Journal of Agricultural Research, 15(3), pp. 330-342. 46

Katharios-Lanwermeyer, S., Xi, C., Jakubovics, N., & Rickard, A. (2014). Mini-review: microbial coaggregation: ubiquity and implications for biofilm development. Biofouling, 30(10), 1235-1251. Kergunteuil, A., Campos-Herrera, R., Sánchez-Moreno, S., Vittoz, P., & Rasmann, S. (2016). The abundance, diversity, and metabolic footprint of soil nematodes is highest in high elevation alpine grasslands. Frontiers in Ecology and Evolution, 4, 84. Kiani, H., Haghighi, A., Rostami, A., Azargashb, E., TABAEI, S. J. S., Solgi, A., & Zebardast, N. (2016). Prevalence, risk factors and symptoms associated to intestinal parasite infections among patients with gastrointestinal disorders in Nahavand, Western Iran. Revista do Instituto de Medicina Tropical de São Paulo, 58:42. Kissoyan, K. A. B., Drechsler, M., Stange, E.-L., Zimmermann, J., Kaleta, C., Bode, H. B., & Dierking, K. (2019). Natural C. elegans microbiota protects against infection via production of a cyclic lipopeptide of the viscosin group. Current Biology, 29(6), 1030-1037.e1035. Klapper, M., Braga, D., Lackner, G., Herbst, R., & Stallforth, P. (2018). Bacterial alkaloid biosynthesis: structural diversity via a minimalistic nonribosomal peptide synthetase. Cell Chemical Biology, 25(6), 659-665. Kobayashi, M., Aoki, S., Gato, K., Matsunami, K., Kurosu, M., & Kitigawa, I. (1994). Marine natural products. XXXIV. Trisindoline, a new antibiotic indole trimer, produced by a bacterium of Vibrio sp. separated from the marine sponge Hyrtios altum. Chemical and Pharmaceutical Bulletin, 42(12), 2449-2451. Kong, C., Eng, S.-A., Lim, M.-P., & Nathan, S. (2016). Beyond traditional antimicrobials: A Caenorhabditis elegans model for discovery of novel anti-infectives. Frontiers in Microbiology, 7, 1956. Kong, D.-X., Jiang, Y.-Y., & Zhang, H.-Y. (2010). Marine natural products as sources of novel scaffolds: Achievement and concern. Drug Discovery Today, 15(21- 22):884-6. König, G. M., Kehraus, S., Seibert, S. F., Abdel‐Lateff, A., & Müller, D. (2006). Natural products from marine organisms and their associated microbes. ChemBioChem, 7(2), 229-238. Kothari, V., Sharma, S., & Padia, D. (2017). Recent research advances on Chromobacterium violaceum. Asian Pacific Journal of Tropical Medicine, 10(8), 744-752. Kumar, V., Rao, D., Thomas, T., Kjelleberg, S., & Egan, S. (2011). Antidiatom and antibacterial activity of epiphytic bacteria isolated from Ulva lactuca in tropical waters. World Journal of Microbiology and Biotechnology, 27(7), 1543-1549. Kumar, V. N. (2008). Inhibition of primary colonizers by marine surface-associated bacteria. Master dissertation, University of the South Pacific, Fiji. Kumarasingha, R., Young, N. D., Yeo, T.-C., Lim, D. S., Tu, C.-L., Palombo, E. A., Shaw, J. M., Gasser, R. B., & Boag, P. R. (2019). Transcriptional alterations in Caenorhabditis elegans following exposure to an anthelmintic fraction of the plant Picria fel-terrae Lour. Parasites & Vectors, 12(1), 181. Laing, R., Kikuchi, T., Martinelli, A., Tsai, I. J., Beech, R. N., Redman, E., Holroyd, N., Bartley, D. J., Beasley, H., & Britton, C. (2013). The genome and transcriptome of Haemonchus contortus, a key model parasite for drug and vaccine discovery. Genome Biology, 14(8), 1-16.

47

Lakowski, B., & Hekimi, S. (1998). The genetics of caloric restriction in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the USA, 95(22), 13091-13096. Lamb, J., Elliott, T., Chambers, M., & Chick, B. (2017). Broad spectrum anthelmintic resistance of Haemonchus contortus in Northern NSW of Australia. Veterinary Parasitology, 241, 48-51. Lambshead, P. J. D. (1993). Recent developments in marine benthic biodiversity research. Oceanis, 19:5–24 Leung, M. C. K., Williams, P. L., Benedetto, A., Au, C., Helmcke, K. J., Aschner, M., & Meyer, J. N. (2008). Caenorhabditis elegans: An emerging model in biomedical and environmental toxicology. Toxicological Sciences, 106(1), 5-28. Levsen, A., Svanevik, C. S., Cipriani, P., Mattiucci, S., Gay, M., Hastie, L. C., Bušelić, I., Mladineo, I., Karl, H., Ostermeyer, U., Buchmann, K., Højgaard, D. P., González, Á. F., Pascual, S., & , G. J. (2018). A survey of zoonotic nematodes of commercial key fish species from major European fishing grounds—Introducing the FP7 PARASITE exposure assessment study. Fisheries Research, 202, 4-21. Lewis, J. A., Wu, C. H., Levine, J. H., & Berg, H. (1980). Levamisole-resitant mutants of the nematode Caenorhabditis elegans appear to lack pharmacological acetylcholine receptors. Neuroscience, 5(6), 967-989. Liberati, N. T., Fitzgerald, K. A., Kim, D. H., Feinbaum, R., Golenbock, D. T., & Ausubel, F. M. (2004). Requirement for a conserved Toll/interleukin-1 resistance domain protein in the Caenorhabditis elegans immune response. Proceedings of the National Academy of Sciences of the USA, 101(17), 6593-6598. Liu, W., & Park, S.-W. (2018). Underground mystery: Interactions between plant roots and parasitic nematodes. Current Plant Biology, 15, 25-29. Lok, J. B., & Unnasch, T. R. (2018). Transgenesis in animal parasitic nematodes: Strongyloides spp. and Brugia spp. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_transgenesis/transgenesis.html. Lopes, L., Nicolino, R., Capanema, R., Oliveira, C., Haddad, J., & Eckstein, G. (2015). Economic impacts of parasitic diseases in cattle. CAB Reviews, 10(051), 1-10. Love, S. (2007). Primefact 478, drench resistance and sheep worm control. NSW Department of Primary Industries: Sydney. Lozano, G. L., Guan, C., Cao, Y., Borlee, B. R., Broderick, N. A., Stabb, E. V., & Handelsman, J. (2020). A chemical counterpunch: Chromobacterium violaceum ATCC 31532 produces violacein in response to translation-inhibiting antibiotics. mBio, 11(3), e00948-00920. Lucas-Elío, P., Gómez, D., Solano, F., & Sanchez-Amat, A. (2006). The antimicrobial activity of marinocine, synthesized by Marinomonas mediterranea, is due to hydrogen peroxide generated by its lysine oxidase activity. Journal of Bacteriology, 188(7), 2493-2501. Lucas-Elio, P., Hernandez, P., Sanchez-Amat, A., & Solano, F. (2005). Purification and partial characterization of marinocine, a new broad-spectrum antibacterial protein produced by Marinomonas mediterranea. Biochim Biophys Acta, 1721(1-3), 193- 203. Magnabosco, C., Lin, L.-H., Dong, H., Bomberg, M., Ghiorse, W., Stan-Lotter, H., , K., Kieft, T., Van Heerden, E., & Onstott, T. C. (2018). The biomass and biodiversity of the continental subsurface. Nature Geoscience, 11(10), 707- 717. 48

Magnino, S., Colin, P., Dei-Cas, E., Madsen, M., McLauchlin, J., Nöckler, K., Prieto Maradona, M., Tsigarida, E., Vanopdenbosch, E., & Van Peteghem, C. (2009). Biological risks associated with consumption of reptile products. International Journal of Food Microbiology, 134(3), 163-175. Mai-Prochnow, A., Evans, F., Dalisay-Saludes, D., Stelzer, S., Egan, S., James, S., Webb, J. S., & Kjelleberg, S. (2004). Biofilm development and cell death in the marine bacterium Pseudoalteromonas tunicata. Applied and Environmental Microbiology, 70(6), 3232. Maqbool, I., Wani, Z. A., Shahardar, R. A., Allaie, I. M., & Shah, M. M. (2017). Integrated parasite management with special reference to gastro-intestinal nematodes. Journal of Parasitic Diseases, 41(1), 1-8. Marchetti, P. M., Kelly, V., Simpson, J. P., Ward, M., & Campopiano, D. J. (2018). The carbon chain-selective adenylation enzyme TamA: the missing link between fatty acid and pyrrole natural product biosynthesis. Organic & Biomolecular Chemistry, 16(15), 2735-2740. Marris, E. (2006). Drugs from the deep. Nature 443, 904–905. Marroquin, L. D., Elyassnia, D., Griffitts, J. S., Feitelson, J. S., & Aroian, R. V. (2000). Bacillus thuringiensis (Bt) toxin susceptibility and isolation of resistance mutants in the nematode Caenorhabditis elegans. Genetics, 155(4), 1693-1699. Mathew, M. D., Mathew, N. D., Miller, A., Simpson, M., Au, V., Garland, S., Gestin, M., Edgley, M. L., Flibotte, S., & Balgi, A. (2016). Using C. elegans forward and reverse genetics to identify new compounds with anthelmintic activity. PLoS Neglected Tropical Diseases, 10(10). Matz, C., Deines, P., Boenigk, J., Arndt, H., Eberl, L., Kjelleberg, S., & Jurgens, K. (2004). Impact of violacein-producing bacteria on survival and feeding of bacterivorous nanoflagellates. Applied and Environmental Microbiology, 70(3), 1593-1599. Matz, C., Webb, J. S., Schupp, P. J., Phang, S. Y., Penesyan, A., Egan, S., Steinberg, P., & Kjelleberg, S. (2008). Marine biofilm bacteria evade eukaryotic predation by targeted chemical defense. PloS One, 3(7), e2744. May, K., Brügemann, K., König, S., & Strube, C. (2018). The effect of patent Dictyocaulus viviparus (re)infections on individual milk yield and milk quality in pastured dairy cows and correlation with clinical signs. Parasites & Vectors, 11(1), 24. McElwee, J. J., Schuster, E., Blanc, E., Thomas, J. H., & Gems, D. (2004). Shared transcriptional signature in Caenorhabditis elegans Dauer larvae and long-lived daf-2 mutants implicates detoxification system in longevity assurance. Journal of Biological Chemistry, 279(43), 44533-44543. McLeod, R. (1995). Costs of major parasites to the Australian livestock industries. International Journal for Parasitology, 25(11), 1363-1367. McLeod, R. (2004). Worm control of small ruminants in tropical Asia. Australian Centre for International Agricultural Research 2004 at https://aciar.gov.au/publication/books-and-manuals/worm-control-small- ruminants-tropical-asia Mehrdana, F., & Buchmann, K. (2017). Excretory/secretory products of anisakid nematodes: biological and pathological roles. Acta Veterinaria Scandinavica, 59(1), 42. Meisel, J. D., & Kim, D. H. (2014). Behavioral avoidance of pathogenic bacteria by Caenorhabditis elegans. Trends in Immunology, 35(10), 465-470.

49

Melville, L. A., McBean, D., Fyfe, A., Campbell, S.-J., Palarea-Albaladejo, J., & Kenyon, F. (2016). Effect of anthelmintic treatment strategy on strongylid nematode species composition in grazing lambs in Scotland. Parasites & Vectors, 9, 199- 199. Mertenskötter, A., Keshet, A., Gerke, P., & Paul, R. J. (2013). The p38 MAPK PMK-1 shows heat-induced nuclear translocation, supports chaperone expression, and affects the heat tolerance of Caenorhabditis elegans. Cell Stress & Chaperones, 18(3), 293-306. Midkiff, D., & San-Miguel, A. (2019). Microfluidic technologies for high throughput screening through sorting and on-chip culture of C. elegans. Molecules, 24(23), 4292. Mir, D. A., & Krishnaswamy, B. (2019). Global proteomic response of Caenorhabditis elegans against PemKSa toxin. Frontiers in Cellular and Infection Microbiology, 9, 172. Mitreva, M., McCarter, J. P., Martin, J., Dante, M., Wylie, T., Chiapelli, B., Pape, D., Clifton, S. W., Nutman, T. B., & Waterston, R. H. (2004). Comparative genomics of gene expression in the parasitic and free-living nematodes Strongyloides stercoralis and Caenorhabditis elegans. Genome Research, 14(2), 209-220. Moens, T., Traunspurger, W., & Bergtold, M. (2006). Feeding ecology of free-living benthic nematodes. Freshwater Nematodes. Ecology and Taxonomy. CAB International Publishing, 105-131. Moestrup, Ø. (2000). The flagellates: unity, diversity and evolution. London, UK, Taylor &Francis. Monaghan, R. L., & Tkacz, J. S. (1990). Bioactive microbial products: focus upon mechanism of action. Annual Review of Microbiology, 44(1), 271-331. Moneo, I., Caballero, M. L., González-Muñoz, M., Rodríguez-Mahillo, A. I., Rodríguez- Perez, R., & Silva, A. (2005). Isolation of a heat-resistant allergen from the fish parasite Anisakis simplex. Parasitology Research, 96(5), 285-289. Montalvo-Katz, S., Huang, H., Appel, M. D., Berg, M., & Shapira, M. (2013). Association with soil bacteria enhances p38-dependent infection resistance in Caenorhabditis elegans. Infection and Immunity, 81(2), 514. Morgan, E. R., Aziz, N.-A. A., Blanchard, A., Charlier, J., Charvet, C., Claerebout, E., Geldhof, P., Greer, A. W., Hertzberg, H., & Hodgkinson, J. (2019). 100 questions in livestock helminthology research. Trends in Parasitology, 35(1), 52-71. Murphy, C. T., McCarroll, S. A., Bargmann, C. I., Fraser, A., Kamath, R. S., Ahringer, J., Li, H., & Kenyon, C. (2003). Genes that act downstream of DAF-16 to influence the lifespan of Caenorhabditis elegans. Nature, 424(6946), 277-283. Mushonga, B., Habumugisha, D., Kandiwa, E., Madzingira, O., Samkange, A., Segwagwe, B. E., & Jaja, I. F. (2018). Prevalence of Haemonchus contortus infections in sheep and goats in Nyagatare District, Rwanda. Journal of Veterinary Medicine, 2018. Nappi, J. (2019). Discovery of novel bioactive metabolites from marine epiphytic bacteria and assessment of their ecological role. Unpublished Doctoral dissertation, University of New South Wales, Australia. Nery, S. V., McCarthy, J. S., Traub, R., Andrews, R. M., Black, J., Gray, D., Weking, E., Atkinson, J.-A., Campbell, S., Francis, N., Vallely, A., Williams, G., & Clements, A. (2015). A cluster-randomised controlled trial integrating a community-based water, sanitation and hygiene programme, with mass distribution of albendazole to reduce intestinal parasites in Timor-Leste: the WASH for WORMS research protocol. BMJ Open, 5(12), e009293. 50

Neu, A. K., Månsson, M., Gram, L., & Prol-García, M. J. (2014). Toxicity of bioactive and probiotic marine bacteria and their secondary metabolites in Artemia sp. and Caenorhabditis elegans as eukaryotic model organisms. Applied and Environmental Microbiology, 80(1), 146-153. Ngcamphalala, P., Lamb, J., & Mukaratirwa, S. (2020). Molecular identification of hookworm isolates from stray dogs, humans and selected wildlife from South Africa. Journal of Helminthology, 94. Oclarit, J. M., Okada, H., Ohta, S., Kaminura, K., Yamaoka, Y., Iizuka, T., Miyashiro, S., & Ikegami, S. (1994). Anti-bacillus substance in the marine sponge, Hyatella species, produced by an associated Vibrio species bacterium. Microbios, 78(314), 7-16. Oh, S. W., Mukhopadhyay, A., Dixit, B. L., Raha, T., Green, M. R., & Tissenbaum, H. A. (2006). Identification of direct DAF-16 targets controlling longevity, metabolism and diapause by chromatin immunoprecipitation. Nature Genetics, 38(2), 251-257. Omura, S. (2008). Ivermectin: 25 years and still going strong. International Journal of Antimicrobial Agents, 31(2), 91-98. Orobets, V., Lisovets, E., Zabashta, S., & Ermakov, A. (2019). Control of fish parasites in aquaculture. IOP Conference Series: Earth and Environmental Science, 403, 012065. Ortega-Morales, B. O., Chan-Bacab, M. J., De la Rosa-García, S. d. C., & Camacho- Chab, J. C. (2010). Valuable processes and products from marine intertidal microbial communities. Current Opinion in Biotechnology, 21(3), 346-352. Page, A. P., & Johnstone, I. (2007). The cuticle. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_cuticle/cuticle.html. Page, A. P., Roberts, M., Félix, M.-A., Pickard, D., Page, A., & Weir, W. (2019). The golden death bacillus Chryseobacterium nematophagum is a novel matrix digesting pathogen of nematodes. BMC Biology, 17(1), 10. Palma Esposito, F., Ingham, C. J., Hurtado-Ortiz, R., Bizet, C., Tasdemir, D., & de Pascale, D. (2018). Isolation by Miniaturized Culture Chip of an Antarctic bacterium Aequorivita sp. with antimicrobial and anthelmintic activity. Biotechnology Reports, 20, e00281. Papadi, B., Boudreaux, C., Tucker, J. A., Mathison, B., Bishop, H., & Eberhard, M. E. (2013). Halicephalobus gingivalis: a rare cause of fatal meningoencephalomyelitis in humans. The American Journal of Tropical Medicine and Hygiene, 88(6), 1062-1064. Pascoal, F., Magalhães, C., & Costa, R. (2020). The link between the ecology of the prokaryotic rare biosphere and its biotechnological potential. Frontiers in Microbiology, 11 (231), 1-8. Penesyan, A., Ballestriero, F., Daim, M., Kjelleberg, S., Thomas, T., & Egan, S. (2013). Assessing the effectiveness of functional genetic screens for the identification of bioactive metabolites. Marine Drugs, 11(1), 40-49. Penesyan, A., Kjelleberg, S., & Egan, S. (2010). Development of novel drugs from marine surface associated microorganisms. Marine Drugs, 8(3), 438-459. Peng, D., Wan, D., Cheng, C., Ye, X., & Sun, M. (2018). Nematode-specific cadherin CDH-8 acts as a receptor for Cry5B toxin in Caenorhabditis elegans. Applied Microbiology and Biotechnology, 102(8), 3663-3673.

51

Persson, G., Ekmann, J. R., & Hviid, T. V. F. (2019). Reflections upon immunological mechanisms involved in fertility, pregnancy and parasite infections. Journal of Reproductive Immunology, 136, 102610. Pinkerton, D. M., Banwell, M. G., Garson, M. J., Kumar, N., de Moraes, M. O., Cavalcanti, B. C., Barros, F. W., & Pessoa, C. (2010). Antimicrobial and cytotoxic activities of synthetically derived tambjamines C and E - J, BE-18591, and a related alkaloid from the marine bacterium Pseudoalteromonas tunicata. Chemistry & Biodiversity, 7(5), 1311-1324. Pradel, E., Zhang, Y., Pujol, N., Matsuyama, T., Bargmann, C. I., & Ewbank, J. J. (2007). Detection and avoidance of a natural product from the pathogenic bacterium Serratia marcescens by Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the USA 104(7), 2295-2300. Proksch, P., Edrada, R. A., & Ebel, R. (2002). Drugs from the seas - current status and microbiological implications. Applied Microbiology and Biotechnology, 59(2-3), 125-134. Pujol, N., Cypowyj, S., Ziegler, K., Millet, A., Astrain, A., Goncharov, A., Jin, Y., Chisholm, A. D., & Ewbank, J. J. (2008). Distinct innate immune responses to infection and wounding in the C. elegans epidermis. Current Biology, 18(7), 481- 489. Queirós, L., Pereira, J., Gonçalves, F., Pacheco, M., Aschner, M., & Pereira, P. (2019). Caenorhabditis elegans as a tool for environmental risk assessment: emerging and promising applications for a “nobelized worm”. Critical Reviews in Toxicology, 49(5), 411-429. Queiroz, K. C. S., Milani, R., Ruela-de-Sousa, R. R., Fuhler, G. M., Justo, G. Z., Zambuzzi, W. F., Duran, N., Diks, S. H., Spek, C. A., Ferreira, C. V., & Peppelenbosch, M. P. (2012). Violacein induces death of resistant leukaemia cells via kinome reprogramming, endoplasmic reticulum stress and golgi apparatus collapse. PloS one, 7(10), e45362. Rahul, S., Chandrashekhar, P., Hemant, B., Chandrakant, N., Laxmikant, S., & Satish, P. (2014). Nematicidal activity of microbial pigment from Serratia marcescens. Natural Product Research, 28(17), 1399-1404. Rai, P., Bharati, M., Subba, A., & Saha, D. (2019). Insecticide resistance mapping in the vector of lymphatic filariasis, Culex quinquefasciatus Say from northern region of West Bengal, India. PloS One, 14(5), e0217706. Ramos, P. (2020). Parasites in fishery products - Laboratorial and educational strategies to control. Experimental Parasitology, 211, 107865. Rao, D., Webb, J. S., & Kjelleberg, S. (2005). Competitive interactions in mixed-species biofilms containing the marine bacterium Pseudoalteromonas tunicata. Applied and Environmental Microbiology, 71(4), 1729-1736. Rashid, M., Rashid, M. I., Akbar, H., Ahmad, L., Hassan, M. A., Ashraf, K., Saeed, K., & Gharbi, M. (2019). A systematic review on modelling approaches for economic losses studies caused by parasites and their associated diseases in cattle. Parasitology, 146(2), 129-141. Ray, A., Rentas, C., Caldwell, G. A., & Caldwell, K. A. (2015). Phenazine derivatives cause proteotoxicity and stress in C. elegans. Neuroscience letters, 584, 23-27. Rengarajan, S., & Hallem, E. A. (2016). Olfactory circuits and behaviors of nematodes. Current Opinion in Neurobiology, 41, 136-148. doi:10.1016/j.conb.2016.09.002 Riddle, D., Blumenthal, T., Meyer, B., & Priess, J. (1997). The biological model. C elegans. C. elegans II. 2nd edition. Cold Spring Harbor, New York. Available at https://www.ncbi.nlm.nih.gov/books/NBK20086/ 52

Rinehart, K. L., Gloer, J. B., Hughes, R. G., Renis, H. E., McGovren, J. P., Swynenberg, E. B., Stringfellow, D. A., Kuentzel, S. L., & Li, L. H. (1981). Didemnins: antiviral and antitumor depsipeptides from a Caribbean tunicate. Science, 212(4497), 933-935. Rischer, M., Raguž, L., Guo, H., Keiff, F., Diekert, G., Goris, T., & Beemelmanns, C. (2018). Biosynthesis, synthesis, and activities of Barnesin A, a NRPS-PKS hybrid produced by an anaerobic Epsilonproteobacterium. ACS Chemical Biology, 13(8), 1990-1995. Rocha-Martin, J., Harrington, C., Dobson, A. D., & O'Gara, F. (2014). Emerging strategies and integrated systems microbiology technologies for biodiscovery of marine bioactive compounds. Marine Drugs, 12(6), 3516-3559. Roeber, F., Jex, A. R., & Gasser, R. B. (2013). Impact of gastrointestinal parasitic nematodes of sheep, and the role of advanced molecular tools for exploring epidemiology and drug resistance - an Australian perspective. Parasites & Vectors, 6(1), 153. Romano, G., Costantini, M., Sansone, C., Lauritano, C., Ruocco, N., & Ianora, A. (2017). Marine microorganisms as a promising and sustainable source of bioactive molecules. Marine Environmental Research, 128, 58-69. Ruiz-Lancheros, E., Viau, C., Walter, T. N., Francis, A., & Geary, T. G. (2011). Activity of novel nicotinic anthelmintics in cut preparations of Caenorhabditis elegans. International Journal for Parasitology, 41(3), 455-461. Rusch, D. B., Halpern, A. L., Sutton, G., Heidelberg, K. B., Williamson, S., Yooseph, S., Wu, D., Eisen, J. A., Hoffman, J. M., Remington, K., Beeson, K., Tran, B., Smith, H., Baden-Tillson, H., Stewart, C., Thorpe, J., Freeman, J., Andrews-Pfannkoch, C., Venter, J. E., Li, K., Kravitz, S., Heidelberg, J. F., Utterback, T., Rogers, Y.- H., Falcón, L. I., Souza, V., Bonilla-Rosso, G., Eguiarte, L. E., Karl, D. M., Sathyendranath, S., Platt, T., Bermingham, E., Gallardo, V., Tamayo-Castillo, G., Ferrari, M. R., Strausberg, R. L., Nealson, K., Friedman, R., Frazier, M., & Venter, J. C. (2007). The Sorcerer II Global Ocean Sampling Expedition: Northwest Atlantic through Eastern Tropical Pacific. PLOS Biology, 5(3), e77. Sackett, D., Holmes, P., Abbott, K., Jephcott, S., & Barber, M. (2006). Assessing the economic cost of endemic disease on the profitability of Australian beef cattle and sheep producers. MLA (Meat and Livestock) Report AHW, 87. Samuel, B. S., Rowedder, H., Braendle, C., Félix, M.-A., & Ruvkun, G. (2016). Caenorhabditis elegans responses to bacteria from its natural habitats. Proceedings of the National Academy of Sciences of the USA 113(27), E3941. Sasidharan, A., Sasidharan, N. K., Amma, D. B. N. S., Vasu, R. K., Nataraja, A. V., & Bhaskaran, K. (2015). Antifungal activity of violacein purified from a novel strain of Chromobacterium sp. NIIST (MTCC 5522). Journal of Microbiology, 53(10), 694-701. Schaeffer, J. M., Blizzard, T. A., Ondeyka, J., Goegelman, R., Sinclair, P. J., & Mrozik, H. (1992). [3H] Paraherquamide binding to Caenorhabditis elegans: Studies on a potent new anthelmintic agent. Biochemical Pharmacology, 43(4), 679-684. Schulenburg, H., & Félix, M.-A. (2017). The natural biotic environment of Caenorhabditis elegans. Genetics, 206(1), 55-86. Scott, I., Pomroy, W., Kenyon, P., Smith, G., Adlington, B., & Moss, A. (2013). Lack of efficacy of monepantel against Teladorsagia circumcincta and Trichostrongylus colubriformis. Veterinary Parasitology, 198(1-2), 166-171.

53

Sczesny-Moraes, E. A., Bianchin, I., da Silva, K. F., Catto, J. B., Honer, M. R., & Paiva, F. (2010). Anthelmintic resistance of gastrointestinal nematodes in sheep, Mato Grosso do Sul, Brazil. Pesquisa Veterinária Brasileira, 30(3), 229-236. Sekurova, O. N., Schneider, O., & Zotchev, S. B. (2019). Novel bioactive natural products from bacteria via bioprospecting, genome mining and metabolic engineering. Microbial Biotechnology, 12(5), 828-844. Shaheerah Alsounosi Alshareef, A. M. A. R. A. A. E. S. S., & Mohamed, C. (2019). Parasitic Contamination of raw vegetables sampled from different farm locations in Brack Al-Shati, Libya. Journal of Pure and Applied Sciences, 18(4), 210-214. Shalaby, H., Ashry, H., Saad, M., & Farag, T. (2020). In vitro effects of Streptomyces tyrosinase on the egg and adult worm of Toxocara vitulorum. Iranian Journal of Parasitology, 15(1), 67-75. Shamsi, S. (2019). Seafood-borne parasitic diseases: a “one-health” approach is needed. Fishes, 4(1), 9. Shinn, A. P., Pratoomyot, J., Bron, J. E., Paladini, G., Brooker, E. E., & Brooker, A. J. (2015). Economic costs of protistan and metazoan parasites to global mariculture. Parasitology, 142(1), 196-270. Sifri, C. D., Begun, J., Ausubel, F. M., & Calderwood, S. B. (2003). Caenorhabditis elegans as a model host for Staphylococcus aureus pathogenesis. Infection and Immunity, 71(4), 2208-2217. Singh, V., & Aballay, A. (2006). Heat-shock transcription factor (HSF)-1 pathway required for Caenorhabditis elegans immunity. Proceedings of the National Academy of Sciences of the USA, 103(35), 13092. Skerratt, J., Bowman, J., Hallegraeff, G., James, S., & Nichols, P. (2002). Algicidal bacteria associated with blooms of a toxic dinoflagellate in a temperate Australian estuary. Marine Ecology Progress Series, 244, 1-15. Skovhus, T. L., Holmström, C., Kjelleberg, S., & Dahllöf, I. (2007). Molecular investigation of the distribution, abundance and diversity of the genus Pseudoalteromonas in marine samples. FEMS Microbiology Ecology, 61(2), 348- 361. Skovhus, T. L., Ramsing, N. B., Holmström, C., Kjelleberg, S., & Dahllöf, I. (2004). Real-time quantitative PCR for assessment of abundance of Pseudoalteromonas species in marine samples. Applied and Environmental Microbiology, 70(4), 2373-2382. Smiley, R. W. (2010). Root-lesion nematodes: Biology and management in Pacific Northwest wheat cropping systems. Available from https://catalog.extension.oregonstate.edu/pnw617. Soliman, G. M., Ameen, H. H., Abdel-Aziz, S. M., & El-Sayed, G. M. (2019). In vitro evaluation of some isolated bacteria against the plant parasite nematode Meloidogyne incognita. Bulletin of the National Research Centre, 43(1), 171. Spiegler, V., Liebau, E., & Hensel, A. (2017). Medicinal plant extracts and plant-derived polyphenols with anthelmintic activity against intestinal nematodes. Natural product reports, 34(6), 627-643. Steinberg, P. D., & De Nys, R. (2002). Chemical mediation of colonisation of seaweed surfaces. Journal of Phycology, 38(4), 621-629. Stelzer, S., Egan, S., Larsen, M. R., Bartlett, D. H., & Kjelleberg, S. (2006). Unravelling the role of the ToxR-like transcriptional regulator WmpR in the marine antifouling bacterium Pseudoalteromonas tunicata. Microbiology, 152(5), 1385- 1394.

54

Stiernagle, T. (2006). Maintenance of C. elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_strainmaintain/strainmaintain.html. Sunagawa, S., Coelho, L. P., Chaffron, S., Kultima, J. R., Labadie, K., Salazar, G., Djahanschiri, B., Zeller, G., Mende, D. R., Alberti, A., Cornejo-Castillo, F. M., Costea, P. I., Cruaud, C., Ovidio, F., Engelen, S., Ferrera, I., Gasol, J. M., Guidi, L., Hildebrand, F., Kokoszka, F., Lepoivre, C., Lima-Mendez, G., Poulain, J., Poulos, B. T., Royo-Llonch, M., Sarmento, H., Vieira-Silva, S., Dimier, C., Picheral, M., Searson, S., Kandels-Lewis, S., Bowler, C., de Vargas, C., Gorsky, G., Grimsley, N., Hingamp, P., Iudicone, D., Jaillon, O., Not, F., Ogata, H., Pesant, S., Speich, S., Stemmann, L., Sullivan, M. B., Weissenbach, J., Wincker, P., Karsenti, E., Raes, J., Acinas, S. G., & Bork, P. (2015). Structure and function of the global ocean microbiome. Science, 348(6237), 1261359. Tan, M.-W., Mahajan-Miklos, S., & Ausubel, F. M. (1999). Killing of Caenorhabditis elegans by Pseudomonas aeruginosa used to model mammalian bacterial pathogenesis. Proceedings of the National Academy of Sciences of the USA, 96(2), 715. Tang, Y. T., Gao, X., Rosa, B. A., Abubucker, S., Hallsworth-Pepin, K., Martin, J., Tyagi, R., Heizer, E., Zhang, X., Bhonagiri-Palsikar, V., Minx, P., Warren, W. C., Wang, Q., Zhan, B., Hotez, P. J., Sternberg, P. W., Dougall, A., Gaze, S. T., Mulvenna, J., Sotillo, J., Ranganathan, S., Rabelo, E. M., Wilson, R. K., Felgner, P. L., Bethony, J., Hawdon, J. M., Gasser, R. B., Loukas, A., & Mitreva, M. (2014). Genome of the human hookworm Necator americanus. Nature Genetics, 46(3), 261-269. Thomas, T., Evans, F. F., Schleheck, D., Mai-Prochnow, A., Burke, C., Penesyan, A., Dalisay, D. S., Stelzer-Braid, S., Saunders, N., Johnson, J., Ferriera, S., Kjelleberg, S., & Egan, S. (2008). Analysis of the Pseudoalteromonas tunicata genome reveals properties of a surface-associated life style in the marine environment. PloS One, 3(9), e3252. Thomas, T. R. A., Kavlekar, D. P., & LokaBharathi, P. A. (2010). Marine drugs from sponge-microbe association--a review. Marine Drugs, 8(4), 1417-1468. Thompson, J., Greco, N., Eastwood, R., Sharma, S., & Scurrah, M. (2000). Integrated control of nematodes of cool season food legumes. In Linking Research and Marketing Opportunities for Pulses in the 21st Century (pp. 491-506): Springer. Troemel, E. R., Chu, S. W., Reinke, V., Lee, S. S., Ausubel, F. M., & Kim, D. H. (2006). p38 MAPK regulates expression of immune response genes and contributes to longevity in C. elegans. PLoS Genetics, 2(11), e183. Tsukimoto, M., Nagaoka, M., Shishido, Y., Fujimoto, J., Nishisaka, F., Matsumoto, S., Harunari, E., Imada, C., & Matsuzaki, T. (2011). Bacterial production of the tunicate-derived antitumor cyclic depsipeptide didemnin B. Journal of Natural Products, 74(11), 2329-2331. Urban Jr, J. F., Hu, Y., Miller, M. M., Scheib, U., Yiu, Y. Y., & Aroian, R. V. (2013). Bacillus thuringiensis-derived Cry5B has potent anthelmintic activity against Ascaris suum. PLoS Neglected Tropical Diseases, 7(6), e2263. Vadlamudi, R. K., Weber, E., Ji, I., Ji, T. H., & Bulla, L. A. (1995). Cloning and expression of a receptor for an insecticidal toxin of Bacillus thuringiensis. Journal of Biological Chemistry, 270(10), 5490-5494. Van den Brom, R., Moll, L., Kappert, C., & Vellema, P. (2015). Haemonchus contortus resistance to monepantel in sheep. Veterinary Parasitology, 209(3-4), 278-280. 55

Vellai, T., McCulloch, D., Gems, D., & Kovács, A. L. (2006). Effects of sex and insulin/insulin-like Growth Factor-1 signaling on performance in an associative learning paradigm in Caenorhabditis elegans. Genetics, 174(1), 309. Venter, J. C., Remington, K., Heidelberg, J. F., Halpern, A. L., Rusch, D., Eisen, J. A., Wu, D., Paulsen, I., Nelson, K. E., Nelson, W., Fouts, D. E., Levy, S., Knap, A. H., Lomas, M. W., Nealson, K., White, O., , J., Hoffman, J., Parsons, R., Baden-Tillson, H., Pfannkoch, C., Rogers, Y.-H., & Smith, H. O. (2004). Environmental genome shotgun sequencing of the Sargasso Sea. Science, 304(5667), 66. Verschave, S. H., Charlier, J., Rose, H., Claerebout, E., & Morgan, E. R. (2016). Cattle and nematodes under global change: transmission models as an ally. Trends in Parasitology, 32(9), 724-738. Villate, L., Morin, E., Demangeat, G., Van Helden, M., & Esmenjaud, D. (2012). Control of Xiphinema index populations by fallow plants under greenhouse and field conditions. Phytopathology, 102(6), 627-634. Viney, M. (2017). How can we understand the genomic basis of nematode parasitism? Trends in Parasitology, 33(6), 444-452. Vlamakis, H., Chai, Y., Beauregard, P., Losick, R., & Kolter, R. (2013). Sticking together: building a biofilm the Bacillus subtilis way. Nature Reviews Microbiology, 11(3), 157-168. Vuong, D., Capon, R. J., Lacey, E., Gill, J. H., Heiland, K., & Friedel, T. (2001). Onnamide F: A new nematocide from a southern Australian marine sponge, Trachycladus laevispirulifer. Journal of Natural Products, 64(5), 640-642. Wang, H., Park, H., Liu, J., & Sternberg, P. W. (2018). An efficient genome editing strategy to generate putative null mutants in Caenorhabditis elegans using CRISPR/Cas9. G3: Genes|Genomes|Genetics, 8(11), 3607. Wang, J., Nakad, R., & Schulenburg, H. (2012). Activation of the Caenorhabditis elegans FOXO family transcription factor DAF-16 by pathogenic Bacillus thuringiensis. Developmental & Comparative Immunology, 37(1), 193-201. Weeks, J. C., Robinson, K. J., Lockery, S. R., & Roberts, W. M. (2018). Anthelmintic drug actions in resistant and susceptible C. elegans revealed by electrophysiological recordings in a multichannel microfluidic device. International Journal for Parasitology: Drugs and Drug Resistance, 8(3), 607- 628. Wei, J.-Z., Hale, K., Carta, L., Platzer, E., Wong, C., Fang, S.-C., & Aroian, R. V. (2003). Bacillus thuringiensis crystal proteins that target nematodes. Proceedings of the National Academy of Sciencesof the USA, 100(5), 2760. Weitere, M., Erken, M., Majdi, N., Arndt, H., Norf, H., Reinshagen, M., Traunspurger, W., Walterscheid, A., & Wey, J. K. (2018). The food web perspective on aquatic biofilms. Ecological Monographs, 88(4), 543-559. Welsh, R. M., Zaneveld, J. R., Rosales, S. M., Payet, J. P., Burkepile, D. E., & Thurber, R. V. (2016). Bacterial predation in a marine host-associated microbiome. The ISME Journal, 10(6), 1540-1544. White, E. C., Houlden, A., Bancroft, A. J., Hayes, K. S., Goldrick, M., Grencis, R. K., & Roberts, I. S. (2018). Manipulation of host and parasite microbiotas: Survival strategies during chronic nematode infection. Science Advances, 4(3), eaap7399. Whittaker, J. H., Robertson, A. P., Kimber, M. J., Day, T. A., & Carlson, S. A. (2016). Intestinal Enterobacteriaceae that protect nematodes from the effects of benzimidazoles. Journal of Bacteriology & Parasitology, 7(5).

56

Wietz, M., Månsson, M., Bowman, J. S., Blom, N., Ng, Y., & Gram, L. (2012). Wide distribution of closely related, antibiotic-producing Arthrobacter strains throughout the Arctic Ocean. Applied and Environmental Microbiology, 78(6), 2039. Won, N.-I., Lee, G.-E., Ko, K.-B., Oh, D.-C., Na, Y.-H., & Park, J.-S. (2017). Identification of a bioactive compound, violacein, from Microbulbifer sp. isolated from a marine sponge Hymeniacidon sinapium on the west coast of Korea. Microbiology and Biotechnology Letters, 45(2), 124-132. Wu, Y.-H., Cheng, H., Xu, L., Jin, X.-B., Wang, C.-S., & Xu, X.-W. (2017). Physiological and genomic features of a novel violacein-producing bacterium isolated from surface seawater. PloS one, 12(6), e0179997. Xiong, Z.-Q., Wang, J.-F., Hao, Y.-Y., & Wang, Y. (2013). Recent advances in the discovery and development of marine microbial natural products. Marine Drugs, 11(3), 700-717. Xu, Y., Kersten, R. D., Nam, S.-J., Lu, L., Al-Suwailem, A. M., Zheng, H., Fenical, W., Dorrestein, P. C., Moore, B. S., & Qian, P.-Y. (2012). Bacterial biosynthesis and maturation of the didemnin anti-cancer agents. Journal of the American Chemical Society, 134(20), 8625-8632. Yamazaki, M., Okuyama, E., Kobayashi, M., & Inoue, H. (1981). The structure of paraherquamide, a toxic metabolite from Penicillium paraherquei. Tetrahedron Letters, 22(2), 135-136. Yu, J., Du, G., Li, R., Li, L., Li, Z., Zhou, C., Chen, C., & Guo, D. (2015). Nematicidal activities of bacterial volatiles and components from two marine bacteria, Pseudoalteromonas marina strain H-42 and Vibrio atlanticus strain S-16, against the pine wood nematode, Bursaphelenchus xylophilus. Nematology, 17(9), 1011- 1025. Zajíčková, M., Nguyen, L. T., Skálová, L., Raisová Stuchlíková, L., & Matoušková, P. (2020). Anthelmintics in the future: current trends in the discovery and development of new drugs against gastrointestinal nematodes. Drug Discovery Today, 25(2), 430-437. Zhang, F., Berg, M., Dierking, K., Félix, M.-A., Shapira, M., Samuel, B. S., & Schulenburg, H. (2017). Caenorhabditis elegans as a model for microbiome research. Frontiers in Microbiology, 8, 485. Zhang, W., Ding, W., Li, Y.-X., Tam, C., Bougouffa, S., Wang, R., Pei, B., Chiang, H., Leung, P., & Lu, Y. (2019). Marine biofilms constitute a bank of hidden microbial diversity and functional potential. Nature Communications, 10(1), 1-10. Zhang, Y., Lu, H., & Bargmann, C. I. (2005). Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis elegans. Nature, 438(7065), 179-184. Zhao, X.-D., Zhang, B.-W., Fu, L.-J., Li, Q.-L., Lin, Y., & Yu, X.-Q. (2020). Possible insecticidal mechanism of Cry41-related toxin against Myzus persicae by enhancing Cathepsin B activity. Journal of Agricultural and Food Chemistry, 68(16), 4607-4615. Zimmermann, J., Obeng, N., Yang, W., Pees, B., Petersen, C., Waschina, S., Kissoyan, K. A., Aidley, J., Hoeppner, M. P., & Bunk, B. (2020). The functional repertoire contained within the native microbiota of the model nematode Caenorhabditis elegans. The ISME Journal, 14(1), 26-38. Zou, C. G., Tu, Q., Niu, J., Ji, X. L., & Zhang, K. Q. (2013). The DAF-16/FOXO transcription factor functions as a regulator of epidermal innate immunity. PLoS Pathogens, 9(10), e1003660.

57

Zwart, R. S., Thudi, M., Channale, S., Manchikatla, P. K., Varshney, R. K., & Thompson, J. P. (2019). Resistance to plant-parasitic nematodes in chickpea: Current status and future perspectives. Frontiers in plant science, 10, 966. WHO (World Health Organization) (2012). Soil-transmitted helminth infections. Available at https://www.who.int/news-room/fact-sheets/detail/soil-transmitted-helminth infections. Accessed on 2 July 2020 WHO (World Health Organization) (2017). Guideline: preventive chemotherapy to control soil-transmitted helminth infections in at-risk population groups. Geneva, World Health Organization. Available at https://www.who.int/nutrition/publications/guidelines/deworming/en/. Accessed on 2 June 2020. WHO (World Health Organization) (2020) Soil-transmitted helminth infections. Available at https://www.who.int/news-room/fact-sheets/detail/soil-transmitted-helminth infections. Accessed on 30 June 2020. World Bank Report (2008) World Development Report 2008: Agriculture for Development. Available at https://documents.worldbank.org/en/publication/documents reports/documentdetail/587251468175472382/world-development-report 2008-agriculture-for-development. Accessed on 29 May 2020.

58

CHAPTER 2

Characterisation of the fast nematode-killing activity of the marine epiphytic bacterium Pseudoalteromonas tunicata D2

59

2.1 INTRODUCTION

Diseases resulting from parasitic helminth (i.e. nematode) infections are a global concern but especially for resource-limited countries with poor hygiene and inadequate clean water supply (Ziegelbauer et al., 2012; Mahmud et al., 2015). Globally approximately, 1.5 billion people suffer from parasitic helminth infections with the majority of cases occurring in East Asia, sub-Saharan Africa, America and China (WHO, 2016). In humans, parasitic nematode infections are associated with malnutrition, anaemia, growth retardation, diminishing fitness, reduced cognition and result in numerous fatalities per year (Stephenson et al., 2000; Mahmud et al., 2015; Plummer et al., 2016). In addition to human illness, parasitic nematodes can also infect agricultural crops and aquaculture products, hence reducing yield and threatening food security (Nicol et al., 2011; Coyne et al., 2018).

For the past 40 years, strategies to control parasitic nematodes have almost exclusively relied on intensive chemotherapy to relieve symptoms and diminish transmission. However, overuse and frequent parasite exposure to single broad-spectrum therapeutic drugs can increase the prevalence of anthelmintic resistance among the harmful parasites (Wolstenholme et al., 2004; Shalaby, 2013; Geurden et al., 2015; Ballesteros et al., 2018; Becker et al., 2018). There is now evidence of parasitic nematode resistance to all of the currently available nematicides including Piperazine, Benzimidazoles, Imidazothiazole, Spiroindoles, macrocyclic lactones, Cyclooctadepsipeptides, Thiazolide (nitazoxanide) and Amino-acetonitrile derivatives (monepantel) (Holden-Dye and Walker, 2014). Therefore, finding a new anthelmintic compound has become a global urgency.

Whilst the search for new candidate anthelmintic drugs is ongoing, marine derived bioactive compounds particularly from the surface-associated microbiota are a promising resource (Penesyan et al., 2010; Romano et al., 2017). Bacteria from the genus Pseudoalteromonas have been recognised for their ability to produce a range of commercial and pharmaceutical-relevant bioactivities against micro- and macrofoulers (Holmstrom and Kjelleberg, 1999; Bernbom et al., 2011; Paulsen et al., 2019) (see section 1.9, Chapter 1). P. tunicata D2 is arguably the most comprehensively studied microorganism within the genus owing to its low and high-molecular weight compounds

60 that are toxic against a wide spectrum of environmentally and medicinally-relevant organisms including nematodes (Egan et al., 2002; Franks et al., 2006; Ballestriero et al., 2010) (see section 1.8, Chapter 1). Functional screening of a P. tunicata D2 genomic fosmid library discovered a heterologous clone (designated HG8) that was toxic against the nematode C. elegans (Ballestriero et al., 2010). Genetic characterisation of HG8 revealed that a gene of unknown function (NCBI GeneBank Accession Number ZP_01132244.1) was partly responsible for its antinematode activity (Ballestriero et al., 2010). However, heterologous expression of ZP_01132244.1 alone did not show significant antinematode activity (Ballestriero et al., unpublished data), suggesting that other genetic determinants are required for the nematode killing activity.

Given the need for new antinematode drugs, this chapter aims to further characterise and identify the bioactive agent responsible for the antinematode activity of the clone HG8. Screening of the HG8 transposon mutant library identified a mutant of the HG8 clone (designated 7C8) showing attenuated antinematode activity due to the disruption of an uncharacterised gene designated hp1 (NCBI GeneBank Accession Number ZP_01132246.1). A 25.17 kDa Nematode killing protein-1 (Nkp-1) was expressed by the hp1 gene and abundantly produced by the HG8 clone. I further show that activity of Nkp- 1 is likely to be specific in targeting nematodes as exposing the Nkp-1 expressing clones to a range of microorganisms did not impact their growth. The discovery of this novel antinematode protein will hopefully contribute to the development of new therapies that can be applied to parasitic nematode control management in the future.

61

2.2 MATERIALS AND METHODS

2.2.1 Bacterial strains and culture condition

All bacterial strains and vectors used in this study are listed in Table 2.1. Unless otherwise stated, E. coli strains and human bacterial isolates were grown in Lysogeny Broth (also known as Luria Bertani broth (LB10, Appendix II) and Nematode Growth Media (Brenner, 1974) (also known as NGM, Appendix II) at 37°C. Marine bacteria were cultivated in Marine Broth 2216 (Difco Laboratories, Maryland) at 25°C. Yeast and fungal isolates were grown on Potato Dextrose Agar (Oxoid, Australia) at 25°C. Solid media was prepared with the addition of 1.5% (w/v) of agar (Oxoid, Australia). For long term storage, all bacterial strains were kept in 30% (v/v) glycerol at -80°C. Where required, L-(+)-Arabinose (0.2% w/v) and antibiotics such as chloramphenicol (12.5 µg/mL), ampicillin (50 µg/mL) and kanamycin (50 µg/mL) were incorporated into the media.

Table 2.1 Bacterial strains, vectors and P. tunicata ORFs used in this study

Reference or Strain or vector Relevant characteristic or genotype source Strain E. coli EPI300-T1R F-mcrA ∆ (mrrhsdRMSmcrBC) Epicentre ɸ80dlacZ∆M15∆lacX74 recA1 endA1 araD139 ∆(ara, leu) 7697 galU galK λ-rpsL nupG trfA tonA dhfr EPI300 F– λ– mcrA Δ(mrr-hsdRMS-mcrBC) Epicentre Φ80dlacZΔM15 Δ(lac)X74 recA1 endA1 araD139 Δ(ara, leu)7697 galU galK rpsL (StrR) nupG' trfA dhfr OP50 Uracil auxotroph (Brenner, 1974) DH5α ϕ80dlacZΔM15 Δ(lacZYA- argF)U169 recA1 endA1 hsdR17 CMSI, UNSW (rk– mk+) supE44 thi-1 gyrA relA1, carrying pBAD24 vector HP1 EPI300 transformed with pBAD24hp1 (NCBI Accession: ZP_01132246.1) This study HP2 EPI300 transformed with pBAD24hp2 (NCBI Accession: ZP_01132245.1) This study

62

HP1/HP2 EPI300 transformed with pBAD24hp1/hp2 (NCBI Accession: This study ZP_01132246.1/ZP_01132245.1) BD24 EPI300 transformed with empty pBAD24 vector This study HG8 EPI300-T1R transformed with pCC1FOS™ :: (Ballestriero et ZP_ 01132230.1 to ZP_ 01132246.1 al., 2010) 20G8 Clone isolated from a metagenomic library that (Penesyan et was non-toxic to nematodes and produced purple al., 2013) pigmentation under L-(+)-Arabinose induction 7C8 HG8 transposon mutant library showing mutation (Daim, 2012) at hp1 7C8::hp1 HG8 transposon mutant in hp1 complemented This study with pBAD24hp1 7C8::hp2 HG8 transposon mutant in hp1 complemented This study with pBAD24hp2 7C8::hp1/hp2 HG8 transposon mutant in hp1 complemented This study with pBAD24hp1/hp2 7C8::pBAD24 HG8 transposon mutant in hp1 complemented This study with empty pBAD24 vector Human bacterial isolates Staphylococcus Clinical isolate SOVS, UNSW epidermidis 019 Staphylococcus Clinical isolate SOVS, UNSW aureus 6538 Escherichia coli Clinical isolate SOVS, UNSW 008

Serratia Clinical isolate American marcescens Type Culture ATCC 13880 Collection (ATCC®) Pseudomonas Clinical isolate American aeruginosa Type Culture ATCC 9027 Collection (ATCC®) Marine bacterial isolates Aquimarina sp.AD1 Marine bacterium isolated from seaweed CMSI, UNSW Aquimarina sp.BL5 Marine bacterium isolated from seaweed CMSI, UNSW Nautella italica Marine bacterium isolated from seaweed CMSI, UNSW R11 Phaeobacter Marine bacterium isolated from seaweed CMSI, UNSW gallaeciensis LSS9 sp. Marine bacterium isolated from seaweed CMSI, UNSW BL110 Agarivorans sp. Marine bacterium isolated from seaweed CMSI, UNSW BL7

Fungal or yeast isolatesb Penicillium sp. Marine fungal isolate CMSI, UNSW Aspergillus unguis Marine fungal isolate CMSI, UNSW

63

Galactomyces Marine fungal isolate CMSI, UNSW geotrichum Bloxamia sp. Marine fungal isolate CMSI, UNSW Pleosporales sp. Marine fungal isolate CMSI, UNSW Unknown yeast Marine yeast isolate CMSI, UNSW Vectors pCC1FOS™ :: ZP_ Fosmid backbone for genomic library of (Burke et al., 01132230.1 to Pseudoalteromonas tunicata D2 carrying a wild 2007; ZP_01132246.1a type D2 insert (13.8 kb) expressing putative Ballesteriero et antinematode activity, Cmr al., 2010) pCC1FOS™ :: ZP_ 01132230.1 to HG8 fosmid mutated by EZ-Tn5™ transposon on (Daim, 2012) ZP_01132246.1 P. tunicata wild type gene ZP_01132246.1, Cmr, with EZ-Tn5™ Δ Kanr ZP_01132246.1a

pBAD24a F-, Δ(argF-lac)169, (Guzman et φ80dlacZ58(M15), glnX44(AS), λ- al., 1995) , rfbC1, gyrA96(NalR), recA1, endA1, spoT1, thiE 1, hsdR17, pBAD24 pBAD24hp1a P. tunicata D2 wild type gene hp1 (NCBI This study Accession: ZP_01132246.1) cloned downstream the pBAD promoter, Ampr pBAD24hp2a P. tunicata D2 wild type gene hp2 (NCBI This study Accession: ZP_01132245.1) cloned downstream the pBAD promoter, Ampr pBAD24hp1/hp2a P. tunicata D2 wild type genes hp1 and hp2 This study (NCBI Accession: ZP_01132246.1 and ZP_01132245.1) cloned downstream the pBAD promoter, Ampr

a Inducible expression with the presence of L-Arabinose 0.2% (w/v) b Fungal strains were isolated by Ms. Juliana Ferrari, School of Biological, Earth and Environmental Sciences (BEES), University of New South Wales (UNSW), Sydney, Australia CMSI denotes the Centre for Marine Science and Innovation, UNSW SOVS denotes the School of Optometry and Vision Science, UNSW

2.2.2 Maintenance and synchronisation of the nematode C. elegans

C. elegans N2 strain Bristol (Brenner, 1974) were maintained on NGM seeded with E. coli OP50 as the food source (Brenner, 1974) and synchronised according to Stiernagle (2006) with modifications as described by Ballestriero et al. (2010). NGM plates containing mixed population of C. elegans with gravid hermaphrodite adults were washed

64 with M9 buffer (Appendix II) and pooled. Next, 5 mL of this nematode solution were transferred to a new tube and 4 mL of fresh bleaching solution (Appendix II) were added to release the eggs from the gravid nematodes. After five minutes, the reaction was stopped by adding 50 mL of S-basal buffer (Appendix II) and centrifuging at 1300 x g for 3 minutes. The supernatant was removed and the pelleted eggs were washed twice using an equal volume of S-basal buffer (Appendix II). After the final wash, the eggs were transferred onto the NGM plates and incubated at 25°C until L1 larvae hatched. Then, 100 µL of E. coli OP50 bacterial culture was supplemented to the NGM plates to support the growth of C. elegans L1 larvae to the L4 stage.

2.2.3 DNA sequence analysis and annotation

The full HG8 fosmid sequence was downloaded from the NCBI (National Center for Biotechnology Information) database (Accession numbers: from ZP_01132230.1 to ZP_01132246.1) and analysed for homologous sequences against the NCBI, IMG/MER (Integrated Microbial Genomes & Microbiome Samples) (Chen et al., 2017) and UniProt (UniProt Consortium, 2018) using BLAST (Basic Local Alignment Search Tool) (Altschul et al., 1997). The conserved domain and family motifs of each gene were determined through the Conserved Domain Database (CDD NCBI) (Marchler-Bauer et al., 2017), the SMART database of protein domains, families and functional sites; (Schultz et al., 1998; Letunic et al., 2015) and the Pfam database (Finn et al., 2016). The multiple sequence alignment of the antinematode protein was analysed using T-Coffee (Di Tommaso et al., 2011) and displayed using Boxshade (ExPASy) (available at https://embnet.vital-it.ch/software/BOX_form.html).

2.2.4 Generation of HG8 mutant libraries1

Transposon mutagenesis of the HG8 fosmid clone was performed using the EZ- Tn5™ promoter Insert Kit (Epicentre) as per manufacturer’s instructions.

1 This section was performed by Ms. Malak Daim, Honours student, School of Biotechnology and Biomolecular Science (BABS), University of New South Wales (UNSW), Sydney, Australia. 65

Briefly, 0.01 pmol of EZ-Tn5™ transposon was mixed with 0.013 pmol HG8 fosmid DNA and electroporated into the electrocompetent E. coli Phage T1- Resistant TransforMax™ EPI300™-T1R (Epicentre) using the BioRad Micropulser. Cells were recovered in SOC media (Appendix II) for 2 hours at 37°C and screened for successful mutants on LB10 agar supplemented with chloramphenicol (12.5 µg/mL) and kanamycin (50 µg/mL). At least 800 mutant clones were screened to ensure the 3x coverage of all ORFs encoded on the HG8 fosmid. Selected mutants were grown overnight at 37°C, 60 rpm in 96-wells plates (Becton Dickinson Co Biosciences) containing LB10 broth with antibiotics. Glycerol (30 % v/v) was supplemented to each well and the 96-well plates were kept at -80°C.

2.2.4.1 Transposon library screening2

Frozen stocks of HG8 transposon mutant libraries were replicated onto omnitory plates (Nunc brand, Germany) containing LB10 agar with antibiotics and L-(+)-Arabinose (0.2% w/v) (Sigma-Aldrich) using a 96 Solid Pin Multi-Blot Replicator tool (V&P Scientific Inc.). After 48 hours of incubation at 25°C, the synchronized L4 stage C. elegans (30-40 worms) were transferred onto the plate for selective grazing against the mutant libraries in triplicate. Complete colony grazing after 4 days indicated the mutant as no longer toxic against the nematode and it was thereafter subjected to nematode killing assay as described in section 2.2.9. The HG8 clone and a randomly selected non-toxic strain from the original P. tunicata fosmid library (BG6) were used as the positive and negative control respectively as described by Ballestriero et al. (2010). Next, the HG8 mutants resulting in > 60 % C. elegans survival in the nematode killing assay (section 2.2.9) were sequenced bi-directionally for transposon mapping using the EZ-Tn5™ promoter Insert Kit (Epicentre) Forward (5’ ACCTACAACAAAGCTCTCATCAACC 3’) and Reverse (5’GCAATGTAACATCAGAGATTTTGAG 3’) primer pairs as described in section 2.2.5.

2 This section was performed by Ms. Malak Daim, Honours student, School of Biotechnology and Biomolecular Science (BABS), University of New South Wales (UNSW), Sydney, Australia.

66

2.2.5 Sequencing analysis

A standard protocol for Sanger sequencing by the Ramaciotti Centre for Genomics (University of New South Wales, Australia) (https://www.ramaciotti.unsw.edu.au/sequencing/sangersequencing/) was performed with exception to the PCR cycling condition. Briefly, sequencing reaction containing the DNA (~40 ng), Forward or Reverse primer (3.2 pmol), BigDye terminator V3.1 (1 µL), BigDye buffer (3.5 µL) and nuclease-free water up to 20 µL of reaction volume were run under the following thermocycler condition; 95°C for 5 minutes, 99 cycles of 95°C for 30 seconds, 55°C for 10 seconds and 60°C for 4 minutes. The PCR products were then purified by ethanol/EDTA precipitation method (Appendix II) and were vacuum dried for 5 minutes. DNA sequencing was performed by the Ramaciotti Centre for Genomics, UNSW using the Applied Biosystems 3730 DNA Analyser. All sequencing results were submitted for BLAST (Basic Local Alignment Search Tool) (Altschul et al., 1997) analysis and nucleotide alignment through NCBI (National Center for Biotechnology Information) and IMG/MER (Integrated Microbial Genomes & Microbiome Samples) servers (Chen et al., 2017).

2.2.6 Cloning of P. tunicata gene inserts into the pBAD24 vector

The cloning procedures of P. tunicata gene insert into the pBAD24 vector was performed as indicated in Figure 2.1A. In brief, P. tunicata DNA were generated through PCR using the HG8 fosmid as the template. PCR reactions containing Forward and Reverse primers pairs for each gene (10 µM, 1.25 µL of each) (Table 2.2), HG8 fosmid (40 ng), Phusion® High-Fidelity PCR Master Mix with HF Buffer (NEB) (25 µL) and nuclease-free water (Ambion®) up to 50 µL of reaction volume were run under the thermocycler condition of 98 ºC for 30 seconds, 30 cycles of initial denaturation at 98 ºC for 30 seconds, annealing at 60 ºC for 30 seconds and extension at 72 ºC for 30 seconds and a final extension at 72 ºC for 7 minutes. The amplicons were purified using the QIAquick® PCR Purification Kit (Qiagen) as per manufacturer’s protocol and quantified using Qubit ® 3.0 fluorometer (Thermo Fisher Scientific, Australia). The purified amplicons were

67 phosphorylated using the T4 Polynucleotide Kinase (NEB) prior to blunt-ended ligation to the pBAD24 vector. The vector was initially extracted from the overnight culture of E. coli DH5α using the Purelink Quick Plasmid Miniprep Kit (Invitrogen). The vector backbone was linearized using the NcoI restriction enzyme (NEB) through a unique cut at C▼CATGG (1318 bp to 1323 bp) restriction site, blunt-end repaired by the DNA Polymerase I, Large (Klenow) Fragment (NEB) and dephosphorylated at the 5’ end using the Shrimp Alkaline Phosphatase (rSAP) (NEB) as per manufacturer’s instructions. The phosphorylated amplicons were ligated alongside the pBAD24-NcoI-Klenow-rSAP treated vector using the T4 DNA Ligase (NEB) and used for DNA transformation in section 2.2.6.1.

2.2.6.1 Transformation of pBAD24 recombinant vector into EPI300 and 7C8 E. coli strains and screening against C. elegans

The pBAD24 recombinant vectors were transformed into the prepared electrocompetent E. coli EPI300 cells as described by Seidman et al. (1997) (Figure 2.1A). Transformants were recovered in SOC media for 45 minutes at 200 rpm, 37°C and screened for successful recombinant clones on LB10 agar with 50 µg/mL ampicillin. Positive clones with correct insert DNA orientation were established by colony pick PCR using the listed primer pairs (Table 2.2). A correct orientation of DNA insert was defined when the start codon was located downstream of the AraBAD promoter region (Figure 2.1B). The recombinant vectors were extracted and submitted for sequencing analysis by the Ramaciotti Centre for Genomics, UNSW, Australia as described in section 2.2.5. Following sequencing, the recombinant vectors were transformed into the initially prepared electrocompetent 7C8 mutant cells (Seidman et al., 1997) and recovered in SOC media as previously described. Successful transformants were screened on LB10 agar with chloramphenicol, kanamycin, and ampicillin and further validated by colony pick PCR using the primer pairs as described in Table 2.2. The successful EPI300 recombinants and the complemented 7C8 E. coli mutants were challenged against C. elegans using the nematode killing assay as described in section 2.2.9.

68

Table 2.2 List of primer pairs used in this study. The P. tunicata gene start codon within the primer sequence was underlined whilst the primers that bind to the pBAD24 vector were shown in bold.

Amplification Insert size of P. tunicata Gene size Forward primer (5' - 3') Reverse primer (5' - 3') (bp) target gene hp1 708 hp1_F; ATGAGTACTACAATTTG GAACGATG hp1_R; CTGGCTTGTTATCGCCATTT 753 hp2 831 hp2_F; CTAGTGTCGCCCTCATCG hp2_R; CCCCAATATGTTCGAATCCA 862 hp1/hp2 1539 hp1_F; ATGAGTACTACAATTTG GAACGATG hp2_R; CCCCAATATGTTCGAATCCA 1744

Determination of correct insert orientation hp1 708 hp1_F; ATGAGTACTACAATTTG GAACGATG pBAD24_R; CTGGCAGTTCCCTACTCTCG 995 pBAD24_F; ATGCCATAGCATTTTTATCCA hp1_R; CTGGCTTGTTATCGCCATTT 868 hp2 831 hp2_F; CTAGTGTCGCCCTCATCG pBAD24_R; CTGGCAGTTCCCTACTCTCG 1104 pBAD24_F; ATGCCATAGCATTTTTATCCA hp2_R; CCCCAATATGTTCGAATCCA 977 hp1/hp2 1539 hp1_F; ATGAGTACTACAATTTG GAACGATG pBAD24_R; CTGGCAGTTCCCTACTCTCG 1906 pBAD24_F; ATGCCATAGCATTTTTATCCA hp2_R; CCCCAATATGTTCGAATCCA 1780

69

A

P. tunicata gene amplification pBAD24 extraction

5’ 3’ pBAD24 E. coli DH5α Purification

Plasmid linearization

pBAD24 NcoI

Phosphorylation

5’ 3’ Blunt-end repair

5’ 3’ Ligation and transformation

Dephosphorylation pBAD24 5’ 3’

E. coli EPI300

B pBAD24 vector P. tunicata hp1

5’ 3’

3’ 5’

AraBAD promoter Start codon

Figure 2.1 Schematic diagram of the gene cloning protocol (A) The specific P. tunicata D2 gene of interest was amplified by PCR, purified and then phosphorylated using the T4 polynucleotide kinase. The extracted pBAD24 vector was linearized using the NcoI restriction enzyme, blunt-end repaired using the DNA polymerase and dephosphorylated using the Shrimp Alkaline Phosphatase (rSAP). P. tunicata D2 gene insert was then ligated alongside the pBAD24 backbone followed by transformation into the heterologous host E. coli EP300. (B) The correct orientation of the P. tunicata D2 gene insert with the start codon (highlighted in red box) located downstream of the AraBAD promoter (shaded in blue) was confirmed by PCR and sequencing.

70

2.2.7 Protein extraction

The E. coli clones HP1, HG8 and BD24 were cultured in 2 x Yeast Tryptone (2YT) broth (Appendix II) with appropriate antibiotics. The 2YT culture media was seeded with a 0.6% (v/v) bacterial overnight culture, grown shaking (200 rpm) at 37°C until the OD600nm reached 0.6 and thereafter heterologous gene expression induced with L- Arabinose shaking (200 rpm) at 25°C for 18 hours. Cells were harvested via centrifugation at 6000 x g for 7 minutes in 4oC, washed with equal volume of ice-cold phosphate buffer saline (PBS) (Appendix II) and kept in -80°C prior to the extraction. Each of the cell pellets (~1.5 g) were resuspended in 10 mL PBS and mechanically disrupted through ultra-sonication (Consonic, Australia) on ice for three minutes at 50% amplitude with 2 seconds on and off interval. The resulting cell lysate was spun at 10 000 x g for 30 minutes at 4°C. The supernatant containing soluble cell fraction was removed and stored at 4°C until further use. The remaining pellets (containing insoluble proteins) were washed twice each with ice cold Wash buffer A (Appendix II) and Wash buffer B (Appendix II) before the final wash twice with PBS at 4°C. The insoluble protein was resuspended in 10 mL PBS and assayed for antinematode activity using the 24-well microtiter plate assay method as described in section 2.2.13. Total protein content in each fraction was analysed using the Lowry method (Appendix II) (Waterborg, 2009) and quantified based on the standard curve generated from a serially diluted Bovine Serum Albumin (BSA) (Appendix III) (Waterborg, 2009).

2.2.8 SDS PAGE, mass spectrometry and protein modelling

SDS PAGE was performed according to Laemmli’s method (Laemmli and Favre, 1973; Cleveland et al., 1977). Briefly, the soluble and insoluble protein extracts of E. coli clones HP1, HG8 and BD24 were heated in 4x sample loading buffer (Appendix II) at 80°C for 10 minutes and cooled down to ambient temperature. Protein samples were run through a 12% Polyacrylamide precast gel (BioRad) in the Mini- PROTEAN® System Electrophoresis Cells (BioRad) containing the Running buffer (Appendix II) at 120 V for 90 minutes. Polyacrylamide gels were fixed and stained

71 using the Coomassie blue staining (Appendix II) and de-stained using acetic acid/methanol de-staining solution (Appendix II). Protein bands of interest were excised from the gel and sent to the Bioanalytical Mass Spectrometry Facility (BMSF), University of New South Wales, Australia for mass-spectrometry analysis and protein identification. Sequence obtained were analysed for protein BLAST (Basic Local Alignment Search Tool) (Altschul et al., 1997) and nucleotide alignment through NCBI (National Center for Biotechnology Information). Advanced protein modelling, prediction of ligand binding sites and intensive protein analysis by computational method was performed using Phyre2 (Kelley et al., 2015) and SWISS- MODEL (Waterhouse et al., 2018).

2.2.9 Nematode killing assay

Assay plates were prepared by spreading the overnight culture of test bacteria (30 µL) onto 35 mm × 10 mm plates (SARSTEDT) containing LB10 agar with appropriate antibiotics and L-(+)-Arabinose (0.2 % w/v). Plates were incubated at 25°C for four days. Following incubation, 30 to 40 synchronised L4 stage C. elegans individuals were transferred onto the test bacterial lawn and incubated at 25°C. The proportion of live nematodes were recorded every 24 hours under the dissecting stereomicroscope (Olympus SZ-CTV). Nematodes were considered dead when they showed no response to touch (using sterilised worm picker). A test bacterial strain was considered toxic to C. elegans when more than 50% of the assayed nematodes were killed within 3 days (Nappi, 2019). Experiments were performed in triplicates and monitored until the maximum lifespan of the non-toxic control either E. coli OP50, E. coli EP300 or E. coli BD24 was reached (usually 18 to 20 days). C. elegans exposed to E. coli HG8 were used as the positive control. Statistical analysis on nematode survival was performed using the log-rank (Mantel-Cox) method (Mantel, 1966; Harrington, 2005) with the GraphPad Prism software 8.3.0 (GraphPad Software, La Jolla, CA, USA). A One-way ANOVA followed by Tukey’s pairwise comparison test was used to determine the difference of C. elegans survival in the nematode killing assay. Results are presented as the means ± standard error from triplicate samples. A p-value < 0.05

72 was considered to be significant in line with other recent studies by Ballestriero et al. (2010), Souza et al. (2019) and Büchter et al. (2020).

2.2.10 Preparation of heat-killed bacterial clones

A nematode killing assay was performed as described in section 2.2.9 with the exception that the cultivated bacterial lawns on the assay plates were heat killed by exposing the bacterial lawns to 65°C for 1 hour and allowing them to cool down to 20°C prior to the experiment. To confirm non-viability of the heat-killed bacteria, cells were streaked onto LB10 agar with appropriate antibiotics before and on the last day of the assay. After the experiment, C. elegans were mechanically disrupted using sterilized pestle in 200 µL M9 buffer and the nematode lysate was spread plated onto the LB10 agar plates with appropriate antibiotics. The bacterial cells were confirmed as non-viable when there was no bacterial growth observed on the agar plate after 24 hours of incubation in 37°C.

2.2.11 Nematode egg hatching assay

Assessment of C. elegans egg hatching after exposure to E. coli clones HP1, HG8 and BD24 was performed according to Sant'anna et al. (2013). Nematode eggs were harvested using synchronisation of the gravid hermaphrodites (Stiernagle, 2006) (see section 2.2.2). After the final washing step, eggs were pelleted by centrifugation at 1000 x g for 10 minutes and resuspended in 1 mL of M9 buffer. Next, ~ 60 eggs were inoculated onto each assay plate (seeded with a test bacterial strain according to section 2.2.9) and incubated at 25°C for 7 hours. Numbers of L1 hatchlings on each plate were quantified under the dissecting microscope. The experiment was performed in triplicate. Results were analysed with Prism software version 8.3.0 (GraphPad Software, La Jolla, CA, USA) using a One-way ANOVA and Tukey’s pairwise comparison test. A p value < 0.05 was considered significant

73

2.2.12 Nematode brood size assay

Assessment of C. elegans brood size after exposure to E. coli clones HP1, HG8 and BD24 was performed according to Bischof et al. (2006). A single L4 stage C. elegans was transferred onto a nematode killing assay plate containing a lawn of the test bacterial strain (see section 2.2.9). After incubation at 25°C for 72 hours, the number of nematode progenies arising from an individual was quantified. The experiment was performed in triplicate (Bischof et al., 2006). Results were analysed with Prism software version 8.3.0 (GraphPad Software, La Jolla, CA, USA) using a One-way ANOVA and Tukey’s pairwise comparison test. A p value < 0.05 was considered significant.

2.2.13 Nematode killing assay using 24-well microtiter plates

C. elegans survival against the E. coli clones, the bacterial cell-free supernatant and the protein extracts (obtained from section 2.2.7) were carried out separately in 24- well microtiter plates as reported by Huang et al. (2014) and Wei et al. (2017) with modifications. Wells were incorporated with 90 µL of bacterial cells (washed with

PBS, re-dissolved in LB10 broth and standardised to OD600 nm = 2.0), appropriate antibiotics, L-(+)-Arabinose and M9 buffer up to 300 µL of final volume. To test activity of the cell-free supernatants, wells were incorporated with 90 µL of cell-free supernatant, 40 µL of concentrated E. coli OP50 (OD600nm = 2.0) and M9 buffer up to 300 µL. To test soluble and insoluble protein fractions,150 µL of each protein extract (equivalent to ~0.7 mg of total protein content) were placed in individual wells supplemented with 40 µL of concentrated E. coli OP50 (washed with PBS and re- dissolved in PBS, OD600nm = 2.0) and M9 buffer up to a final volume of 300 µL. Twenty to thirty L4 stage nematodes were transferred into each well of the 24-well microtiter plate and incubated in 25°C for 72 hours. The proportion of live nematodes were recorded every 24 hours. Nematodes were considered dead when they showed no response to touch (using sterilised worm picker). Experiment was conducted in triplicate and data was presented as the percentage of nematode survival over time of

74 assay. Statistical analysis on nematode survival was performed using the log-rank (Mantel-Cox) method (Mantel, 1966; Harrington, 2005) with the GraphPad Prism software 8.3.0 (GraphPad Software, La Jolla, CA, USA). A One-way ANOVA followed by Tukey’s pairwise comparison test was used to determine the difference of C. elegans survival in the nematode killing assay. Results are presented as the means ± standard error from triplicate samples. A p-value < 0.05 was considered to be significant (Huang et al., 2014).

2.2.14 Antibacterial and antifungal assay

Antibacterial and antifungal assays using a disc diffusion method (James et al., 1996; Puskarova et al., 2017) were performed to determine if the E. coli clones toxic to C. elegans had a broad spectrum of antimicrobial activity. Briefly, LB10 agar with L- Arabinose were seeded with the overnight culture of human bacterial isolates or fungal or yeast suspension (Table 2.1). For fungal or yeast isolates, suspension were prepared by scratching the 3-4 days-old fungal mycelia or yeast cells grown on the Potato Dextrose Agar (Oxoid) (Appendix II) and mixed with 10 mL sterilized water. The fungal spores and yeast suspension were adjusted to 1 x 106 conidia/mL using the Neubauer’s chamber and spread onto the LB10 agar plates added with antibiotics and L-(+)-Arabinose using sterilised cotton swabs. Blank discs (CTO998B, Oxoid) were overlaid onto the seeded LB10 agar. A 10 µL aliquot of the test bacterial suspension

(washed with PBS and resuspended in LB10 broth, OD600nm = 2.0) was inoculated onto each disc. For marine bacterial isolates that do not grow on LB10, a modified protocol was used. Briefly, fresh LB10 agar plates with L-Arabinose were overlaid with blank discs inoculated with the test bacterial suspension (10 µL). Soft marine agar 2216 (Difco™) (Appendix II) containing the marine target strains (Table 2.1) (0.4 mL bacteria/ 3 mL soft agar) was overlaid onto the LB10 agar (James et al., 1996) and solidified at room temperature. All of the assay plates were incubated at 25°C for 48 to 72 hours. Blank discs inoculated with chloramphenicol (50 µg/ mL) or ketoconazole (180 µg/mL) and LB10 broth were used as the positive and negative control respectively. E. coli clone 20G8 (Table 2.1) that produces a purple pigmentation under the control of the same inducible promoter was used as an indicator of successful L-

75

(+)-Arabinose induction. The antibacterial and antifungal activity of the test clones was assessed via the visualisation of growth clearing zone around the disc (James et al., 1996; Puskarova et al., 2017).

2.3 RESULTS

2.3.1 Annotation of the 13.8 kb P. tunicata D2 DNA fragment encoded in the HG8 fosmid

The majority of the predicted ORFs encoded in the 13.8 kb P. tunicata D2 DNA fragment (Figure 2.2, Table 2.3) were annotated as conserved hypothetical proteins. Only three of the 17 ORFs (namely NCBI Accession Number ZP_01132243.1, ZP_01132239.1 and ZP_01132234.1) encode gene products with a match to previously characterised proteins including a 2',3'-cyclic nucleotide 2'- phosphodiesterase/3'-nucleotidase bifunctional periplasmic precursor protein (ZP_01132243.1), probable outer membrane protein A (ompA) (ZP_01132239.1) and Transposase IS66 (ZP_01132234.1). The remaining 14 hypothetical proteins could not be assigned a known function.

However, further investigation identified several conserved domain that might be important for the gene function (Table 2.3). For example, the hypothetical protein ZP_01132244.1 harbours a conserved ABC_transp_aux domain that is found in a number of uncharacterised hypothetical proteins (e.g. hypothetical protein from Borrelia turicatae (strain 91E135) (NCBI Accession: WP_011772685.1) (Porcella et al., 2008) and proteins involved in membrane transport (e.g. ABC_transp_aux domain-containing protein from Pseudomonas aeruginosa MH27 (NCBI Accession: WP_023657096.1) (Tielen et al., 2014) and gliding motility (e.g. gliding motility- associated ABC transporter ATP-binding subunit GldA from Flavobacterium johnsoniae (NCBI Accession: WP_014164802.1) (Agarwal et al., 1997) and Intraflagellar transport protein IFT52 from Chlamydomonas reinhardtii (NCBI Accession: XP_001692161.1) (Taschner et al., 2014). In addition, the hypothetical

76 protein ZP_01132242.1 harbours conserved BON and OsmY domains that are found in other characterised BON-domain containing proteins in several microorganisms e.g. Paraglaciecola sp., Pseudomonas , Rheinheimera baltica, Acidobacteria bacterium and Xanthomonadales bacterium (NCBI Accession; NCT48573.1, WP_110682546.1, WP_027671440.1, RPH59456.1 and RPH96582.1 respectively) (Yeats and Bateman, 2003; Dalcin Martins et al., 2018; Probst et al., 2020). Other ORFs annotated as the hypothetical proteins (ZP_01132233.1, ZP_01132232.1 and ZP_01132231.1) contained conserved domains related to transposase function, including two with homology to predicted transposases encoded in P. ulvae. Additionally, six ORFs; ZP_01132244.1, ZP_01132243.1, ZP_01132242.1, ZP_01132241.1, ZP_01132240.1 and ZP_01132239.1 all showed homology to genes encoding putative membrane bound proteins. Further descriptions on each of the gene are listed in the Table 2.3.

77

HG8 fosmid

13.8 kb P. tunicata insert

Tn

Uncharacterised hypothetical protein 300 bp 2,3-cyclic nucleotide 2-phosphodiesterase / 3-nucleotidase bifunctional periplasmic Probable outer membrane protein (OmpA Family protein) Transposase IS66 Transposon mutation (Tn) Intergenic region

Figure 2.2 Schematic diagram of 13.8 kb P. tunicata D2 insert encoded in the HG8 fosmid. Previous transposon disruption of one of the hypothetical proteins (NCBI Accession Number ZP_01132244.1) diminished the nematode killing activity by HG8 clone (Ballestriero et al., 2010). 78

Table 2.3 Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata gene Protein Query Identified Nucleotide Signal Predicted gene % Accession size Closest homolog covered E-value conserved domain size (bp) peptide function Identity Number (aa) (%) and E-value

Uncharacterised Uncharacterised hypothetical protein ZP_01132246.1 708 235 No hypothetical Vibrio ordalii FS-238 85.53 100 3.1 e-149 No (denoted hp1) protein (IMG Locus Tag: A1QSDRAFT_00964)

Uncharacterised Uncharacterised hypothetical protein ZP_01132245.1 831 276 No hypothetical Vibrio ordalii FS-238 78.8 100 2.40E-159 No (denoted hp2) protein (IMG Locus Tag: A1QSDRAFT_00963)

ABC_transp_aux Uncharacterised (Accession: hypothetical protein pfam09822). E- Uncharacterised Pseudoalteromonas ZP_01132244.1 value: 6.63e-03 2175 724 Yes hypothetical luteoviolacea 79.06 100 0 (denoted hp3) protein DSM6061 (IMG P-loop_NTP Locus Tag: (Accession: Ga0128231_114841) cl21455). E-value: 9.18e-04

79

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

Identified P. tunicata Protein Query Nucleotide Signal Predicted gene % conserved gene Accession size Closest homolog covered E-value size (bp peptide function Identity domain and E- Number (aa) (%) value UshA (2',3'-cyclic- nucleotide 2'- phosphodiesterase/ 5'- or 3'- nucleotidase, 5'- nucleotidase 2',3'-cyclic-nucleotide family (Accession: 2',3'-cyclic 2'-Phosphodiesterase / COG0737). E- nucleotide 2'- 3'-nucleotidase value: 5.44e-136 phosphodiesterase/3'- ZP_01132243.1 Pseudoalteromonas 2040 679 Yes nucleotidase 75.57 100 0 CpdB (denoted hp4) luteoviolacea bifunctional (bifunctional 2',3'- DSM6061 (IMG periplasmic cyclic nucleotide Locus Tag: precursor protein 2'- Ga0128231_114840) phosphodiesterase/ 3'-nucleotidase periplasmic precursor protein) Accession: PRK09420. E- value: 0.0

80

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata gene Query Nucleotide Protein Signal Predicted gene % Identified conserved Accession Closest homolog covered E-value size (bp size (aa) peptide function Identity domain and E-value Number (%) OsmY (Accession: COG2823) - Osmotically-inducible protein OsmY, contains BON domain. E- value: 2.87e-22 Periplasmic protein (Accession: PRK10568). BON-domain E-value: 7.53e-17 Uncharacterised containing protein ZP_01132242.1 BON (Accession: 837 278 No hypothetical Paraglaciecola sp. 56.38 98 3.00E-96 (denoted hp5) pfam04972) - BON protein (NCBI Accession: domain; This domain is NCT48573.1) found in a family of osmotic shock protection proteins. E-value: 6.52e-13 LysM domain/BON superfamily protein (Accession: PRK10568). E-value: 4.98e-05

81

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata gene Protei Signal Query Nucleotid Predicted gene % Identified conserved Accession n size peptid Closest homolog covered E-value e size (bp function Identity domain and E-value Number (aa) e (%)

lmo0937 family Membrane protein Uncharacterise ZP_01132241.1 Marinobacter sp. 168 55 No d hypothetical 83.33 87 3.00E-17 No (denoted hp6) Arc7-DN-1) (NCBI protein Accession: WP_117618226.1

DUF 4938 domain- DUF 4398 (Domain of containing protein Uncharacterise unknown function) ZP_01132240.1 Ectothiorhodospira 417 138 No d hypothetical 49.58 86 2.00E-24 (Accession: (denoted hp7) sp. BSL-9 (NCBI protein pfam14346). E- Accession: value: 6.00e-10 WP_063464708.1)

82

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata gene Predicted Query Nucleotide Protein Signal Closest % Identified conserved Accession gene covered E-value size (bp size (aa) peptide homolog Identity domain and E-value Number function (%) OmpA_C - like domain- containing protein (Accession : cd07185). E-value: 3.51e-42 OmpA : Outer membrane protein OmpA and related peptidoglycan-associated (lipo)proteins (Accession : OmpA family COG2885). Probable Nitrosomonas E-value: 6.71e-40 outer sp. ls79A3 ZP_01132239.1 Putative outer membrane 882 293 No membrane (NCBI 46.02 97 3.00E-69 (denoted OmpA) lipoprotein (Accession: protein Accession: PRK10510). (OmpA) WP_0139642 E-value: 1.43e-25 71.1) Pal_lipo: peptidoglycan- associated lipoprotein (Accession: TIGR02802). E-value: 6.87e-25 OmpA: OmpA family (Accession: pfam00691). E-value: 1.32e-22

83

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata Protein Query Nucleotide Signal Predicted gene % Identified conserved gene Accession size Closest homolog covered E-value size (bp peptide function Identity domain and E-value Number (aa) (%)

Transposase Uncharacterised Pseudoalteromonas ZP_01132238.1 7.00E- 201 66 No hypothetical luteoviolacea (NCBI 73.08 39 No (denoted hp9) 04 protein Accession: WP_081215943.1)

Transposase

Uncharacterised Pseudoalteromonas ZP_01132237.1 2.00E- 294 97 No hypothetical ulvae (NCBI 90.48 86 No (denoted hp10) 47 protein Accession: WP_086742174.1) Transposase

Uncharacterised Pseudoalteromonas ZP_01132236.1 7.00E- 201 66 No hypothetical ulvae (NCBI 82.14 84 No (denoted hp11) 22 protein Accession: WP_086743239.1) Uncharacterised hypothetical protein Uncharacterised ZP_01132235.1 Pseudoalteromonas 4.00E- 963 320 No hypothetical 58.33 97 No (denoted hp12) sp. DL-6 (NCBI 122 protein Accession: WP_131691654.1)

84

Table 2.3 (Continue) Homologs and predicted conserved domains of the 17 ORFs in the 13.8 kb P. tunicata D2 HG8 insert

P. tunicata Protein Query Nucleotide Signal Predicted gene % Identified conserved gene Accession size Closest homolog covered E-value size (bp) peptide function Identity domain and E-value Number (aa) (%) IS66 family transposase DDE_Tnp_IS66_C ZP_01132234.1 Transposase Vibrio diabolicus 3.00E- (Accession : (denoted 147 48 No 73.33 93 IS66 (NCBI Accession: 15 pfam13817). E-value: tpsIS66) WP_005398543.1) 2.26e-14 IS66 family transposase DDE_Tnp_IS66 Uncharacterised Pseudoalteromonas sp. ZP_01132233.1 3.00E- (Accession : 162 53 No hypothetical TAB23 (NCBI 84.91 100 (denoted hp14) 23 pfam03050). E- protein Accession: value: 2.00e-11 WP_024608700.1)

Transposase Uncharacterised ZP_01132232.1 Pseudoalteromonas 1.00E- 231 76 No hypothetical 91.3 60 No (denoted hp15) ulvae (NCBI Accession: 21 protein OUL57932.1) Transposase Uncharacterised ZP_01132231.1 Pseudoalteromonas 1.00E- 201 66 No hypothetical 85.71 84 No (denoted hp16) ulvae (NCBI Accession: 22 protein WP_086743237.1)

Uncharacterised

Uncharacterised hypothetical protein ZP_01132230.1 1.00E- 279 92 No hypothetical Pseudoalteromonas 85.39 96 No (denoted hp17) 45 protein ulvae (NCBI Accession: WP_086743969.1)

85

2.3.2 Generation of transposon mutants in HG8 with attenuated nematode killing activity

A total of 960 HG8 transposon mutant clones were successfully generated and screened for diminished toxic activity against C. elegans. In total, 69.8% (670 mutants) were not grazed by nematodes, 24.7% (237 mutants) were grazed at least in one replicate and the remaining 5.5% (53 mutants) were completely grazed in all three replicates (Figure 2.3A).

To confirm loss of antinematode activity for grazed HG8 transposon mutants, 44 randomly chosen mutants from those that were grazed in at least one replicate were assessed for toxic activity using the nematode killing assay. Approximately, half (21/44) of these mutants were considered non-toxic with > 60 % nematode survival after 5 days. Five of the confirmed non-toxic mutants (denoted 8E10, 7C8, 1B11, 3B4 and 8D11) were randomly selected for transposon mapping. Sequencing of the DNA region directly adjacent to the transposon insertion site revealed that mutant 8E10 was disrupted in the gene ZP_01132244.1, whilst mutants 7C8, 1B11, 3B4 and 8D11 were all disrupted in the neighbouring gene (NCBI Accession; ZP_01132246.1) (Figure 2.3B). Genes ZP_01132244.1, ZP_01132246.1 and the adjacent ORF; ZP_01132245.1 are annotated as genes encoding uncharacterised hypothetical proteins) whilst ZP_01132243.1 is annotated as 2',3'-cyclic nucleotide 2'- phosphodiesterase/3'-nucleotidase bifunctional periplasmic precursor protein.

86

A 5.5 % Not grazed (toxic) Grazed at least in 1 replicate Grazed in all three replicates 24.7 %

69.8 %

Total = 960 mutants

B

hp4 hp3 hp2 hp1

Figure 2.3 HG8 transposon mutant generation (A) Fraction of HG8 transposon mutants with different levels of toxicity against C. elegans. (B) Transposon mapping of five randomly selected non-toxic mutants (8E10, 8D11, 3B4, 1B11, 7C8) showing attenuated nematode killing activity identified transposon disruption in genes encoding for uncharacterised hypothetical proteins; ZP_01132244.1 (denoted hp3) and ZP_01132246.1 (denoted hp1) (both genes are coloured in blue).

87

2.3.3 Attenuated activity of the 7C8 mutant is restored upon complementation with hp1

Since transposon mapping of mutants 7C8, 1B11, 3B4 and 8D11 demonstrated that these are sister clones (i.e. with the transposon inserted in the same location), only one mutant (7C8) was selected for further analysis. The role of hp1 and hp2 in antinematode activity was confirmed via genetic complementation followed by activity in the nematode killing assay. After 2 days, C. elegans exposed to the mutant clone expressing a functional hp1 in trans had reduced survival compared to those exposed to the 7C8 mutant strain (One-way ANOVA followed by Tukey’s test; F (6, 14) = 155.2; p < 0.0001, Figure 2.4, Supplementary Materials; Table S2.1) and a similar response as those exposed to the original toxic HG8 clone (One-way ANOVA followed by Tukey’s test; F (6, 14) = 155.2; p = 0.0553, Figure 2.4, Supplementary Materials; Table S2.1). In contrast, no significant differences in survival between C. elegans exposed to the mutant 7C8 and 7C8 expressing hp2 was observed (One-way

ANOVA followed by Tukey’s test; F (6, 14) = 155.2; p = 0.5838, Figure 2.4, Supplementary Materials; Table S2.1). While complementation of 7C8 with a plasmid expressing both hp1 and hp2 restored some antinematode activity (One-way ANOVA followed by Tukey’s test; F (6, 14) = 155.2; p = 0.0486, Figure 2.4, Supplementary Materials; Table S2.1), a higher proportion of C. elegans survived exposure to this strain compared to those exposed 7C8 expressing hp1 alone (One-way ANOVA followed by Tukey’s test; F (6, 14) = 155.2; p < 0.0001, Figure 2.4, Supplementary Materials; Table S2.1). These results suggest that hp1 is important for the antinematode activity while the co-expression of hp1 and hp2 may partially suppress the toxic activity of hp1.

88

100 7C8::hp1

7C8::hp2

)

% (

7C8::hp1/hp2

l

a v

i 7C8

v

r u

s 50

7C8::pBAD24

e

d o

t HG8

a m

e OP50 N

0 0 5 10 15 20 25 Day

Figure 2.4. C. elegans survival on the 7C8 mutants complemented with the genes encoding for the uncharacterised hypothetical protein of P. tunicata D2 (hp1, hp2 or hp1/hp2) expressed in trans. Statistical test revealed significant difference among the treatments (Log-rank (Mantel-Cox) test; df = 6; chi square; 70.55; p < 0.0001). A One- way ANOVA and the Tukey’s pairwise comparison test indicate that, after two days, C. elegans survival was significantly reduced when exposed to the 7C8::hp1 compared to the mutant 7C8 (p < 0.0001) and showed no difference to the original toxic clone HG8 (p = 0.0553). In contrast, no significant difference was found as a result of 7C8::hp2 and 7C8 treatments against the nematode (p = 0.5838). A decrease of nematode survival was also shown after exposure to the 7C8::hp1 compared to the 7C8::hp1/hp2 ( p < 0.0001). Each data point represents means of nematode survival ± standard error from triplicate samples.

2.3.4 Expression of hp1 gene in E. coli results in reduced C. elegans survival

To further validate the toxicity expressed by the hp1 gene, E. coli clones expressing hp1 (HP1), hp2 (HP2) or hp1/hp2 (HP1/HP2) were assessed for their ability to reduce C. elegans survival. Exposure to E. coli expressing hp1 after two days resulted in significant reduction in nematode survival compared to nematodes exposed to the E. coli expressing hp2 (One-way ANOVA followed by Tukey’s test; F (5, 12) = 46.62; p = 0.0001, Figure 2.5, Supplementary Materials; Table S2.2). These results validate that expression of hp1 alone is toxic against C. elegans. However, co-expression of hp1and hp2 significantly reduced the toxic activity (One-way ANOVA followed by Tukey’s

89 test; F (5, 12) = 46.62; p = 0.0006, Figure 2.5, Supplementary Materials; Table S2.2). Additionally, the killing activity of the single hp1 gene was slightly lower compared to the original HG8 clone (One-way ANOVA followed by Tukey’s test; F (5, 12) = 46.62; p = 0.0198, Figure 2.5, Supplementary Materials; Table S2.2), suggesting the flanking genes or the protein expression background may provide some auxiliary function required for full toxicity against nematodes.

100 HP1

HP2

)

% (

HP1/HP2

l

a v

i BD24

v

r u

s 50

EPI300

e

d o

t HG8

a

m

e N

0 0 5 10 15 20 25 Day

Figure 2.5 C. elegans survival on E. coli clones HP1, HP2 and HP1/HP2. The HG8 strain was set as the positive control whilst the E. coli BD24 and E. coli EPI300 strains were used as the negative controls. Significant differences among the treatments were detected (Log-rank (Mantel-Cox) test; df = 5; Chi square = 78.57; p < 0.0001). A One- way ANOVA and the Tukey’s pairwise comparison test indicates that, after two days, a significant decrease of survival was observed on the nematodes when exposed to E. coli HP1 compared to E. coli HP2 (p = 0.0001). In contrast, C. elegans exposed to E. coli HP1/HP2 showed an increased survival compared to E. coli HP1 (p = 0.0006). However, the killing activity resulted from the E. coli HP1 was slightly reduced compared to HG8 (p = 0.0198). Each data point represents means of nematode survival ± standard error from triplicate samples.

90

2.3.5 Exposure to E. coli HP1 and HG8 does not affect C. elegans egg hatching efficiency but decreases the brood size

Here I determined the impact of E. coli clones HP1 and HG8 exposure on C. elegans egg hatching and resulting progeny (brood size). There was no significant difference in the proportion of eggs hatching after exposure to E. coli HP1 or HG8 bacterial lawns compared to the non-toxic E. coli BD24 clone (One-way ANOVA followed by

Tukey’s test; F (2, 6) = 0.4181; p = 0.6891 and p = 0.9908 respectively, Figure 2.6A, Supplementary Materials; Table S2.3). However, the L1 larvae fed E. coli clones HP1 and HG8 did not survive past 24 hours. To determine the impact of HP1 and HG8 on nematode brood size, a single C. elegans hermaphrodite was challenged against the toxic E. coli clones and the number of progenies resulting from the single nematode quantified. A significant reduction in progeny was observed with both E. coli HP1 and HG8 strains compared to the number of progenies on the BD24 clone (One-way

ANOVA followed by Tukey’s test; F (2, 6) = 169.7; p < 0.0001, Figure 2.6B, Supplementary Materials; Table S2.3). These findings suggest that E. coli HP1 and HG8 did not affect C. elegans eggs hatching efficiency, however they were toxic to young progeny which results in diminishing C. elegans brood size.

91

A

100

80

)

%

(

g 60 n

i

h

c

t

a h

40

g g E 20

0 1 8 4 P G 2 H H D B

B

40

e

d

o

t a

m 30

e

n

/

s

e

i n

e 20 g

o

r

p

f

o

r

e 10

b m

u N 0 1 8 4 P G 2 H D H B

Figure 2.6 Effect of E. coli clones HP1, HG8 and the negative control BD24 exposure on (A) C. elegans egg hatching efficiency and (B) C. elegans brood size. No significant difference was observed on the proportion of C. elegans eggs hatching on E. coli HP1 and HG8 compared to BD24 (p > 0.05). However, the number of C. elegans progeny on the E. coli HP1 and HG8 bacterial lawn was significantly reduced compared to the BD24 bacterial strain (p < 0.0001). Results shown represent means ± standard error from triplicate samples.

92

2.3.6 Heat-killing treatment significantly reduces the toxic activity of E. coli HP1 and HG8

To determine if alive E. coli HP1 and HG8 are required for nematode killing activity, the toxic bacteria were heat-killed before being assayed against C. elegans. No bacterial growth was recovered from the heat-killed bacterial lawn (before the nematode killing assay) and from the nematode lysate (after the nematode killing assay) indicating that the bacteria were not viable after heat treatment. After two days, the survival of nematodes on the heat-killed E. coli HP1 and HG8 was significantly increased compared to viable E. coli HP1 and HG8 strains (One-way ANOVA followed by Tukey’s test; F (5, 12) = 81.95; p = 0.0001, and p < 0.0001 respectively, Figure 2.7, Supplementary Materials; Table S2.4). These results suggest that living bacterial cells are necessary for continuous killing activity against C. elegans and/or the active compound expressed by the E. coli HP1 and HG8 is heat sensitive.

100 Heat-killed HP1

Heat-killed HG8

)

% (

Heat-killed BD24

l

a v

i HP1

v

r u

s 50

HG8

e

d o

t BD24

a

m

e N

0 0 5 10 15 20 25 Day

Figure 2.7 C. elegans survival on live and heat-killed E. coli clones HP1 and HG8 and the negative control E. coli BD24 bacterial strains. After two days, an increase of survival was observed on nematodes exposed to the heat-killed E. coli HP1 and HG8 compared to nematodes exposed to live bacteria of from the same strain (p = 0.0001, and p < 0.0001 respectively). Each data point represents means of nematode survival ± standard error from triplicate samples.

93

2.3.7 E. coli HP1 and HG8 bioactivities are nematode specific

To determine if E. coli HP1 and HG8 possess broad spectrum inhibitory activities in addition to the nematode toxicity, they were assayed against a range of human and marine isolates including bacteria, fungi and yeast. Purple pigmentation by 20G8 clone indicated a positive L-(+)-Arabinose induction on the test plates. There were no visible signs of growth inhibition surrounding the wells containing the test E. coli strains (Table 2.4 and Table 2.5 respectively). These results suggest that the activity of E. coli HP1 and HG8 is likely not broad range against other microorganisms

94

Table 2.4 Antimicrobial assay of E. coli clones HP1, HG8 and BD24 on human and marine bacterial strains. Positive and negative inhibitory activities are indicated by the “+” and “-” symbols respectively while Cmp denotes chloramphenicol as the positive control. R1, R2 and R3 indicate the assays replicates.

Bacteria HP1 HG8 BD24 LB10 LB10/ LB10/ Human pathogens R1 R2 R3 Cmp 20G8 R1 R2 R3 Cmp 20G8 R1 R2 R3 Cmp 20G8 /MB MB MB Staphylococcus - - - - + Purple - - - - + Purple - - - - + Purple aureus Staphylococcus - - - - + Purple - - - - + Purple - - - - + Purple epidermidis Pseudomonas - - - - + Purple - - - - + Purple - - - - + Purple aeruginosa Serratia marcescens - - - - + Purple - - - - + Purple - - - - + Purple Escherichia coli - - - - + Purple - - - - + Purple - - - - + Purple

Marine isolates Nautella italica R11 - - - - + Purple - - - - + Purple - - - - + Purple Phaeobacter - - - - + Purple - - - - + Purple - - - - + Purple gallaeciensis LSS9 Alteromonas sp. - - - - + Purple - - - - + Purple - - - - + Purple BL110 Aquimarina sp. AD1 - - - - + Purple - - - - + Purple - - - - + Purple Aquimarina sp. BL5 - - - - + Purple - - - - + Purple - - - - + Purple Agarivorans sp. BL7 - - - - + Purple - - - - + Purple - - - - + Purple

95

Table 2.5 Antifungal assay of E. coli clones HP1, HG8 and BD24 on marine fungal and yeast isolates. Positive and negative inhibitory activities are indicated by the “+” and “-” symbols respectively while Keto denotes ketoconazole as the positive control. R1, R2 and R3 indicate the assays replicates.

Fungi and yeast strains HP1 HG8 BD24 Fungi R1 R2 R3 LB10 Keto 20G8 R1 R2 R3 LB10 Keto 20G8 R1 R2 R3 LB10 Keto 20G8 Aspergillus unguis - - - - + Purple - - - - + Purple - - - - + Purple Penicillium sp. - - - - + Purple - - - - + Purple - - - - + Purple Galactomyces geotrichum - - - - + Purple - - - - + Purple - - - - + Purple Bloxamia sp. - - - - + Purple - - - - + Purple - - - - + Purple Pleosporales sp. - - - - + Purple - - - - + Purple - - - - + Purple

Yeast Unknown marine yeast - - - - + Purple - - - - + Purple - - - - + Purple

96

2.3.8 The antinematode compound produced by E. coli HP1 and HG8 is not secreted extracellularly

To determine whether the bioactive compound responsible for the antinematode activity is secreted by E. coli clones HP1 and HG8 to the extracellular environment, cell-free supernatants of these strains were assessed for nematode killing activity. After 48 hours of exposure to E. coli HP1 and HG8 cell-free supernatant, nematode survival was increased compared to the intact bacterial cells treatment (cells were standardised to OD600nm = 2.0 before the nematode killing assay) (One-way ANOVA followed by Tukey’s test; F (6, 14) = 159.1; p < 0.0001, Figure 2.8, Supplementary Materials; Table S2.5). These results suggest that the nematode killing compound is retained intercellularly and is not secreted to the external environment.

100 OP50 cells

HP1 cells

)

% (

HG8 cells

l

a v

i BD24 cells

v

r u

s 50

HP1 supernatant

e

d o

t HG8 supernatant

a m

e BD24 supernatant N

0 0 20 40 60 80 Hours

Figure 2.8 C. elegans survival on E. coli HP1, HG8, BD24 and their corresponding cell-free supernatants. E. coli OP50 and the non-toxic BD24 strains were used as the negative control. Before the assay, the cells were standardised to OD600nm = 2.0. After 48 hours, the nematodes survival was increased after exposure to the cell-free supernatants of E. coli HP1 and HG8 compared to the intact bacterial cells (p < 0.0001). Each data point represents means of nematode survival ± standard error from triplicate samples.

97

2.3.9 E. coli HP1 and HG8 protein extracts are toxic to C. elegans

Soluble and insoluble protein fractions of E. coli HP1 and HG8 were assayed against C. elegans to investigate their potential of nematode killing activity. After 48 hours, treatment against E. coli HP1 and HG8 protein fractions resulted in diminished C. elegans survival with toxic activity more pronounced in the insoluble protein fraction compared to the soluble fraction (One-way ANOVA followed by Tukey’s test; F (6, 14) = 427.2; p < 0.0001, Figure 2.9, Supplementary Materials; Table S2.6). In contrast, C. elegans treated with the soluble and insoluble protein fractions of the non-toxic BD24 control clone demonstrated a normal nematode survival (One-way ANOVA followed by Tukey’s test; F (6, 14) = 427.2; p = 0.9783, Figure 2.9, Supplementary Materials; Table S2.6). These results suggest that the nematode killing activity likely results from a proteinaceous compound associated with E. coli HP1 and HG8 cells. The toxic compound is hereafter denoted as Nematode killing protein-1 (Nkp-1).

100 Control (M9 buffer + OP50)

HP1 soluble protein

)

% (

HG8 soluble protein

l

a v

i BD24 soluble protein

v r

u HP1 insoluble protein

s 50

e

d HG8 insoluble protein

o

t a

m BD24 insoluble protein

e N

0 0 20 40 60 80 Hours

Figure 2.9 C. elegans survival in the soluble and insoluble protein fractions of E. coli clones HP1, HG8 and BD24 using the 24-well microtiter plates. The total protein in each fraction was ~ 0.7 mg/mL. The non-toxic E. coli BD24 protein fractions and M9 buffer added with E. coli OP50 were used as the negative controls. Exposure to the protein extracts of E. coli HP1 and HG8 reduced the nematode survival and the decrease was more pronounced in the insoluble protein compared to the soluble protein in each strain (p < 0.0001). Each data point represents means of nematode survival ± standard error from triplicate samples.

98

2.3.10 SDS PAGE, mass spectrometry analysis and 3D protein modelling of the Nematode killing protein-1 (Nkp-1)

Protein bands corresponding to the estimated size of Nkp-1 (~25.17 kDa) were observed through SDS PAGE in the protein fractions of E. coli HP1 and HG8 strains (Figure 2.10, indicated by red arrows). These protein bands were excised from the gel and identified as Nkp-1 using MALDI-TOF analysis and protein BLAST.

The Nkp-1 peptide sequence (consisting of 235 aa) (Appendix IV) was aligned (Figure 2.11) to the sequences of closely related protein homologs of Nkp-1 (obtained from the BLASTp results, see Table 2.2). These also include a characterised (produced by closely related bacteria from the genus Vibrio i.e. Vibrio cholerae (Class Gammaproteobacteria, Family Vibrionaceae) (Moustafa et al., 2004; O Neal et al., 2005) and a hypothetical protein from Bacillus halodurans (PDB ID: 1W9S) harbouring the carbohydrate binding module (van Bueren et al., 2005). Homology modelling of Nkp-1 using PhyRe2 resulted in a 3D protein model (Figure 2.12B) based on the highest confidence scoring template, a β-1,3-glucan binding CBM6 module (Galactose-binding domain-like, Family 6 carbohydrate binding module) of hypothetical protein BH0236 from B. halodurans (PDB ID: 1W9S) (van Bueren et al., 2005). In total, 88 residues of Nkp-1 from residue 14 to residue 101 (37% of the Nkp- 1 sequence) located in the N-terminal of the protein were modelled with 48% confidence (Figure 2.12A). The Nkp-1 aligned sequence was modelled to have secondary structure consisting of anti-parallel beta strands or β barrel which is composed of tandem repeats that coil and twist to construct a closed toroidal structure. A second 3D model was built using SWISS-MODEL based on the PsCBM35-2 domain containing carbohydrate binding module 35 (CBM35) of α-1,6- glucosyltransferase (classified under the glycoside hydrolase (GH) Family 31 alpha- glucosidase) from Paenibacillus sp. 598K (PDB ID: 5X7O) (Fujimoto et al., 2017) (Figure 2.12C). The alignment of Nkp-1 to the PsCBM35-2 domain of α-1,6- glucosyltransferase resulted in 23.8% identity from residue 16 to residue 105 (Figure 2.12A).

99

Figure 2.10 SDS Polyacrylamide gel (12%) showing soluble and insoluble protein fractions of E. coli clones HP1, HG8 and the control strains; BD24 and EPI300. Key: S; soluble protein fraction, In; Insoluble protein fraction, - ; no L-(+)-Arabinose induction, + ; induced with L-(+)-Arabinose (0.2% w/v). Protein bands (indicated by red arrows) were confirmed as the Nematode killing protein-1 (Nkp-1) encoded by P. tunicata D2 hp1 gene (NCBI Accession: ZP_01132246.1). The Nkp-1 was not detected in the protein fractions of E. coli BD24 or EPI300 control strains as protein bands approximated the size of ~25 kDa found in those bacterial protein extracts were belonged to either TEM extended spectrum beta-lactamase or outer membrane protein A of E. coli.

100

Figure 2.11 Multiple sequence alignment of Nkp-1 protein sequence and closely related protein sequences. These include the uncharacterised hypothetical protein (hp) from V. ordalii and P. aliena and characterised protein sequences; Cholera Subunit A from V. cholerae (denoted as Cholera) and a hypothetical protein BH0236 from B. halodurans (PDB ID: 1W9S) harbouring a β-1,3-glucan binding CBM6 module (van Bueren et al., 2005) (denoted as 1W9S).The total amino acid residues for each of the protein sequences are; Nkp-1 (235 aa), hp V. ordalii (235 aa), hp P. aliena (235 aa), Cholera (258 aa) and 1W9S (1020 aa). The consensus residues in each aligned protein sequence are highlighted in black and grey whilst the identified gaps may indicate deletion or insertion of amino acids. The sequence alignment was created using T-Coffee (Di Tommaso et al., 2011) and displayed using Boxshade (ExPASy).

101

A

B C Model A Model B

C-

C-

N- N-

Figure 2.12 Schematic diagrams representing Nkp-1 protein sequence and the 3D protein models generated by different software. (A) Nkp-1 protein coverage that is used by PhyRe2 and SWISS-MODEL to generate the Nkp-1 3D protein model A and B respectively. (B) Model A Nkp-1 was generated by PhyRe2 using the structure of beta-1,3-glucan binding CBM6 module of hypothetical protein BH0236 from Bacillus halodurans (PDB ID: 1W9S). The model dimensions (Å): X:26.001 Y:40.693 Z:33.630 indicates the permitted rotation of the C-N bond in the polypeptide chain. (C) Model B Nkp-1 was generated by SWISS-MODEL using the structure of PsCBM35-2 domain containing carbohydrate binding module-35 (CBM35) of α-1,6- glucosyltransferase from Paenibacillus sp. 598K (PDB ID: 5X7O). The protein alpha helices are shown as arrows whilst the beta strands are shown as ribbons. The coil and turn protein are indicated in line. The arrowhead points towards the carboxy termini.

102

2.4 DISCUSSION

This chapter aimed to characterise the gene(s) directly or indirectly responsible for the nematode killing activity of HG8. A research pipeline comprising transposon mutations, genetic complementation, in-silico studies and in-vitro screening against C. elegans was performed. To characterise the putative nematode killing compound, protein fractions of the active clones were assessed for toxicity against nematodes and subjected to protein separation and mass spectrometry analysis. Given the 3- dimensional structure being more conserved than the sequence (Illergård et al., 2009; Nadzirin and Firdaus-Raih, 2012), a putative 3D protein model of Nkp-1 was built and analysed in attempt to discover homologous protein with characterised biological potential. In order to assess if the protein was toxic to other microbial targets, a range of environmental and pathogenic microorganisms were screened against the toxic clones. This work has successfully led to several key findings that are further discussed in the following sections.

2.4.1 The gene hp1 is responsible for the expression of Nkp-1; a novel nematode killling protein with a putative carbohydrate binding module

Nine hundred and sixty E. coli mutant clones of HG8 were successfully screened against C. elegans. Given an insert size of 13.8 kb and the smallest ORF of 81 bp, the screening of 960 mutant clones represented greater than a 3x coverage of the DNA insert. Sequencing and gene complementation of non-toxic mutants revealed that expression of a single gene hp1 was sufficient for toxic activity against C. elegans. The hp1 gene encodes for a novel Nematode killing protein (Nkp-1), which is toxic to C. elegans in both soluble and insoluble forms. While the intact E. coli HP1 was active against C. elegans, its cell-free supernatant was non-toxic, suggesting that the concentration of Nkp-1 in the supernatant is too dilute or that it is not secreted to the environment. Other studies have also observed nematode killing compounds that are retained within the cell of the producers. For example, prodigiosin in Serratia marcescens cell extracts was toxic against the plant parasitic nematodes; Radopholus

103 similis and Meloidogyne javanica (Rahul et al., 2014). Additionally, the insoluble nematicidal Cry5B crystal protein was also purified from the intact cells of Bacillus thuringiensis (Charuchaibovorn et al., 2019).

Modelling of the Nkp-1 peptide sequence led to the generation of two protein models (Model A and B) both consisting of one of two different putative carbohydrate binding modules (CBMs) (i.e. CBM6 or CBM35) (Figure 2.13). The structural features of both carbohydrate binding modules are composed of canonical β-sandwich folds which are connected by loops of different length (http://www.cazy.org/) (Cantarel et al., 2008). While the carbohydrate binding specificity of both CBMs are diverse (Correia et al., 2010), CBM6 specifically binds to the β-1,4-glucan (or cellulose), mixed-linked β- 1,3-1,4-glucan (or lichenan), agarose, xylan, β-1,3-glucan (or laminarin) and chitin (Abbott et al., 2009) while CBM35 binds to uronic acids (4,5Δunsaturated deoxyGalA (Pel-CBM35), glucuronic acid (Montanier et al., 2009), β1,4-mannan (Bolam et al., 2004) or α-1,6- and α-1,4-glucans (Suzuki et al., 2014; Fujimoto et al., 2017).

Proteins with carbohydrate binding modules are ubiquitous in nature (Michel et al., 2009). The CBM6 and CBM35 carbohydrate binding modules are found in several enzymes i.e. endoglucanase, β-agarase, α-1,3-glucanase, glucuronoxylanase, esterase and mannanase (Maglione et al., 1992; Sainz-Polo et al., 2014; Zhang et al., 2014; Katsimpouras et al., 2016; Teh et al., 2019; Yano et al., 2019) and bacterial toxins i.e. Crystal toxins (Cry) by Bacillus thuringiensis (Ficko-Blean and Boraston, 2012; Feng et al., 2015; Rodríguez-González et al., 2020). These carbohydrate binding modules enable enzyme binding to the target substrate (Michel et al., 2009) or to toxin binding to the membrane or glycoconjugate receptors of the target host cells (Ficko-Blean and Boraston, 2012; Feng et al., 2015; Berninsone, 2018). As an example, insoluble Cry proteins produced by B. thuringiensis are potent nematicidal and pesticidal toxins that harbour the CBM6 and CBM35 carbohydrate binding module (Feng et al., 2015; Geng et al., 2017; Jing et al., 2019; Rodríguez-González et al., 2020). The Cry toxins generally consist of three structural domains where domain I is responsible for the protein toxicity, domain II is associated with toxin binding to the host glycoconjugate receptor and domain III is important for host receptor recognition as well as for the

104 toxin binding (Figure 2.13) (Bravo et al., 2011; Xu et al., 2014; Feng et al., 2015; Liu et al., 2020).

All protein templates and Nkp-1 models harbouring the CBM6 (Figure 2.13A) and CBM35 modules (Figure 2.13B) consist of an antiparallel β-sandwich protein arrangement. This β-sandwich structure is thought to be key to the specificity of Cry toxin binding to the target nematode and insect’s intestinal cell receptor (Deist et al., 2014; Feng et al., 2015), with studies showing that swapping this structure between different Cry toxins results in altered specificity and toxicity of the toxin towards several target organisms (de Maagd et al., 2000; Naimov et al., 2001; Karlova et al., 2005; Bravo et al., 2013). The β-sandwich structure (domain III) is also important for Cry protein structural integrity and stability due to the network of hydrogen bond and van der Waals interactions between the protein residues and the side chains (Li et al., 1991; López Pazos, 2007; Deist et al., 2014). Given the similar arrangement of antiparallel β-sandwich in Nkp-1, it is possible that these carbohydrate binding modules plays a role in recognition and binding to one of the many glycoconjugate receptors present in the C. elegans intestine (Berninsone, 2018; Jankowska et al., 2018). While the specific Nkp-1-targetting glycoconjugate receptor in C. elegans warrants further investigations, this discovery may indicate the initial mechanism of Nkp-1 killing activity against C. elegans involving attachment to the host membrane receptor prior to the initiation of nematicidal activity.

105

Domain III A

1W9S Model A Nkp-1

Domain I Domain II

Cry1Aa PsCBM35-2 B

Model B Nkp-1 5X7O

alpha-1,6-glucosyltransferase

Figure 2.13 Characterised protein templates showing resemblance to Nkp-1 3D protein models. (A) Antiparallel β-sandwich protein structure harbouring the CBM6 (Pigott and Ellar, 2007) is shown in the domain III of Cry1Aa produced by B. thuringiensis, β-1,3-glucan binding structure of B. halodurans hypothetical protein (PDB ID: 1W9S) and the Model A Nkp-1 from P. tunicata D2. (B) A β-sandwich protein fold harbouring the CBM35 (Sainz-Polo et al., 2014) is displayed by the alpha- 1,6-glucosyltransferase from Paenibacillus sp. 598K and its PsCBM35-2 domain (PDB ID: 5X7O) and the Model B Nkp-1. The CBM35 was previously found in the domain III of Cry1Ie toxin produced by B. thuringiensis (Feng et al., 2015) (picture not shown). Analogous β-sandwich protein fold observed in all of the protein structures may suggest a putative similar protein function. The Cry1Aa protein structure was adapted from (Pigott and Ellar, 2007) with permission whilst 1W9S, 5X70 and alpha-1,6-glucosyltransferase were retrieved from the Protein Data Bank (PDB). The Nkp-1 protein models A and B are generated by PhyRe2 and SWISS Model respectively.

106

2.4.2 Different expression system and heterologous recombinant cell background effects Nkp-1 expression level in E. coli HG8 and HP1 strains

Whilst the expression of hp1 gene alone in the HP1 strain is sufficient for toxicity towards C. elegans, the activity was not as strong as the original HG8 wild type clone. This observation could be linked to a higher Nkp-1 expression in HG8 compared to HP1 indicated by a strong protein band representing Nkp-1 in its insoluble protein fraction (see Figure 2.10). One possible explanation for this finding is the difference between expression system and heterologous recombinant host backgrounds that were required to express hp1 in both HG8 and HP1 strains. The pCC1FOS Fosmid with the strong T7 RNA polymerase promoter system flanking the cloning site was used as the vector backbone to construct the genomic libraries of P. tunicata D2 (i.e. HG8 strain) (Burke et al., 2007; Ballestriero et al., 2010). However, in HP1 strain, hp1 expression is under the control of the less efficient PBAD promoter requiring L-(+)-Arabinose induction (Guzman et al., 1995; Balzer et al., 2013; Ahmad et al., 2018). Attempts were made to increase Nkp-1 production by increasing L-(+)-Arabinose concentrations, however induction above 0.2% (w/v) L-(+)-Arabinose concentration [1.0% (w/v)] resulted in decreasing intensity of the Nkp-1 protein band (Figure S2.1, Supplementary Materials). This might also be linked to the deleterious effect of high arabinose concentration against E. coli EPI300 host cells (Guzman et al., 1995).

Overexpression of Nkp-1 under the T7 promoter in the pET28b(+) vector and the E. coli BL21 (λDE3) recombinant strain was attempted, however this was unsuccessful as indicated by the faint protein bands (Figure S2.2, Supplementary Material). So far, little is known about the intracellular toxicity of Nkp-1 to the producing strain. However, toxicity often associated with an overexpression system may result in a low desired protein productivity (Dumon-Seignovert et al., 2004; Giacalone et al., 2006; Narayanan et al., 2011). In this case, a tightly regulatable expression system i.e. the

PBAD promoter as chosen for the current study is recommended to efficiently control the toxic protein production (Narayanan et al., 2011). Further troubleshooting, molecular cloning (employing different expressing vector, promoters and expressing

107 organisms) and culture condition optimisation are required to further increase Nkp-1 production for future applications.

2.4.3 The products of other genes in HG8 might have an indirect effect on Nkp- 1 toxicity against C. elegans

In addition to the different expression systems and host cell backgrounds between HP1 and HG8, several P. tunicata D2 auxiliary ORFs in HG8 may play a role in mitigating the toxicity of Nkp-1towards C. elegans. Firstly, the concomitant expression of hp1 and hp2 resulted in decreasing toxicity against nematodes compared to the expression of hp1 on its own (Figure 2.3.5). One explanation for this observation is that hp2 may function as an antitoxin-like factor that suppresses Nkp-1 toxicity. Toxin-antitoxin systems are ubiquitous in microorganisms including in the Cry toxin producer; B. thuringiensis (Fico and Mahillon, 2006; Short et al., 2015) and other Pseudoalteromonas species (Kopel et al., 2014; Qi et al., 2019). Interestingly, hp2 does not share homology with any of the classical antitoxin systems as reviewed by (Lee and Lee, 2016), suggesting that HP2 potentially acts as a novel antitoxin protein based on its suppressive activity. Given its predictive role as an antitoxin, HP2 may indirectly facilitate the overexpression of Nkp-1 in the HG8 strain via the temporary suppression of Nkp-1 toxicity towards E. coli in the HG8 cell background. Thus, further explaining the difference in the antinematode activity between E. coli strains HP1 (which does not harbour the hp2 gene) and HG8.

Secondly, several other genes encoded on the HG8 fosmid may enhance Nkp-1 toxicity by promoting the colonisation of Nkp-1 expressing E. coli cells to the C. elegans gut, leading to the nematodes demise. For example, the hp5 (ZP_01132242.1) encoding for a hypothetical protein (HP5) harbouring a conserved Bon (bacterial OsmY and nodulation)-domain enhance E. coli HG8 attachment and colonisation owing to its conserved OsmY (Osmotically-inducible protein Y) and LysM (Lysin Motif) (Yeats and Bateman, 2003) domains. OsmY domain proteins are widespread in (Yan et al., 2019) and functionally important to prevent shrinkage of cytoplasmic cavity as a cellular response to osmotic shock (Zheng et al., 2015).

108

Recent findings also suggest OsmY proteins function as important for virulence factors in several pathogens including Yersinia ruckeri (Mendez et al., 2018), Salmonella typhimurium (Bader et al., 2003), Cronobacter sakazakii (Ye et al., 2015) and E. coli (Dong and Schellhorn, 2009), presumably by protecting the pathogen from osmotic stress during the initial process of infection (Zheng et al., 2015). Additionally, this Bon domain-like protein may facilitate E. coli HG8 attachment to C. elegans body owing to the LysM conserved carbohydrate binding module that enables protein recognition of N-acetylglucosamine (GlcNAc) containing polysaccharides such as chitin (Mesnage et al., 2014), one of the main polysaccharides found on the surface of the C. elegans pharynx (Heustis et al., 2012). In addition to the Bon-domain like protein, colonisation of C. elegans by E. coli HG8, maybe enhanced by the expression of OmpA (ZP_01132239.1), a gene encoding for an outer membrane protein A (OmpA)-like protein. OmpA’s role in mediating host cell colonisation and invasion (including C. elegans) is well described and has been observed for numerous bacterial pathogens (Bartra et al., 2012; Martinez et al., 2014; Kim et al., 2016; Skerniškytė et al., 2019).

Finally, HG8 also harbours genes previously thought to play a role in bacterial persistence. For example, BLASTp analysis revealed that hp3 (ZP_01132244.1) encodes a protein with conserved domains of P-loop NTP (P-loop containing Nucleotide Triphosphate Hydrolases (NTPase) and ATP binding cassette (ABC)_transp_aux. The P-loop NTP is also referred as Walker A and Walker B motif that bind and hydrolyse nucleotides typically ATP or GTP (Leipe et al., 2002). The ATP binding cassette (ABC)_transp_aux domain is found in a number of uncharacterised hypothetical proteins (e.g. hypothetical protein from Borrelia turicatae (strain 91E135) (NCBI Accession: WP_011772685.1) (Porcella et al., 2008) and proteins involved in membrane transport (e.g. ABC_transp_aux domain- containing protein from Pseudomonas aeruginosa MH27 (NCBI Accession: WP_023657096.1) (Tielen et al., 2014) and gliding motility (e.g. gliding motility- associated ABC transporter ATP-binding subunit GldA from Flavobacterium johnsoniae (NCBI Accession: WP_014164802.1) (Agarwal et al., 1997) and Intraflagellar transport protein IFT52 from Chlamydomonas reinhardtii (NCBI Accession: XP_001692161.1). While these features, together with the presence of a

109 signal peptide, may suggest hp3 function as part of an unknown bacterial ABC transporter, the direct mechanism on how hp3 facilitate Nkp-1 nematicidal activity remains unclear. Interestingly, previous studies indicate that NTPase’s can improve bacterial persistence in the target host by degrading bacterial cyclic nucleotides including pathogen-associated molecular patterns (PAMP), hence leading to undetected bacterial invasion (Woodward et al., 2010; Dey and Bishai, 2014). This may further explain the previous loss of HG8 nematicidal activity due to a transposon mutation in hp3 which resulted in diminishing E. coli HG8 colonisation in C. elegans intestine, hence leading to increased nematode survival compared to the wild type clone (Ballestriero et al., 2010).

Similarly, hp4 (ZP_01132243.1) encoding for a predicted periplasmic 2,3-cyclic nucleotide 2-phosphodiesterase/3-nucleotidase may also play a role in evading the host immune response. A recent characterisation of the homologous enzyme CpdB from Escherichia coli BL21 revealed its novel activity as a cyclic dinucleotide phosphodiesterase showing a high efficiency towards 3’-AMP (Adenosine 3'- monophosphate) and 2’,3’-cyclic mononucleotides (Lopez-Villamizar et al., 2016). Bacterial cyclic dinucleotides are recognised PAMPs that trigger the innate immune response of the associated host including C. elegans (Kim, 2015; Tor et al., 2020), hence targeted degradation of these PAMPs by HP4 (2,3-cyclic nucleotide 2- phosphodiesterase / 3-nucleotidase bifunctional periplasmic) may result in increased bacterial persistence within the worm gut. In support of this idea, past studies have shown that mutations in cpdB ,a homolog of HP4 in Salmonella enterica result in increased colonisation and persistence of the strain in the cecum of chickens (Liu et al., 2015).

In summary, whilst Nkp-1 (expressed by hp1) serves as the nematicidal protein, the intense killing activity of E. coli HG8 against C. elegans may be a function of the other auxiliary genes that are responsible for bacterial persistence, adhesion and biofilm formation (see proposed model in Figure 2.14). Further mutations and genetic complementation of the respective genes are required to confirm this hypothesis.

110

Figure 2.14 Schematic diagram representing the proposed function of key genes encoded on HG8 that are involved in antinematode activity. The Nkp-1 protein is expressed by the hp1 gene during the cell growth. An antitoxin-like factor is encoded by hp2 to prevent the toxicity of Nkp-1 in the cell cytoplasm by forming the Nkp-1-HP2 antitoxin complex. Several lines of findings indicate that the antitoxin factors found in other bacteria for example E. coli, S. aureus, Enterococcus faecium and Acinetobacter baumannii are easily degraded (Amitai et al., 2004; Fernández-García et al., 2016). In light of this, the HP2 antitoxin-like factor is also speculated to be labile and easily degraded by the cell protease. In order to continually neutralise the Nkp-1, the HP2 antitoxin must be constantly expressed. Indirect function of auxiliary genes may also be important to enhance HG8 toxicity. These include hp3 and hp4 (NCBI Accession: ZP_01132244.1 and ZP_01132243.1 respectively) that are suggested to degrade the cyclic nucleotides/PAMPs resulting in prevention of HG8 bacterial invasion in C. elegans. In addition, hp5 expressing BON-domain containing HP5 and OmpA expressing an OmpA-like protein (NCBI Accession: ZP_01132242.1 and ZP_01132239.1 respectively) are believed to be responsible for bacterial adhesion and colonisation of the C. elegans pharynx and intestine.

111

2.4.4 E. coli HP1 and HG8 strains appear to be specifically toxic to nematodes

Despite testing a wide variety of bacterial and fungal isolates I did not find any evidence of antifungal and antibacterial activity by the Nkp-1 expressing clones, indicating that Nkp-1 is unlikely to have a wide spectrum of antagonistic activity. However, given the prediction that Nkp-1 binds to glycoconjugate receptors via a conserved carbohydrate binding domains (CBM6 or CBM35) (Figure 2.13), future studies should include screening for activity towards a wider variety of eukaryotic target organisms. In particular the screening of Nkp-1 biological potential against insects and protozoa should provide more information on the target specificity of this protein, since the glycoconjugate receptors among nematodes, protozoa and insects are highly conserved (Veríssimo et al., 2019).

2.4.5 High temperature diminished E. coli HP1 and HG8 toxic activity against C. elegans

A significant increase in the survival of C. elegans was observed when exposed to heat- killed E. coli HP1 and HG8 strains compared to the viable strains. This observation may suggest that live E. coli HP1 and HG8 cells are required for the toxic activity against the assayed nematodes. Previous studies have shown that live bacterial cells are required for full toxicity. For example, C. elegans survival on the heat-inactivated Burkholderia pseudomallei and an E. coli violacein-producing clone was increased compared to nematodes exposed to their viable counterparts (Gan et al., 2002; Ballestriero et al., 2014). There are also studies showing that nematode killing activity is associated with an increase in colonisation by strains of bacterial pathogens Salmonella and Staphylococcus aureus (Irazoqui et al., 2010; Portal-Celhay et al., 2012). Moreover, the ‘slow killing’ activity of C. elegans by P. aeruginosa is correlated with bacterial colonisation of the nematode gut resulting in diminishing pharyngeal pumping and nematode immobilisation leading to death (Tan et al., 1999; Marko et al., 2018). Likewise, C. elegans colonisation by two Enterococci species (E. faecalis and E. faecium) results in intestinal lumen distension due to the increasing bacterial proliferation, severe

112 damage to microvilli of intestinal cells and impairment of defecation rhythm (Yuen and Ausubel, 2018). In contrast, damages were not observed on nematodes exposed to the heat-killed bacteria (Yuen and Ausubel, 2018). One conclusion that can be drawn from these findings is the additive effect of live toxic bacterial cells in causing damage and eventual death of the C. elegans host.

2.5 CONCLUSION

In conclusion, the nematode killing activity by E. coli HG8 has been successfully characterised. A single gene hp1 is responsible for the nematode killing activity, however its full toxicity in HG8 may also be a result of a combination of strong expression by the T7 promoter of the expression vector and the function of auxiliary genes located on the fosmid insert. These additional HG8 encoded genes may indirectly increase toxicity of Nkp-1 expressing E. coli via improving bacterial adhesion and persistence resulting in increased bacterial loads.

Further analysis of the Nkp-1 protein sequence identified a putative carbohydrate binding module, which is proposed to enable binding of Nkp-1 to the glycoconjugate receptor of C. elegans prior to toxicity. However, the specific nematode glycoconjugate receptor that Nkp-1 binds to is still unknown and this hypothesis warrants further experimental validation. Furthermore, while the Nkp-1 expressing clone was confirmed as toxic, its killing mechanism against C. elegans remains to be elucidated. In addition, the animal behavioural and cellular response in confronting the challenge of Nkp-1 expressing heterologous clones is still undetermined. These questions will be addressed in the following Chapter 3.

113

2.6 REFERENCES

Abbott, D. W., Ficko-Blean, E., van Bueren, A. L., Rogowski, A., Cartmell, A., Coutinho, P. M., Henrissat, B., Gilbert, H. J., & Boraston, A. B. (2009). Analysis of the structural and functional diversity of plant cell wall specific family 6 carbohydrate binding modules. Biochemistry, 48(43), 10395-10404. Agarwal, S., Hunnicutt, D. W., & McBride, M. J. (1997). Cloning and characterization of the Flavobacterium johnsoniae (Cytophaga johnsonae) gliding motility gene, gldA. Proceedings of the National Academy of Sciences of the USA, 94(22), 12139-12144. Ahmad, I., Nawaz, N., Darwesh, N. M., ur Rahman, S., Mustafa, M. Z., Khan, S. B., & Patching, S. G. (2018). Overcoming challenges for amplified expression of recombinant proteins using Escherichia coli. Protein Expression and Purification, 144, 12-18. Altschul, S. F., Madden, T. L., Schäffer, A. A., Zhang, J., Zhang, Z., Miller, W., & Lipman, D. J. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Research, 25(17), 3389-3402. Amitai, S., Yassin, Y., & Engelberg-Kulka, H. (2004). MazF-mediated cell death in Escherichia coli: a point of no return. Journal of Bacteriology, 186(24), 8295- 8300. Bader, M. W., Navarre, W. W., Shiau, W., Nikaido, H., Frye, J. G., McClelland, M., Fang, F. C., & Miller, S. I. (2003). Regulation of Salmonella typhimurium virulence gene expression by cationic antimicrobial peptides. Molecular Microbiology, 50(1), 219-230. Ballesteros, C., Pulaski, C. N., Bourguinat, C., Keller, K., Prichard, R. K., & Geary, T. G. (2018). Clinical validation of molecular markers of macrocyclic lactone resistance in Dirofilaria immitis. International Journal for Parasitology: Drugs and Drug Resistance, 8(3), 596-606. Ballestriero, F., Daim, M., Penesyan, A., Nappi, J., Schleheck, D., Bazzicalupo, P., Di Schiavi, E., & Egan, S. (2014). Antinematode activity of Violacein and the role of the insulin/IGF-1 pathway in controlling violacein sensitivity in Caenorhabditis elegans. PloS One, 9(10), e109201-e109201. Ballestriero, F., Thomas, T., Burke, C., Egan, S., & Kjelleberg, S. (2010). Identification of compounds with bioactivity against the nematode Caenorhabditis elegans by a screen based on the functional genomics of the marine bacterium Pseudoalteromonas tunicata D2. Applied and Environmental Microbiology, 76(17), 5710-5717. Balzer, S., Kucharova, V., Megerle, J., Lale, R., Brautaset, T., & Valla, S. (2013). A comparative analysis of the properties of regulated promoter systems commonly used for recombinant gene expression in Escherichia coli. Microbial Cell Factories, 12, 26. Bartra, S. S., Gong, X., Lorica, C. D., Jain, C., Nair, M. K., Schifferli, D., Qian, L., Li, Z., Plano, G. V., & Schesser, K. (2012). The outer membrane protein A (OmpA) of Yersinia pestis promotes intracellular survival and virulence in mice. Microbial Pathogenesis, 52(1), 41-46.

114

Becker, S. L., Liwanag, H. J., Snyder, J. S., Akogun, O., Belizario, V., Jr., Freeman, M. C., Gyorkos, T. W., Imtiaz, R., Keiser, J., Krolewiecki, A., Levecke, B., Mwandawiro, C., Pullan, R. L., Addiss, D. G., & Utzinger, J. (2018). Toward the 2020 goal of soil-transmitted helminthiasis control and elimination. PLoS Neglected Tropical Diseases, 12(8), e0006606. Bernbom, N., Ng, Y. Y., Kjelleberg, S., Harder, T., & Gram, L. (2011). Marine bacteria from Danish coastal waters show antifouling activity against the marine fouling bacterium Pseudoalteromonas sp. strain S91 and zoospores of the green alga Ulva australis independent of bacteriocidal activity. Applied and Environmental Microbiology, 77(24), 8557-8567. Berninsone, P. M. (2018). Carbohydrates and glycosylation. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_glycosylation/glycosylation.html. Bischof, L. J., Huffman, D. L., & Aroian, R. V. (2006). Assays for toxicity studies in C. elegans with Bt crystal proteins. In C. elegans (pp. 139-154): Springer. Bolam, D. N., Xie, H., Pell, G., Hogg, D., Galbraith, G., Henrissat, B., & Gilbert, H. J. (2004). X4 modules represent a new family of carbohydrate-binding modules that display novel properties. Journal of Biological Chemistry, 279(22), 22953-22963. Bravo, A., Gómez, I., Porta, H., García-Gómez, B. I., Rodriguez-Almazan, C., Pardo, L., & Soberón, M. (2013). Evolution of Bacillus thuringiensis Cry toxins insecticidal activity. Microbial Biotechnology, 6(1), 17-26. Bravo, A., Likitvivatanavong, S., Gill, S. S., & Soberón, M. (2011). Bacillus thuringiensis: a story of a successful bioinsecticide. Insect Biochemistry and Molecular Biology, 41(7), 423-431. Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics, 77(1), 71-94. Büchter, C., Koch, K., Freyer, M., Baier, S., Saier, C., Honnen, S., & Wätjen, W. (2020). The beauvericin impairs development, fertility and life span in the nematode Caenorhabditis elegans accompanied by increased germ cell apoptosis and lipofuscin accumulation. Toxicology Letters, 334, 102-109. Burke, C., Thomas, T., Egan, S., & Kjelleberg, S. (2007). The use of functional genomics for the identification of a gene cluster encoding for the biosynthesis of an antifungal tambjamine in the marine bacterium Pseudoalteromonas tunicata. Environmental Microbiology, 9(3), 814-818. Cantarel, B. L., Coutinho, P. M., Rancurel, C., Bernard, T., Lombard, V., & Henrissat, B. (2008). The Carbohydrate-Active EnZymes database (CAZy): an expert resource for Glycogenomics. Nucleic Acids Research, 37, D233-D238. Charuchaibovorn, S., Sanprasert, V., Sutthanont, N., Hu, Y., Abraham, A., Ostroff, G. R., Aroian, R. V., Jaleta, T. G., Lok, J. B., & Nuchprayoon, S. (2019). Bacillus thuringiensis Cry5B is active against Strongyloides stercoralis in vitro. The American Journal of Tropical Medicine and Hygiene, 101(5), 1177-1182. Chen, I. M. A., Markowitz, V. M., Chu, K., Palaniappan, K., Szeto, E., Pillay, M., Ratner, A., Huang, J., Andersen, E., Huntemann, M., Varghese, N., Hadjithomas, M., Tennessen, K., Nielsen, T., Ivanova, N. N., & Kyrpides, N. C. (2017). IMG/M: integrated genome and metagenome comparative data analysis system. Nucleic Acids Research, 45, D507-D516. Cleveland, D. W., Fischer, S. G., Kirschner, M. W., & Laemmli, U. K. (1977). Peptide mapping by limited proteolysis in sodium dodecyl sulfate and analysis by gel electrophoresis. Journal of Biological Chemistry, 252(3), 1102-1106.

115

Correia, M. A. S., Abbott, D. W., Gloster, T. M., Fernandes, V. O., Prates, J. A. M., Montanier, C., Dumon, C., Williamson, M. P., Tunnicliffe, R. B., Liu, Z., Flint, J. E., Davies, G. J., Henrissat, B., Coutinho, P. M., Fontes, C. M. G. A., & Gilbert, H. J. (2010). Signature active site architectures illuminate the molecular basis for ligand specificity in Family 35 carbohydrate binding module. Biochemistry, 49(29), 6193-6205. Coyne, D. L., Cortada, L., Dalzell, J. J., Claudius-Cole, A. O., Haukeland, S., Luambano, N., & Talwana, H. (2018). Plant-parasitic nematodes and food security in Sub- Saharan Africa. Annual Review of Phytopathology, 56(1), 381-403. Daim, M. F. (2012) Characterising the antinematodal activity of the marine bacterium Pseudoalteromonas tunicata when screened against Caenorhabditis elegans. Honours Thesis. The University of New South Wales, Australia Dalcin Martins, P., Danczak, R. E., Roux, S., Frank, J., Borton, M. A., Wolfe, R. A., Burris, M. N., & Wilkins, M. J. (2018). Viral and metabolic controls on high rates of microbial sulfur and carbon cycling in wetland ecosystems. Microbiome, 6(1), 138. de Maagd, R. A., Weemen-Hendriks, M., Stiekema, W., & Bosch, D. (2000). Bacillus thuringiensis delta-endotoxin Cry1C domain III can function as a specificity determinant for Spodoptera exigua in different, but not all, Cry1-Cry1C hybrids. Applied and Environmental Microbiology, 66(4), 1559. Deist, B. R., Rausch, M. A., Fernandez-Luna, M. T., Adang, M. J., & Bonning, B. C. (2014). Bt toxin modification for enhanced efficacy. Toxins, 6(10), 3005-3027. Dey, B., & Bishai, W. R. (2014). Crosstalk between Mycobacterium tuberculosis and the host cell. Seminars in Immunology, 26(6), 486-496. Di Tommaso, P., Moretti, S., Xenarios, I., Orobitg, M., Montanyola, A., Chang, J.-M., Taly, J.-F., & Notredame, C. (2011). T-Coffee: a web server for the multiple sequence alignment of protein and RNA sequences using structural information and homology extension. Nucleic Acids Research, 39, W13-W17. Dong, T., & Schellhorn, H. E. (2009). Global effect of RpoS on gene expression in pathogenic Escherichia coli O157: H7 strain EDL933. BMC Genomics, 10(1), 349. Dumon-Seignovert, L., Cariot, G., & Vuillard, L. (2004). The toxicity of recombinant proteins in Escherichia coli: a comparison of overexpression in BL21(DE3), C41(DE3), and C43(DE3). Protein Expression and Purification, 37(1), 203-206. Egan, S., James, S., Holmström, C., & Kjelleberg, S. (2002). Correlation between pigmentation and antifouling compounds produced by Pseudoalteromonas tunicata. Environmental Microbiology, 4(8), 433-442. Feng, D., Chen, Z., Wang, Z., Zhang, C., He, K., & Guo, S. (2015). Domain III of Bacillus thuringiensis Cry1Ie toxin plays an important role in binding to peritrophic membrane of Asian Corn Borer. PloS One, 10(8), e0136430. Fernández-García, L., Blasco, L., Lopez, M., Bou, G., García-Contreras, R., Wood, T., & Tomas, M. (2016). Toxin-antitoxin systems in clinical pathogens. Toxins, 8(7), 227. Ficko-Blean, E., & Boraston, A. B. (2012). Insights into the recognition of the human glycome by microbial carbohydrate-binding modules. Current Opinion in Structural Biology, 22(5), 570-577. Fico, S., & Mahillon, J. (2006). TasA-tasB, a new putative toxin-antitoxin (TA) system from Bacillus thuringiensis pGI1 plasmid is a widely distributed composite mazE- doc TA system. BMC Genomics, 7(1), 259.

116

Finn, R. D., Coggill, P., Eberhardt, R. Y., Eddy, S. R., Mistry, J., Mitchell, A. L., Potter, S. C., Punta, M., Qureshi, M., Sangrador-Vegas, A., Salazar, G. A., Tate, J., & Bateman, A. (2016). The Pfam protein families database: towards a more sustainable future. Nucleic Acids Research, 44(D1), D279-D285. Franks, A., Egan, S., Holmström, C., James, S., Lappin-Scott, H., & Kjelleberg, S. (2006). Inhibition of fungal colonization by Pseudoalteromonas tunicata provides a competitive advantage during surface colonization. Applied and Environmental Microbiology, 72(9), 6079-6087. Fujimoto, Z., Suzuki, N., Kishine, N., Ichinose, H., Momma, M., Kimura, A., & Funane, K. (2017). Carbohydrate-binding architecture of the multi-modular α-1, 6- glucosyltransferase from Paenibacillus sp. 598K, which produces α-1, 6- glucosyl-α-glucosaccharides from starch. Biochemical Journal, 474(16):2763- 2778. Gan, Y. H., Chua, K. L., Chua, H. H., Liu, B., Hii, C. S., Chong, H. L., & Tan, P. (2002). Characterization of Burkholderia pseudomallei infection and identification of novel virulence factors using a Caenorhabditis elegans host system. Molecular Microbiology, 44(5), 1185-1197. Geng, C., Liu, Y., Li, M., Tang, Z., Muhammad, S., Zheng, J., Wan, D., Peng, D., Ruan, L., & Sun, M. (2017). Dissimilar crystal proteins Cry5Ca1 and Cry5Da1 synergistically act against Meloidogyne incognita and delay Cry5Ba-based nematode resistance. Applied and Environmental Microbiology, 83(18), e03505- 03516. Geurden, T., Chartier, C., Fanke, J., di Regalbono, A. F., Traversa, D., von Samson- Himmelstjerna, G., Demeler, J., Vanimisetti, H. B., Bartram, D. J., & Denwood, M. J. (2015). Anthelmintic resistance to ivermectin and moxidectin in gastrointestinal nematodes of cattle in Europe. International Journal for Parasitology: Drugs and Drug Resistance, 5(3), 163-171. Giacalone, M. J., Gentile, A. M., Lovitt, B. T., Berkley, N. L., Gunderson, C. W., & Surber, M. W. (2006). Toxic protein expression in Escherichia coli using a rhamnose-based tightly regulated and tunable promoter system. Biotechniques, 40(3), 355-364. Guzman, L.-M., Belin, D., Carson, M. J., & Beckwith, J. (1995). Tight regulation, modulation, and high-level expression by vectors containing the arabinose PBAD promoter. Journal of Bacteriology, 177(14), 4121-4130. Harrington, D. (2005). Linear Rank Tests in Survival Analysis. In Encyclopedia of Biostatistics. John Wiley & Sons, Ltd. 10.1002/0470011815.b2a11047 Heustis, R. J., Ng, H. K., Brand, K. J., Rogers, M. C., Le, L. T., Specht, C. A., & Fuhrman, J. A. (2012). Pharyngeal polysaccharide deacetylases affect development in the nematode C. elegans and deacetylate chitin in vitro. PloS One, 7(7), e40426. Holden-Dye, L., & Walker, R. (2014). Anthelmintic drugs and nematocides: studies in Caenorhabditis elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_anthelminticdrugs.2/anthelminticdrug s.2.html Holmstrom, C., & Kjelleberg, S. (1999). Marine Pseudoalteromonas species are associated with higher organisms and produce biologically active extracellular agents. FEMS Microbiology Ecology, 30(4), 285-293. Huang, X., Li, D., Xi, L., & Mylonakis, E. (2014). Caenorhabditis elegans: a simple nematode infection model for Penicillium marneffei. PloS one, 9(9), e108764.

117

Illergård, K., Ardell, D. H., & Elofsson, A. (2009). Structure is three to ten times more conserved than sequence—a study of structural response in protein cores. Proteins: Structure, Function, and Bioinformatics, 77(3), 499-508. Irazoqui, J. E., Troemel, E. R., Feinbaum, R. L., Luhachack, L. G., Cezairliyan, B. O., & Ausubel, F. M. (2010). Distinct pathogenesis and host responses during infection of C. elegans by P. aeruginosa and S. aureus. PLoS Pathogens, 6(7), e1000982. James, S. G., Holmström, C., & Kjelleberg, S. (1996). Purification and characterization of a novel antibacterial protein from the marine bacterium D2. Applied and Environmental Microbiology, 62(8), 2783-2788. Jankowska, E., Parsons, L. M., Song, X., Smith, D. F., Cummings, R. D., & Cipollo, J. F. (2018). A comprehensive Caenorhabditis elegans N-glycan shotgun array. Glycobiology, 28(4), 223-232. Jing, X., Yuan, Y., Wu, Y., Wu, D., Gong, P., & Gao, M. (2019). Crystal structure of Bacillus thuringiensis Cry7Ca1 toxin active against Locusta migratoria manilensis. Protein Science, 28(3), 609-619. Karlova, R., Weemen-Hendriks, M., Naimov, S., Ceron, J., Dukiandjiev, S., & de Maagd, R. A. (2005). Bacillus thuringiensis δ-endotoxin domain III enhances activity against Heliothis virescens in some, but not all Cry1-Cry1Ac hybrids. Journal of Invertebrate Pathology, 88(2), 169-172. Katsimpouras, C., Dimarogona, M., , P., Christakopoulos, P., & Topakas, E. (2016). A thermostable GH26 endo-β-mannanase from Myceliophthora thermophila capable of enhancing lignocellulose degradation. Applied Microbiology and Biotechnology, 100(19), 8385-8397. Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N., & Sternberg, M. J. E. (2015). The Phyre2 web portal for protein modeling, prediction and analysis. Nature Protocols, 10, 845. Kim, D. H. & Ewbank, J. (2015). Signaling in the innate immune response. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at https://www.ncbi.nlm.nih.gov/books/NBK19673/. Kim, S. W., Oh, M. H., Jun, S. H., Jeon, H., Kim, S. I., Kim, K., Lee, Y. C., & Lee, J. C. (2016). Outer membrane Protein A plays a role in pathogenesis of Acinetobacter nosocomialis. Virulence, 7(4), 413-426. Kopel, M., Helbert, W., Henrissat, B., Doniger, T., & Banin, E. (2014). Draft genome sequence of Pseudoalteromonas sp. strain PLSV, an ulvan-degrading bacterium. Genome Announcements, 2(6), e01257-01214. Laemmli, U. K., & Favre, M. (1973). Maturation of the head of bacteriophage T4: I. DNA packaging events. Journal of Molecular Biology, 80(4), 575-599. Lee, K.-Y., & Lee, B.-J. (2016). Structure, biology, and therapeutic application of toxin- antitoxin systems in pathogenic bacteria. Toxins, 8(10), 305. Leipe, D. D., Wolf, Y. I., Koonin, E. V., & Aravind, L. (2002). Classification and evolution of P-loop GTPases and related ATPases. Journal of Molecular Biology, 317(1), 41-72. Letunic, I., Doerks, T., & Bork, P. (2015). SMART: recent updates, new developments and status in 2015. Nucleic Acids Research, 43(D1), D257-D260. Li, J., Carroll, J., & Ellar, D. J. (1991). Crystal structure of insecticidal δ-endotoxin from Bacillus thuringiensis at 2.5 Å resolution. Nature, 353(6347), 815-821.

118

Liu, H., Chen, L., Wang, X., Si, W., Wang, H., Wang, C., Liu, S., & Li, G. (2015). Decrease of colonization in the chicks' cecum and internal organs of Salmonella enterica serovar Pullorum by deletion of cpdB by red system. Microbial Pathogenesis, 80, 21-26. Liu, Y., Zhou, Z., Wang, Z., Zhong, B., Shu, C., & Zhang, J. (2020). Replacement of loop2 and 3 of Cry1Ai in domain II affects specificity to the economically important insect Bombyx mori. Journal of Invertebrate Pathology, 169, 107296. Lopez-Villamizar, I., Cabezas, A., Pinto, R. M., Canales, J., Ribeiro, J. M., Cameselle, J. C., & Costas, M. J. (2016). The Characterization of Escherichia coli CpdB as a recombinant protein reveals that, besides having the expected 3 -nucleotidase and 2 ,3 -cyclic mononucleotide phosphodiesterase activities, it is also active as cyclic dinucleotide phosphodiesterase. PloS One, 11(6), e0157308. López Pazos, S. (2007). Three-dimensional structure of Bacillus thuringiensis toxins: a review. Acta Biológica Colombiana, 12(2), 19-32. Maglione, G., Matsushita, O., Russell, J. B., & Wilson, D. B. (1992). Properties of a genetically reconstructed Prevotella ruminicola endoglucanase. Applied and Environmental Microbiology, 58(11), 3593. Mahmud, M. A., Spigt, M., Bezabih, A. M., Pavon, I. L., Dinant, G.-J., & Velasco, R. B. (2015). Efficacy of handwashing with soap and nail clipping on intestinal parasitic infections in school-aged children: A factorial cluster randomized controlled trial. PLoS Medicine, 12(6), e1001837. Mantel, N. (1966). Evaluation of survival data and two new rank order statistics arising in its consideration. Cancer Chemotherapy Reports, 50(3), 163-170. Marchler-Bauer, A., Bo, Y., Han, L., He, J., Lanczycki, C. J., Lu, S., Chitsaz, F., Derbyshire, M. K., Geer, R. C., Gonzales, N. R., Gwadz, M., Hurwitz, D. I., Lu, F., Marchler, G. H., Song, J. S., Thanki, N., Wang, Z., Yamashita, R. A., Zhang, D., Zheng, C., Geer, L. Y., & Bryant, S. H. (2017). CDD/SPARCLE: functional classification of proteins via subfamily domain architectures. Nucleic Acids Research, 45, D200-D203. Marko, V. A., Kilmury, S. L. N., MacNeil, L. T., & Burrows, L. L. (2018). Pseudomonas aeruginosa type IV minor pilins and PilY1 regulate virulence by modulating FimS-AlgR activity. PLoS Pathogens, 14(5), e1007074. Martinez, E., Cantet, F., Fava, L., Norville, I., & Bonazzi, M. (2014). Identification of OmpA, a Coxiella burnetii protein involved in host cell invasion, by multi- phenotypic high-content screening. PLoS Pathogens, 10(3), e1004013. Mendez, J., Cascales, D., Garcia-Torrico, A. I., & Guijarro, J. A. (2018). Temperature- dependent gene expression in Yersinia ruckeri: tracking specific genes by bioluminescence during in vivo colonization. Frontiers in Microbiology, 9, 1098. Mesnage, S., Dellarole, M., Baxter, N. J., Rouget, J.-B., Dimitrov, J. D., Wang, N., Fujimoto, Y., Hounslow, A. M., Lacroix-Desmazes, S., Fukase, K., Foster, S. J., & Williamson, M. P. (2014). Molecular basis for bacterial peptidoglycan recognition by LysM domains. Nature Communications, 5(1), 4269. Michel, G., Barbeyron, T., Kloareg, B., & Czjzek, M. (2009). The family 6 carbohydrate- binding modules have coevolved with their appended catalytic modules toward similar substrate specificity. Glycobiology, 19(6), 615-623.

119

Montanier, C., van Bueren, A. L., Dumon, C., Flint, J. E., Correia, M. A., Prates, J. A., Firbank, S. J., Lewis, R. J., Grondin, G. G., Ghinet, M. G., Gloster, T. M., Herve, C., Knox, J. P., Talbot, B. G., Turkenburg, J. P., Kerovuo, J., Brzezinski, R., Fontes, C. M., Davies, G. J., Boraston, A. B., & Gilbert, H. J. (2009). Evidence that family 35 carbohydrate binding modules display conserved specificity but divergent function. Proceedings of the National Academy of Sciences of the USA, 106(9), 3065-3070. Moustafa, I., Connaris, H., Taylor, M., Zaitsev, V., Wilson, J. C., Kiefel, M. J., Von Itzstein, M., & Taylor, G. (2004). Sialic acid recognition by Vibrio cholerae neuraminidase. Journal of Biological Chemistry, 279(39), 40819-40826. Naimov, S., Weemen-Hendriks, M., Dukiandjiev, S., & de Maagd, R. A. (2001). Bacillus thuringiensis delta-endotoxin Cry1 hybrid proteins with increased activity against the Colorado Potato . Applied and Environmental Microbiology, 67(11), 5328. Nappi, J. (2019). Discovery of novel bioactive metabolites from marine epiphytic bacteria and assessment of their ecological role. Unpublished Doctoral dissertation, University of New South Wales, Australia. Narayanan, A., Ridilla, M., & Yernool, D. A. (2011). Restrained expression, a method to overproduce toxic membrane proteins by exploiting operator–repressor interactions. Protein Science, 20(1), 51-61. Nadzirin, N., & Firdaus-Raih, M. (2012). Proteins of unknown function in the Protein Data Bank (PDB): an inventory of true uncharacterized proteins and computational tools for their analysis. International Journal of Molecular Sciences, 13(10), 12761-12772. Nicol, J., Turner, S., Coyne, D., Den Nijs, L., Hockland, S., & Maafi, Z. T. (2011). Current nematode threats to world agriculture. In Genomics and Molecular Genetics of Plant-Nematode Interactions (pp. 21-43): Springer. O Neal, C., Jobling, M., Holmes, R., & Hol, W. (2005). Structural biology: Structural basis for the activation of cholera toxin by human ARF6-GTP. Science, 5737:1093-1096 Paulsen, S. S., Strube, M. L., Bech, P. K., Gram, L., & Sonnenschein, E. C. (2019). Marine chitinolytic Pseudoalteromonas represents an untapped reservoir of bioactive potential. mSystems, 4(4), e00060-00019. Penesyan, A., Ballestriero, F., Daim, M., Kjelleberg, S., Thomas, T., & Egan, S. (2013). Assessing the effectiveness of functional genetic screens for the identification of bioactive metabolites. Marine Drugs, 11(1), 40-49. Penesyan, A., Kjelleberg, S., & Egan, S. (2010). Development of novel drugs from marine surface associated microorganisms. Marine Drugs, 8(3), 438-459. Pigott, C. R., & Ellar, D. J. (2007). Role of receptors in Bacillus thuringiensis crystal toxin activity. Microbiology and Molecular Biology Reviews, 71(2), 255-281. Plummer, M., de Martel, C., Vignat, J., Ferlay, J., Bray, F., & Franceschi, S. (2016). Global burden of cancers attributable to infections in 2012: A synthetic analysis. The Lancet Global Health, 4(9), e609-e616. Porcella, S., Raffel, S., Schrumpf, M., Montgomery, B., Smith, T., & Schwan, T. (2008). The genome sequence of Borrelia hermsii and Borrelia turicatae: Comparative analysis of two agents of endemic N. America relapsing fever. Submitted (DEC- 2004) to the EMBL/GenBank/DDBJ databases. Portal-Celhay, C., Bradley, E. R., & Blaser, M. J. (2012). Control of intestinal bacterial proliferation in regulation of lifespan in Caenorhabditis elegans. BMC Microbiology, 12(49), 1-17. 120

Probst, A. J., Elling, F. J., Castelle, C. J., Zhu, Q., Elvert, M., Birarda, G., Holman, H.-Y. N., Lane, K. R., Ladd, B., Ryan, M. C., Woyke, T., Hinrichs, K.-U., & Banfield, J. F. (2020). Lipid analysis of CO2-rich subsurface aquifers suggests an autotrophy-based deep biosphere with lysolipids enriched in CPR bacteria. The ISME Journal, 14, 1547–1560. Puskarova, A., Buckova, M., Krakova, L., Pangallo, D., & Kozics, K. (2017). The antibacterial and antifungal activity of six essential oils and their cyto/genotoxicity to human HEL 12469 cells. Scientific Reports 7(1), 8211. Qi, W., Colarusso, A., Olombrada, M., Parrilli, E., Patrignani, A., Tutino, M. L., & Toll- Riera, M. (2019). New insights on Pseudoalteromonas haloplanktis TAC125 genome organization and benchmarks of genome assembly applications using next and third generation sequencing technologies. Scientific Reports, 9(1), 1-14. Rahul, S., Chandrashekhar, P., Hemant, B., Chandrakant, N., Laxmikant, S., & Satish, P. (2014). Nematicidal activity of microbial pigment from Serratia marcescens. Natural Product Research, 28(17), 1399-1404. Rodríguez-González, Á., Porteous-Álvarez, A. J., Val, M. D., Casquero, P. A., & Escriche, B. (2020). Toxicity of five Cry proteins against the insect pest Acanthoscelides obtectus (Coleoptera: Chrisomelidae: Bruchinae). Journal of Invertebrate Pathology, 169, 107295. Romano, G., Costantini, M., Sansone, C., Lauritano, C., Ruocco, N., & Ianora, A. (2017). Marine microorganisms as a promising and sustainable source of bioactive molecules. Marine Environmental Research, 128, 58-69. Sainz-Polo, M. A., Valenzuela, S. V., González, B., Pastor, F. I. J., & Sanz-Aparicio, J. (2014). Structural analysis of glucuronoxylan-specific Xyn30D and its attached CBM35 domain gives insights into the role of modularity in specificity. Journal of Biological Chemistry, 289(45), 31088-31101. Sant'anna, V., Vommaro, R. C., & de Souza, W. (2013). Caenorhabditis elegans as a model for the screening of anthelminthic compounds: ultrastructural study of the effects of albendazole. Experimental Parasitology, 135(1), 1-8. Schultz, J., Milpetz, F., Bork, P., & Ponting, C. P. (1998). SMART, a simple modular architecture research tool: Identification of signaling domains. Proceedings of the National Academy of Sciences of the USA, 95(11), 5857-5864. Seidman, C. E., Struhl, K., Sheen, J., & Jessen, T. (1997). Introduction of plasmid DNA into cells. Current Protocols in Molecular Biology, 37(1), 1.8. 1-1.8. 10. Shalaby, H. A. (2013). Anthelmintics Resistance; How to overcome it? Iranian Journal of Parasitology, 8(1), 18-32. Short, F. L., Monson, R. E., & Salmond, G. P. C. (2015). A Type III protein-RNA toxin- antitoxin system from Bacillus thuringiensis promotes plasmid retention during spore development. RNA Biology, 12(9), 933-937. Skerniškytė, J., Karazijaitė, E., Deschamps, J., Krasauskas, R., Briandet, R., & Sužiedėlienė, E. (2019). The mutation of conservative Asp268 residue in the peptidoglycan-associated domain of the OmpA protein affects multiple Acinetobacter baumannii virulence characteristics. Molecules, 24(10), 1972. Souza, C., Simpson-Louredo, L., Mota, H. F., Faria, Y. V., Cabral, F. O., Colodette, S. D. S., Canellas, M., Cucinelli, A., Luna, M. D. G., Santos, C. D. S., Moreira, L. O., & Mattos-Guaraldi, A. L. (2019). Virulence potential of Corynebacterium striatum towards Caenorhabditis elegans. Antonie van Leeuwenhoek, 112(9), 1331-1340. Stephenson, L. S., Latham, M. C., & Ottesen, E. (2000). Malnutrition and parasitic helminth infections. Parasitology, 121(S1), S23-S38. 121

Stiernagle, T. (2006). Maintenance of C. elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_strainmaintain/strainmaintain.html. Suzuki, N., Fujimoto, Z., Kim, Y. M., Momma, M., Kishine, N., Suzuki, R., Suzuki, S., Kitamura, S., Kobayashi, M., Kimura, A., & Funane, K. (2014). Structural elucidation of the cyclization mechanism of alpha-1,6-glucan by Bacillus circulans T-3040 cycloisomaltooligosaccharide glucanotransferase. Journal of Biological Chemistry, 289(17), 12040-12051. Tan, M.-W., Mahajan-Miklos, S., & Ausubel, F. M. (1999). Killing of Caenorhabditis elegans by Pseudomonas aeruginosa used to model mammalian bacterial pathogenesis. Proceedings of the National Academy of Sciences of the USA, 96(2), 715. Taschner, M., Kotsis, F., Braeuer, P., Kuehn, E. W., & Lorentzen, E. (2014). Crystal structures of IFT70/52 and IFT52/46 provide insight into intraflagellar transport B core complex assembly. Journal of Cell Biology, 207(2), 269-282. Teh, A.-H., Fazli, N. H., & Furusawa, G. (2019). Crystal structure of a neoagarobiose- producing GH16 family β-agarase from Persicobacter sp. CCB-QB2. Applied Microbiology and Biotechnology, 1-9. Tielen, P., Wibberg, D., Blom, J., Rosin, N., Meyer, A.-K., Bunk, B., Schobert, M., Tüpker, R., Schatschneider, S., Rückert, C., Albersmeier, A., Goesmann, A., Vorhölter, F.-J., Jahn, D., & Pühler, A. (2014). Genome sequence of the small- colony variant Pseudomonas aeruginosa MH27, isolated from a chronic urethral catheter infection. Genome Announcements, 2(1), e01174-01113. Tor, Y., Li, Y., Ludford, P. T., Fin, A., & Rovira, A. R. (2020). Fluorescent cyclic dinucleotides. Chemistry–A European Journal, 26, 6076-6084 UniProt Consortium, T. (2018). UniProt: the universal protein knowledgebase. Nucleic Acids Research, 46(5), 2699-2699. van Bueren, A. L., Morland, C., Gilbert, H. J., & Boraston, A. B. (2005). Family 6 carbohydrate binding modules recognize the non-reducing end of beta-1,3-linked glucans by presenting a unique ligand binding surface. Journal of Biological Chemistry, 280(1), 530-537. Veríssimo, C. D. M., Graeff-Teixeira, C., Jones, M. K., & Morassutti, A. L. (2019). Glycans in the roles of parasitological diagnosis and host–parasite interplay. Parasitology, 146(10), 1217-1232. Waterborg, J. H. (2009). The Lowry method for protein quantitation. In The protein protocols handbook (pp. 7-10): Springer. Waterhouse, A., Bertoni, M., Bienert, S., Studer, G., Tauriello, G., Gumienny, R., Heer, F. T., de Beer, T. A. P., Rempfer, C., Bordoli, L., Lepore, R., & Schwede, T. (2018). SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Research, 46(W1), W296-W303. Wei, J.-Z., Siehl, D. L., Hou, Z., Rosen, B., Oral, J., Taylor, C. G., & Wu, G. (2017). An enterotoxin-like binary protein from Pseudomonas protegens with potent nematicidal activity. Applied and Environmental Microbiology, 83(19), 00942- 00917. WHO (World Health Organization) (2016) Soil-transmitted helminth infections. Available at http://www.who.int/mediacentre/factsheets/fs366/en/. Accessed on 1 December 2016.

122

Wolstenholme, A. J., Fairweather, I., Prichard, R., von Samson-Himmelstjerna, G., & Sangster, N. C. (2004). Drug resistance in veterinary helminths. Trends in Parasitology, 20(10), 469-476. Woodward, J. J., Iavarone, A. T., & Portnoy, D. A. (2010). c-di-AMP secreted by intracellular Listeria monocytogenes activates a host type I interferon response. Science, 328(5986), 1703-1705. Xu, C., Wang, B.-C., Yu, Z., & Sun, M. (2014). Structural insights into Bacillus thuringiensis Cry, Cyt and parasporin toxins. Toxins, 6(9), 2732-2770. Yamaguchi, Y., & Inouye, M. (2009). mRNA interferases, sequence‐specific endoribonucleases from the toxin–antitoxin systems. Progress in Molecular Biology and Translational Science, 85, 467-500. Yamaguchi, Y., Park, J. H., & Inouye, M. (2011). Toxin-antitoxin systems in bacteria and archaea. Annual Review of Genetics, 45, 61-79. Yan, Z., Hussain, S., Wang, X., Bernstein, H. D., & Bardwell, J. C. (2019). Chaperone OsmY facilitates the biogenesis of a major family of autotransporters. Molecular Microbiology, 112(5), 1373-1387. Yano, S., Suyotha, W., Oguro, N., Matsui, T., Shiga, S., Itoh, T., Hibi, T., Tanaka, Y., Wakayama, M., & Makabe, K. (2019). Crystal structure of the catalytic unit of GH 87-type α-1,3-glucanase Agl-KA from Bacillus circulans. Scientific Reports, 9(1), 15295. Ye, Y., Gao, J., Jiao, R., Li, H., Wu, Q., Zhang, J., & Zhong, X. (2015). The membrane proteins involved in virulence of Cronobacter sakazakii virulent G362 and attenuated L3101 isolates. Frontiers in Microbiology, 6, 1238. Yeats, C., & Bateman, A. (2003). The BON domain: a putative membrane-binding domain. Trends in Biochemical Sciences, 28(7), 352-355. Yuen, G. J., & Ausubel, F. M. (2018). Both live and dead Enterococci activate Caenorhabditis elegans host defense via immune and stress pathways. Virulence, 9(1), 683-699. Zhang, X., Rogowski, A., Zhao, L., Hahn, M. G., Avci, U., Knox, J. P., & Gilbert, H. J. (2014). Understanding how the complex molecular architecture of mannan- degrading hydrolases contributes to plant cell wall degradation. J Biological Chemistry, 289(4), 2002-2012. Zheng, X., Ji, Y., Weng, X., & Huang, X. (2015). RpoS-dependent expression of OsmY in Salmonella enterica serovar Typhi: activation under stationary phase and SPI- 2-inducing conditions. Current Microbiology, 70(6), 877-882. Ziegelbauer, K., Speich, B., Mäusezahl, D., Bos, R., Keiser, J., & Utzinger, J. (2012). Effect of sanitation on soil-transmitted helminth infection: systematic review and meta-analysis. PLoS Medicine, 9(1), e1001162.

123

SUPPLEMENTARY MATERIALS

Table S2.1 C. elegans survival on 7C8 mutants complemented with P. tunicata D2 uncharacerised hypothetical proteins (hp) wild type (wt)

Overall test; Log-rank (Mantel-Cox) C. elegans survival on 7C8 df = Chi square = p < 0.0001 complemented mutants 6 70.55 One-way ANOVA C. elegans survival on 7C8 df = complemented mutants F = 155.2 p < 0.0001 6 Tukey's test Test strain Day q df p-value

7C8::hp1 vs. 7C8::hp2 2 20.85 14 0.0001 7C8::hp1 vs. 7C8::hp1/hp2 2 18.50 14 0.0001 7C8::hp1 vs. 7C8 2 23.35 14 0.0001 7C8::hp1 vs. 7C8::pBAD24 2 24.74 14 0.0001 7C8::hp1 vs. HG8 2 4.75 14 0.0553 7C8::hp1 vs. OP50 2 25.32 14 0.0001 7C8::hp2 vs. 7C8::hp1/hp2 2 2.34 14 0.6521 7C8::hp2 vs. 7C8 2 2.51 14 0.5838 7C8::hp2 vs. 7C8::pBAD24 2 3.89 14 0.1551 7C8::hp2 vs. HG8 2 25.60 14 0.0001 7C8::hp2 vs. OP50 2 4.47 14 0.0779 7C8::hp1/hp2 vs. 7C8 2 4.85 14 0.0486 7C8::hp1/hp2 vs. 7C8::pBAD24 2 6.24 14 0.0082 7C8::hp1/hp2 vs. HG8 2 23.25 14 0.0001 7C8::hp1/hp2 vs. OP50 2 6.82 14 0.0039 7C8 vs. 7C8::pBAD24 2 1.39 14 0.9505 7C8 vs. HG8 2 28.10 14 0.0001 7C8 vs. OP50 2 1.96 14 0.7989 7C8::pBAD24 vs. HG8 2 29.49 14 0.0001 7C8::pBAD24 vs. OP50 2 0.58 14 0.9995 HG8 vs. OP50 2 30.07 14 0.0001

124

Table S2.2 C. elegans survival on E. coli clones expressing individual P. tunicata D2 hypothetical proteins

Overall test; Log-rank (Mantel-Cox) C. elegans survival on P. tunicata Chi square = D2 individual hypothetical protein df = 5 p < 0.0001 E. coli clones 78.57 One-way ANOVA C. elegans survival on P. tunicata D2 individual hypothetical protein df = 5 F = 46.62 p < 0.0001 E. coli clones Tukey's test Test strain Day q df p-value

HP1 vs. HP2 2 9.998 12 0.0001 HP1 vs. HP1/HP2 2 8.546 12 0.0006 HP1 vs. BD24 2 10.77 12 < 0.0001 HP1 vs. EPI300 2 10.42 12 < 0.0001 HP1 vs. HG8 2 5.527 12 0.0198 HP2 vs. HP1/HP2 2 1.452 12 0.9000 HP2 vs. BD24 2 0.7713 12 0.9929 HP2 vs. EPI300 2 0.4206 12 0.9996 HP2 vs. HG8 2 15.53 12 < 0.0001 HP1/HP2 vs. BD24 2 2.224 12 0.6294 HP1/HP2 vs. EPI300 2 1.873 12 0.7673 HP1/HP2 vs. HG8 2 14.07 12 < 0.0001 BD24 vs. EPI300 2 0.3506 12 0.9998 BD24 vs. HG8 2 16.30 12 < 0.0001 EPI300 vs. HG8 2 15.95 12 < 0.0001

125

Table S2.3 C. elegans eggs hatching competency and brood size on E. coli clones expressing individual P. tunicata D2 hypothetical proteins

One-way ANOVA C. elegans eggs hatching assay df = 2 F = 0.4181 p < 0.0001

C. elegans brood size assay df = 2 F = 169.7 p = 0.6761 Tukey's test C. elegans eggs hatching assay Test strain df q p-value

HP1 vs. HG8 6 1.02 0.7619 HP1 vs. BD24 6 1.20 0.6891 HG8 vs. BD24 6 0.18 0.9908

C. elegans brood size assay Test strain df q p-value

HP1 vs. HG8 6 0.43 0.9508 HP1 vs. BD24 6 22.34 0.0001 HG8 vs. BD24 6 22.77 0.0001

126

Table S2.4 C. elegans survival on live and heat-killed E. coli clones expressing individual P. tunicata D2 hypothetical proteins

Overall test; Log-rank (Mantel-Cox) C. elegans survival on live and heat- killed E. coli clones expressing P. Chi square = df = 5 p < 0.0001 tunicata D2 individual hypothetical 93.57 protein One-way ANOVA C. elegans survival on live and heat- killed E. coli clones expressing P. F = df = 5 p < 0.0001 tunicata D2 individual hypothetical 81.95 protein Tukey's test Test strain Day q df p-value

Heat-killed HP1 vs. heat-killed HG8 2 0.00 12 0.9999 Heat-killed HP1 vs. heat-killed BD24 2 0.55 12 0.9985 Heat-killed HP1 vs. live HP1 2 10.32 12 0.0001 Heat-killed HP1 vs. live HG8 2 21.97 12 0.0001 Heat-killed HP1 vs. live BD24 2 0.00 12 0.9999 Heat-killed HG8 vs. heat-killed BD24 2 0.55 12 0.9985 Heat-killed HG8 vs. HP1 2 10.32 12 0.0001 Heat-killed HG8 vs. live HG8 2 21.97 12 0.0001 Heat-killed HG8 vs. live BD24 2 0.00 12 0.9999 Heat-killed BD24 vs. HP1 2 9.78 12 0.0002 Heat-killed BD24 vs. live HG8 2 21.42 12 0.0001 Heat-killed BD24 vs. live BD24 2 0.55 12 0.9985 HP1 vs. live HG8 2 11.65 12 0.0001 HP1 vs. live BD24 2 10.32 12 0.0001 Live HG8 vs. live BD24 2 21.97 12 0.0001

127

Table S2.5 C. elegans survival on E. coli HP1, HG8 and BD24 cells and cell-free supernatants

One-way ANOVA C. elegans survival on E. coli HP1, df = F = HG8 and BD24 cells and cell-free p < 0.0001 6 159.1 supernatant Tukey's test Test strain Hours df q p-value

OP50 cells vs. HP1 cells 48 14 23.30 < 0.0001 OP50 cells vs. HG8 cells 48 14 28.41 < 0.0001 OP50 cells vs. BD24 cells 48 14 0.00 > 0.9999 OP50 cells vs. HP1 supernatant 48 14 0.00 > 0.9999 OP50 cells vs. HG8 supernatant 48 14 0.93 0.993 OP50 cells vs. BD24 supernatant 48 14 0.00 > 0.9999 HP1 cells vs. HG8 cells 48 14 5.11 0.0349 HP1 cells vs. BD24 cells 48 14 23.30 < 0.0001 HP1 cells vs. HP1 supernatant 48 14 23.30 < 0.0001 HP1 cells vs. HG8 supernatant 48 14 22.36 < 0.0001 HP1 cells vs. BD24 supernatant 48 14 23.30 < 0.0001 HG8 cells vs. BD24 cells 48 14 28.41 < 0.0001 HG8 cells vs. HP1 supernatant 48 14 28.41 < 0.0001 HG8 cells vs. HG8 supernatant 48 14 27.48 < 0.0001 HG8 cells vs. BD24 supernatant 48 14 28.41 < 0.0001 BD24 cells vs. HP1 supernatant 48 14 0.00 > 0.9999 BD24 cells vs. HG8 supernatant 48 14 0.93 0.993 BD24 cells vs. BD24 supernatant 48 14 0.00 > 0.9999 HP1 supernatant vs. HG8 supernatant 48 14 0.93 0.993 HP1 supernatant vs. BD24 supernatant 48 14 0.00 > 0.9999 HG8 supernatant vs. BD24 supernatant 48 14 0.93 0.993

128

Table S2.6 C. elegans survival on E. coli HP1, HG8 and BD24 soluble and insoluble protein fractions

One-way ANOVA C. elegans survival on E. coli HP1, HG8 and df = F = p < BD24 soluble and insoluble proteins 6 427.2 0.0001 Tukey's test Test strain Hours df q p-value

Control (M9 buffer + OP50) vs. HP1 soluble 48 14 6.814 0.0039 protein Control (M9 buffer + OP50) vs. HG8 soluble 48 14 7.985 0.0009 protein Control (M9 buffer + OP50) vs. BD24 soluble 48 14 3.361 0.2767 protein Control (M9 buffer + OP50) vs. HP1 insoluble 48 14 42.48 < 0.0001 protein Control (M9 buffer + OP50) vs. HG8 insoluble 48 14 44.08 < 0.0001 protein Control (M9 buffer + OP50) vs. BD24 48 14 2.197 0.7112 insoluble protein HP1 soluble protein vs. HG8 soluble protein 48 14 1.17 0.9777 HP1 soluble protein vs. BD24 soluble protein 48 14 10.17 < 0.0001 HP1 soluble protein vs. HP1 insoluble protein 48 14 35.67 < 0.0001 HP1 soluble protein vs. HG8 insoluble protein 48 14 37.27 < 0.0001 HP1 soluble protein vs. BD24 insoluble 48 14 9.011 0.0003 protein HG8 soluble protein vs. BD24 soluble protein 48 14 11.35 < 0.0001 HG8 soluble protein vs. HP1 insoluble protein 48 14 34.5 < 0.0001 HG8 soluble protein vs. HG8 insoluble protein 48 14 36.1 < 0.0001 HG8 soluble protein vs. BD24 insoluble 48 14 10.18 < 0.0001 protein BD24 soluble protein vs. HP1 insoluble 48 14 45.84 < 0.0001 protein BD24 soluble protein vs. HG8 insoluble 48 14 47.44 < 0.0001 protein BD24 soluble protein vs. BD24 insoluble 48 14 1.164 0.9783 protein HP1 insoluble protein vs. HG8 insoluble 48 14 1.602 0.9073 protein HP1 insoluble protein vs. BD24 insoluble 48 14 44.68 < 0.0001 protein HG8 insoluble protein vs. BD24 insoluble 48 14 46.28 < 0.0001 protein

129

Figure S2.1 Nkp-1 protein bands from E. coli HP1 soluble protein extracts induced with different L-(+)-Arabinose concentration; 0.10% (w/v), 0.20% (w/v) and 1.0% (w/v) for 0, 2, 4, 8 and 18 hours at 25°C and 200 rpm of culturing condition. After 18 hours of L-(+)-Arabinose induction, a stronger Nkp-1 protein band was observed after induction with 0.20% (w/v) (indicated by red arrow) compared to the induction with 1.0% (w/v). The HP1 soluble protein extracts were run through SDS PAGE using the 12% polyacrylamide precast gel (BioRad).

130

Figure S2.2 Nkp-1 protein bands expressed by E. coli BL21 (λDE3) transformed with the pET28b(+)hp1 vector. The E. coli recombinant clone was grown in LB10 broth supplemented with Kanamycin (50 µg / mL) and incubated at 37°C and 200 rpm of shaking condition until the OD600 nm reached 0.7. The bacterial cultures were then induced with Isopropyl β-D-1-thiogalactopyranoside (IPTG) at different concentrations; 50 µM (5), 100 µM (6), 200 µM (7), 400 µM (8) 600 µM (9) and 800 µM (10) for 18 hours at 25°C, 200 rpm. The E. coli BL21 (λDE3) soluble protein extracts were run through SDS PAGE using the 12% polyacrylamide precast gel (BioRad). Lane 1 – 10 indicate the separation of soluble protein while lane 11 – 18 show the separation of insoluble protein. Weak protein bands were observed in lane 4, 5, 6, 7, 8, 9 and 10 indicating a small amount of Nkp-1 expression by the clones. Key: M; protein marker , 1; Empty pET28b(+) vector (before induction), 2; Empty pET28b(+) vector with 50 µM of IPTG induction, 3; pET28b(+)hp1 before induction (time 0), 4; pET28b(+)hp1 after 18 hours of incubation (0 µM IPTG), 5 pET28b(+)hp1 vector induced with 50 µM IPTG, 6; pET28b(+)hp1 induced with 100 µM IPTG, 7; pET28b(+)hp1 induced with 200 µM IPTG, 8; pET28b(+)hp1 induced with 400 µM IPTG, 9; pET28b(+)hp1 induced with 600 µM IPTG, 10; pET28b(+)hp1 induced with 800 µM IPTG, 11; Empty pET28b(+) vector (before induction), 12; Empty pET28b(+) vector with 50 µM of IPTG induction, 13; pET28b(+)hp1 before induction (time 0), 14; pET28b(+)hp1 after 18 hours of incubation (0 µM IPTG), 15 pET28b(+)hp1 vector induced with 50 µM IPTG, 16; pET28b(+)hp1 induced with 100 µM IPTG, 17; pET28b(+)hp1 induced with 200 µM IPTG, 18; pET28b(+)hp1 induced with 400 µM IPTG.

131

CHAPTER 3

Investigation of the putative mode of action and nematode response to Nkp-1

132

3.1 INTRODUCTION

In nature, predation by protozoa and nematodes is a major factor causing bacterial mortality (Matz and Kjelleberg, 2005; Pernthaler, 2005; Sun et al., 2018; Kurm et al., 2019). Therefore, many bacteria have evolved to produce toxic metabolites against predators as a defence mechanism (see section 1.7, Chapter 1). This kind of microbial- predator interaction can be regarded as a potential repository for antinematode drug discovery.

Antinematode bioactive metabolites can be isolated from a range of bacteria including soil-inhabiting and marine surface-associated strains (Tables 1.1 and 1.3, Chapter 1). These bioactive metabolites include hydrolytic enzymes, toxins, crystal proteins and secondary metabolites (Darby et al., 1999; Griffitts et al., 2005; Niu et al., 2010; Ballestriero et al., 2014; Chen et al., 2015). Some of these metabolites have been characterised and are known to target either the specific regions of the C. elegans body for example cuticle, egg shell, pharynx, intestine, vulva and anus (Sellegounder et al., 2011; Chen et al., 2015) or attack the host internal system such as the nervous system and/or the innate immunity cascade (Navarro et al., 2005; Atkinson et al., 2011). After contact with a nematicidal agent, C. elegans individuals will often undergo physical and cellular damage that leads to deteriorating fecundity, cell death (necrosis or apoptosis) and a shortened lifespan (Sellegounder et al., 2011; Zhang et al., 2016). While some metabolites such as Pp-ANP and tambjamine (produced respectively by Pseudomonas protegens and Pseudoalteromonas tunicata) (Ballestriero et al., 2010; Wei et al., 2017) have non-specific cellular targets, others bind to specific receptors on the nematode host prior to the initiation of the toxic activity. For example, the Cry toxins produced by Bacillus thuringiensis are solubilised in C. elegans’ gut where they bind to a specific receptor on the intestinal cell (i.e. glycolipid and cadherin) prior to oligomerisation and insertion into the cell membrane. As a result, the cell membrane is disrupted, causing vacuole formation due to osmotic imbalance between the cytoplasm and the extracellular environment, necrosis and nematode demise (Griffitts et al., 2005; Zhang et al., 2016; Peng et al., 2018).

133

In nature, C. elegans are in constant contact with a vast number of diverse bacteria, some of which will be beneficial as food source while others will be hazardous (Alegado et al., 2003). In response, nematodes have evolved mechanisms, ranging from resistance to tolerance and avoidance behaviour to minimise the impacts of potentially toxic bacteria (Meisel and Kim, 2014; Anderson and McMullan, 2018). The processes by which nematodes detect and avoid toxic compounds are important factors to consider when developing new nematode control strategies. Nematode avoidance behaviour can be observed promptly after the first encounter of a microbial toxic repellent, such behaviour is known as innate aversion. However, in other cases, avoidance needs to be learnt and this occurs after the initial contact and is followed by what is known as an induced associative learning avoidance behaviour upon subsequent encounters (Zhang et al., 2005). In addition to avoidance behaviour, C. elegans utilizes a well characterised immune cascade to confront harmful bacteria (Ermolaeva and Schumacher, 2014; Brassinga and Sifri, 2019). Upon exposure to toxic microbes, several distinct immunity pathways including the ILR-signalling, p38 MAPK, ERK MAPK, TGF-β, FSHR-1, ZIP- 2, Wnt/Hox and autophagy are triggered. These induced pathways provide protection to the host through the expression of several downstream genes encoding bactericidal reactive oxygen species (ROS), detoxification enzymes (e.g. superoxide dismutases and catalase) and lysozymes (Engelmann and Pujol, 2010; Ermolaeva and Schumacher, 2014). Whilst these immunity gene products serve as a protective role for nematodes, they can also be manipulated by harmful bacteria to enhance their killing effect against the host (Huffman et al., 2004). For example, Salmonella enterica infection in C. elegans relies on the increased ROS production by the nematode host to promote necrosis, resulting in intestinal damage and a shortened lifespan of the animal (Sem and Rhen, 2012).

In the previous chapter, the Nematode killing Protein-1 (Nkp-1) encoded by P. tunicata D2 was characterised. Here I aim to elucidate the mode of action (MOA) employed by the Nkp-1 producing clones (i.e. E. coli HP1 and HG8) against C. elegans and to determine any cellular damage, behavioural and/or immunity responses of the nematode as a result of the clones’ toxic activity. I found that Nkp-1 expressing bacteria colonise C. elegans causing cellular (i.e. necrosis) and physical damage (i.e. vacuole formation, pharynx distortion and internal organ damage). Also, I demonstrate that C. elegans uses 134 its innate avoidance and associated learning behaviour to avoid the Nkp-1 expressing bacteria, while differentially regulating several genes involved in the ILR signalling and p38-MAPK pathways in addition to several downstream effector proteins i.e. lysozyme and superoxide dismutase. In addition, I observed that nematodes exposed to the clone E. coli HP1 exhibited the dar (deformed anal region) phenotype, indicated by swelling of the post anal region (Battisti et al., 2017), suggesting C. elegans also uses the epithelial rectal swelling as a strategy to remove toxic bacteria (Anderson et al., 2019). The information gained in this study is believed to facilitate the development of a new drug for the control of parasitic nematodes in the future.

3.2 MATERIALS AND METHODS

3.2.1 Bacterial strains and culture condition

All bacterial strains and vectors used in this study are listed in Table 3.1. Unless otherwise stated, all of the bacterial strains were routinely grown in Lysogeny Broth (also known as Luria Bertani broth (LB10), Appendix II) and Nematode Growth Medium (Brenner, 1974) (NGM, Appendix II). Where required, antibiotics such as chloramphenicol (12.5 µg/mL), ampicillin (50 µg/mL), kanamycin (50 µg/mL) and L-(+)-Arabinose (0.2% w/v) were incorporated into the media. In every experiment which involved C. elegans’ exposure to strains HP1, HG8 and the non-toxic control BD24 bacterial lawn, assay plates were initially prepared by spreading 30 µL of the bacterial overnight culture onto the LB10 agar with antibiotics and induced with L-(+)-Arabinose for 96 hours at 25°C.

135

Table 3.1 Bacterial strains and vectors used in this study

Reference or Strain or vector Relevant characteristic or genotype source Strain E. coli EPI300-T1R F-mcrA ∆ (mrrhsdRMSmcrBC) Epicentre ɸ80dlacZ∆M15∆lacX74 recA1 endA1 araD139 ∆(ara, leu) 7697 galU galK λ- rpsL nupG trfA tonA dhfr EPI300 F– λ– mcrA Δ(mrr-hsdRMS-mcrBC) Epicentre Φ80dlacZΔM15 Δ(lac)X74 recA1 endA1 araD139 Δ(ara, leu)7697 galU galK rpsL (StrR) nupG' trfA dhfr OP50 Uracil auxotroph (Brenner, 1974) HP1 EPI300 transformed with pBAD24hp1 This study BD24 EPI300 transformed with empty pBAD24 This study vector HG8 EPI300-T1R transformed with (Ballestriero et al., pCC1FOS™ :: ZP_ 01132230.1 to ZP_ 2010) 01132246.1 HP1::GFP E. coli Nkp-1 transformed with p519ngfp This study plasmid HG8::GFP E. coli HG8 strain transformed with This study p519ngfp plasmid BD24::GFP E. coli BD24 strain transformed with This study p519ngfp plasmid A1A P. tunicata D2 genomic library non-toxic (Penesyan et al., to nematode 2012) P. aeruginosa ATCC Clinical sample American Type 9027 Culture Collection (ATCC®) Vector pCC1FOS™ :: ZP_ Fosmid backbone for genomic library of (Burke et al., 2007; 01132230.1 to ZP_ Pseudoalteromonas tunicata D2 carrying Ballesteriero et al., 01132246.1 a wild type D2 insert (13.8 kb) expressing 2010) putative antinematode activity, Cmr pBAD24 F-, Δ(argF-lac)169, (Guzman et al., φ80dlacZ58(M15), glnX44(AS), λ- 1995) , rfbC1, gyrA96(NalR), recA1, endA1, spo T1, thiE1, hsdR17, pBAD24 pBAD24hp1 P. tunicata D2 wild type gene hp1 (NCBI This study Accession: ZP_01132246.1) cloned downstream the pBAD promoter, Ampr p519ngfp High-copy-number plasmid with (Stretton et al., constitutive GFP expression; Kmr 1998)

a Inducible expression with the presence of L-(+)-Arabinose 0.2% (w/v)

136

3.2.2 Cultivation and manipulation of Caenorhabditis elegans

Caenorhabditis elegans N2 Strain Bristol were maintained on Nematode Growth Medium (NGM) agar plates seeded with E. coli OP50 (Brenner, 1974). Nematode synchronisation was performed according to the standard bleaching procedure (Stiernagle, 2006) with modifications as described in section 2.2.2 (Chapter 2).

3.2.3 C. elegans exposure to the liquid culture of Nkp-1 expressing clones

To investigate the effect of C. elegans exposure to E. coli HP1 and HG8 bacterial strains in liquid culture, 20-30 L4 stage nematodes were assayed against the bacterial cultures using the 24-wells microtiter plate as described in section 2.2.9.5 (Chapter 2). C. elegans were assayed for 72 hours at 25°C and the live nematodes in the well plates were visualised using a dissecting stereomicroscope (Olympus SZ-CTV).

3.2.4 Tagging E. coli strains with GFP-expressing plasmid

Electrocompetent cells of E. coli HP1, HG8 and BD24 strains were prepared as described by Seidman et al. (1997). The green fluorescence plasmid p519ngfp (10 ng) were transformed into the electrocompetent cells, recovered in SOC media (Appendix II) and screened for successful transformants on LB10 agar with ampicillin or chloramphenicol and kanamycin after an overnight incubation at 37°C.

3.2.5 Bacterial colonisation assay

To assess the bacterial colonisation of C. elegans, a method as described by Garsin et al. (2001) and (Portal-Celhay et al., 2012) was performed. In brief, the L4 stage nematodes were washed from the NGM plates using the M9 buffer (Appendix II) and exposed to the

137

E. coli strains HP1::GFP, HG8::GFP and BD24::GFP on the LB10 agar. Thirty nematodes were picked from plates for each treatment (i.e. 10 worms/ plate in triplicate) and used for microscopy and image analysis as described in section 3.2.7. The intensity of fluorescence signals from the acquired images was quantified using the ImageJ (http://rsbweb.nih.gov/ij/download.html). Any observed morphological changes to C. elegans body structure e.g. size reduction, pharynx distortion, degenerated internal organ, internal hatching or deformed anal region (dar) were also recorded. To count the number of bacterial cells that colonised C. elegans, 10 live nematodes were picked after 6, 12, 24, 48, 72 and 96 hours of assay and transferred onto the LB10 agar with gentamicin (100 µg/mL) to kill the surface-associating bacteria. Nematodes were subsequently washed with 10 µL of LM buffer (Appendix II) and then washed twice in LM buffer containing 100 µg/mL gentamicin and M9 buffer alone. Removal of surface bacteria was confirmed by inoculating the final M9 buffer wash onto the LB10 agar plates following an overnight incubation at 37°C. Non-viable bacteria on the LB10 media indicated a successful removal of bacteria from the nematode surface. Thereafter washed C. elegans were transferred into a 1.5 mL microcentrifuge tube containing 200 µL of Phosphate Buffer Saline (PBS, Appendix II) with 1% Trion X-100 and mechanically disrupted using sterilized pestle. Nematode lysates were serially diluted in PBS and spread plated onto the LB10 agar with appropriate antibiotics. After an overnight incubation at 37°C, number of colonies were quantified and represented as colony forming unit per nematode (CFU/nematode). Each experiment was performed in triplicate. The results of bacterial colonisation (CFU counting/nematode) and bacterial fluorescence intensity were analysed with Prism software version 8.3.0 (GraphPad Software, La Jolla, CA, USA) using a One-way ANOVA and Tukey’s pairwise comparison test. A p value < 0.05 was considered significant.

3.2.6 Assessment of C. elegans damage due to the Nkp-1 exposure

Protein extracts were obtained from the E. coli HP1, HG8 and BD24 cell pellets as described previously (see section 2.2.7, Chapter 2). Eighty to ninety synchronised L4 stage C. elegans were transferred into the 24-well microtiter plates containing 150 µL of soluble or insoluble protein extracts (at ~ 0.7 mg/mL of total protein), 40 µL of E. coli 138

OP50 (OD600nm = 2.0) and M9 buffer up to 300 µL final volume (see section 2.2.9.5, Chapter 2). After incubation in 25°C for 72 hours, 30 nematodes per treatment (10 nematodes / replicate plate) were picked and washed twice in M9 buffer. Nematodes were observed under the fluorescence microscope as in section 3.2.7 and the number of nematodes showing signs of physical damage e.g. size reduction, pharynx distortion, vacuole formation, internal hatching or degenerated internal organ (gonad and/or intestine) was recorded. Experiment was performed in triplicate. The proportion of nematode with physical damages due to protein treatment was analysed with Prism software version 8.3.0 (GraphPad Software, La Jolla, CA, USA) using a Two-way ANOVA followed by Tukey’s and Sidak’s test. A p value < 0.05 was considered significant.

3.2.7 Determination of E. coli clones HP1 and HG8 killing mechanism against C. elegans

3.2.7.1 Enzymatic assays

The overnight culture of E. coli strains HP1, HG8 and BD24 and their protein extracts were assayed for a range of hydrolytic enzyme activities. Where relevant the bacterium Pseudomonas aeruginosa which actively produces protease (Galdino et al., 2017), lipase (Hu et al., 2018), gelatinase (collagenase) (Nursyam et al., 2018) and chitinase (Sarker et al., 2019) was included as the positive control. LB10 broth or PBS (Appendix II) was used as the negative control. Protease, lipase, gelatinase/collagenase and chitinase activities were assayed on prepared skim milk agar (Waschkowitz et al., 2009), gelatine agar (Balan et al., 2012), Tween 80 agar (Samad et al., 1989) and chitin agar (Lamine et al., 2012) respectively (Appendix II). Prior to the assay, a well was made on each agar plate using the sterilised cork borer (diameter size 8 mm). Fifty microlitres of bacterial cultures or the bacterial protein extracts were inoculated into the well. Plates were incubated in 48 to 96 hours at 25°C. Positive enzymatic activity was indicated by the clearance zone observed surrounding the well. Experiment was performed in triplicate.

139

3.2.7.2 Necrosis assay

C. elegans exposed to the E. coli clone HP1, HG8 and the control strain BD24 were assayed for necrotic cell death following protocols of (Kourtis et al., 2012; Zhang et al., 2016) with modifications. Briefly, synchronized L4 stage C. elegans were washed from the E. coli OP50 lawn on NGM plates using the M9 buffer. Seventy to eighty nematodes were transferred onto the test bacterial lawn followed by incubation at 25°C for 48 hours. At least 30 nematodes per treatment (10 nematodes / replicate plate) were picked from the triplicate test bacterial lawns and washed twice in 20 µL of M9 buffer. The nematodes were transferred into each well of a 96-well microtiter plate containing 10 µM propidium iodide (Sigma Aldrich) in M9 buffer and incubated in the dark at 25°C for three hours. Propidium iodide is a small intercalating dye that is used to study the cell membrane integrity in vitro. It is membrane impermeable and when fed to the intact normal C. elegans, the dye is constrained within the gut lumen and cannot stain the cell (Los et al., 2011). In contrast, cells with destructed membrane integrity will be stained by propidium iodide indicating cell death or necrosis. After propidium iodide staining, the animals were washed with M9 buffer to remove excess stain and visualised using the fluorescence microscope as in the section 3.2.7. C. elegans was confirmed as having necrosis when the dye was visualised in cells adjacent to the intestinal lumen due to loss of membrane cell integrity. The number of C. elegans showing necrosis were quantified. Experiment was performed in triplicate. The proportion of nematode with necrosis was analysed with Prism software version 8.3.0 (GraphPad Software, La Jolla, CA, USA) using a One-way ANOVA and Tukey’s pairwise comparison test. A p value < 0.05 was considered significant.

3.2.8 Microscopy imaging

Images of C. elegans assayed in section 3.2.3 and section 3.2.6.2 were acquired using the fluorescence microscope Olympus BX61 equipped with the cellSens Dimension microscopy imaging software (Olympus). Slide containing live nematodes were prepared following protocols as previously described (Bai et al., 2014) with some modifications.

140

Briefly, C. elegans were mounted onto the agarose pad (2% w/v in M9 buffer) using sterilised worm picker and anesthetised with 10 µL of levamisole (60 µg/mL in M9 buffer). Slides were observed under the microscope using the appropriate filters. To visualise the E. coli HP1::GFP and HG8::GFP bacterial colonisation within C. elegans, the Fluorescein Isothiocyanate (FITC) filter (excitation wavelength: 495 nm, emission wavelength: 519 nm) was used. Any physical damage on the nematodes due to exposure to the bacterial cells or the protein extracts were inspected under the Differential Interference Contrast (DIC) filter. To determine any necrotic cell death, nematodes were observed under the Cyanine-3 (Cy3) filter (excitation wavelength: 550 nm, emission wavelength: 570 nm).

3.2.9 C. elegans avoidance behaviour and aversive olfactory learning against the Nkp-1 and HG8 strains

Innate aversion and induced associative learning avoidance behaviour between the naïve and trained C. elegans against the test strains; E. coli HP1and HG8 and the control clones E. coli BD24 and A1A were investigated using the food choice assay adapted from (Zhang et al., 2005; Ballestriero et al., 2016) with modifications. Prior to the assay, C. elegans eggs obtained from the standard bleaching method (Stiernagle, 2006) (see section 2.2.2, Chapter 2) were hatched to L1 young larvae on fresh NGM plates without E. coli OP50. To develop the naïve C. elegans population, the L1 larvae were transferred onto the LB10 agar plates seeded with 100 µL of control bacteria (i.e. E. coli BD24 or A1A) (Figure 3.1A). To develop the trained C. elegans population, the L1 larvae were transferred onto the LB10 agar plates that were half seeded with 50 µL of the test bacteria (i.e. E. coli HP1 or HG8) and the other half with 50 µL of the control bacteria (i.e. E. coli BD24 or A1A) (Figure 3.1B). Nematodes were then exposed to the bacterial strain/s on each plate for 72 hours in 25°C. Next, 90 to 100 of the naïve and trained nematodes were washed from each plate using the M9 buffer (Appendix II) and transferred onto the middle of assay plates that were equidistantly spotted with 30 µL of the test and control bacterial strains. Unless otherwise stated, 6 µL of 10 mM sodium azide (NaN3) were incorporated into each of the bacterial spots to immobilise nematodes that had reached the bacterial colony. The naïve and trained C. elegans were then allowed to choose either the test or 141 control bacteria as their foods for 1 hour in 25°C. The number of nematodes on either the test or control bacterial spots were counted and used to calculate the choice index (CI) and learning index (LI) as described in the following equation (1) and (2) respectively. A CI index of + 1.0 score indicates maximal attraction of C. elegans towards the control strains whilst a – 1.0 score is an evidence of nematodes repulsion. CIs were calculated from triplicate experiments and differences between treatments assessed via Unpaired Student’s t-test. A positive learning index indicates a learned avoidance behaviour by C. elegans towards the toxic bacteria with value of 2 as the maximum level. Experiment was performed in triplicate. The Unpaired Student’s t-test was applied to analyse the food choice assay results. A p value < 0.05 was considered significant.

(1)

(2)

142

A

B

Figure 3.1 Schematic diagram of C. elegans training and the assessment of innate aversion and induced associative learning avoidance behaviour between the trained and naïve nematodes. (A) Naïve C. elegans were grown on the LB10 agar cultivated with the control bacteria only (i.e. E. coli BD24 or A1A) whilst in (B), the trained C. elegans were exposed to both test (E. coli HP1 or HG8) and control bacterial strains (i.e. E. coli BD24 or A1A). The naïve and trained C. elegans were subsequently transferred onto the centre starting point (indicated by the red ‘x’ symbol) and allowed to choose their preferred bacterial cells.

3.2.10 C. elegans RNA isolation and cDNA synthesis

RNA isolation from C. elegans exposed to the toxic strain E. coli HP1 and the negative control BD24 were performed according to Zhang et al. (Zhang et al., 2012) with minor modifications. In brief, synchronized L4-stage C. elegans were transferred onto the test bacterial lawn on the 9 cm assay plates (800 to 900 nematodes/plate). Nematodes were sampled before the experiment (0 hour), after 6 and 24 hours of exposure to the bacterial clones. After collection nematodes were washed with M9 buffer and flash frozen using 143 the liquid nitrogen. C. elegans were homogenised in lysis bead tubes (Lysing Matrix EMP Bio) by vortexing using the Tissue Lyser (Qiagen) at the maximum speed for 10 minutes in 4°C. Next, the nematodes’ RNA was isolated using the Trizol reagents (Thermo Fisher) as per manufacturer’s protocol. Total nematodes’ RNA concentration was quantified using the Qubit ® 3.0 fluorometer (Thermo Fisher Scientific, Australia) whilst the purity was estimated using the Nanodrop (Thermo Fisher Scientific, Australia) based on the ratio of sample absorbance at 260/280 nm and 260/230 nm. Genomic DNA contamination within the isolated RNA was removed after incubation with RQ DNase (Promega) as per manufacturer’s protocol. cDNAs were synthesized from the total isolated RNA (0.2 µg) with random hexamer mix using the Protoscript II (NEB) following the manufacturer’s instruction.

3.2.11 Quantitative Polymerase Chain Reaction (qPCR)

The relative expression of twelve genes (Table 3.2) responsible for the immunity pathway and selected downstream genes in C. elegans was determined by quantitative PCR (qPCR). The cell division control protein 42 homolog gene (cdc-42), whose expression remained stable among the experimental samples was used as an endogenous reference gene for all reactions (Hoogewijs et al., 2008). Primers used for each gene were obtained from previous studies and listed in Table 3.2. Each primer pair was checked for binding specificity against the cDNA template in-silico using the ApE (A Plasmid Editor) software v2.0.60 (http://jorgensen.biology.utah.edu/wayned/ape/). The efficiency of the PCR reaction for each gene was optimised to close to 100% using the serially diluted pooled cDNA (Supplementary material, Table S3.1). All qPCR reactions were carried out in a Mic Real-Time PCR Cycler (Bio Molecular Systems, Australia) incorporated with built in Lin-Reg PCR software. For controls, tubes containing the pooled cDNA template (1:8-fold dilution) were used as the positive control whilst tubes containing the raw RNA (NRT) and the cDNA was substituted with the molecular grade water (NTC) were used as negative controls. The reactions were run in triplicates in a total volume of 10 μL containing the following; 1 μL of cDNA (1:8-fold dilution), 0.3 μL of each primer 10 µM (forward and reverse, 0.3 µM each final concentration), 5 μL of 2 × PerfeCTa SYBR Green SuperMix (Quantabio, Beverly, MA, USA) and 3.4 μL of ultra-pure nuclease-free 144 water (Invitrogen). The amplification reactions were subjected to an initial denaturation at 95 °C for 2 minutes, followed by 45 cycles of 95 °C for 15 seconds, 60 °C for 30 seconds and 72 °C for 30 seconds. The melting curves were obtained from 65°C to 95 °C at 0.5°C/second and determined for specific amplification indicated by a single peak on the melt curve (Supplementary material, Figure S3.6). Expression data was obtained as Ct values, which corresponds to the cycle number of the amplification reaction, at which the fluorescence of the sample exceeds the background level for the first time, and which were measured within the linear amplification ranges (Karlen et al., 2007). Linear regression was used for baseline correction of each sample as implemented in the program LinRegPCR (http:// LinRegPCR.nl) (Ruijter et al., 2009). The triplicates Ct values from triplicate biological samples were analysed using the micPCR v2.2 software (Bio Molecular Systems, Upper Coomera, QLD, Australia). Relative gene expression data was calculated using the Pfaffl method following the equation in (3) (Pfaffl, 2001). Data were represented as gene expression ratio of nematodes exposed to the toxic E. coli HP1 relative to nematodes treated with the non-toxic E. coli BD24 strain during the early stage of bacterial exposure (after 6 hours) and after 24 hours (before 50% of nematode mortality as a result of E. coli HP1 toxicity). The qPCR results were analysed using a Two-way ANOVA followed by Sidak’s test. A p value < 0.05 was considered significant.

(3)

145

Table 3.2 List of C. elegans genes and primer pairs used in qPCR reaction.

GenBank Product Gene Gene name accession Forward primer (5'-3') Reverse Primer (5'-3') length Reference symbol number (bp) Reference gene cdc-42 Cell division control protein 42 NM_063197.7 CTGCTGGACAGGAAG CTCGGACATTCTCGA 111 (Hoogewijs et homolog ATTACG ATGAAG al., 2008) p38-MAPK pathway pmk-1 Mitogen-activated protein NM_068964.3 CGACTCCACGAGAAG ATATGTACGACGGGC 203 (Li et al., 2018) kinase pmk-1 GAT ATG sek-1 Dual specificity mitogen- AB024087.1 TGCTCAACGAGCTAG ATGTTCGACGGTTTC 268 (Li et al., 2018) activated protein kinase kinase ACG ACG sek-1 nsy-1 Mitogen-activated protein kinase NM_001313625.1 TGCGATGAACTACTA CACCCAAATGACCAA 273 (Li et al., 2018) kinase kinase nsy-1 CGG ATA

ILR signalling pathway daf-2 Insulin-like receptor subunit beta; AF012437.1 TGCTGCCGAGTACGC GCAAGTGGTGTTCGA 154 (Wang et al., Receptor protein-tyrosine kinase; TGTCA CCAAC 2017) hypothtical protein daf-16 Forkhead box protein O AF032112.1 GAGAGCATTGATGGG TGGAGAAACACGAG 190 (Wang et al., CTCCC ACGACG 2017)

146

Table 3.2 (Continue) List of C. elegans genes and primer pairs used in qPCR reaction.

GenBank Product Gene Gene name accession Forward primer (5'-3') Reverse Primer (5'-3') length Reference symbol number (bp)

ERK MAPK pathway ksr-1 Kinase suppressor of U38820.1 ACAATCAGAGGTCCTAATGC GGTGTAAACTTTGACTGCG 159 (Feng et al., Ras A 2017) lin-45 Raf homolog AY455928.1 ATTCTTCGGAACTGTTGCTA ATTTCTGGTTCTCGAACCC 150 (Feng et al., serine/threonine- 2017) protein kinase mek-2 Dual specificity U21107.1 GATTGAACTGGCTGATAGTCT AAATCCAATCACTTCGTCAG 151 (Feng et al., mitogen-activated TAA 2017) protein kinase kinase mek-2 Downstream genes lys-7 Lysozyme-like protein NM_071571. GTCTCCAGAGCCAGACAATCC CCAGTGACTCCACCGCTGTA 143 (Alper et al., 7 6 GG CAC 2007) lys-8 LYSozyme NM_062682. GAGCTAGACAATATGGAATGA CCAAGGACATTCCAGTACCA 121 (Alper et al., 5 CTGTCGG GAGG 2007) bli-3 Dual oxidase 1 NM_058285. TCTTTCAAACAAGGGCGG CCTGGATTCTCATTCACACG 161 (Zou et al., 4 2013) sod-3 Superoxide dismutase NM_078363. AAAGGAGCTGATGGACACTAT AAGTTATCCAGGGAACCGA 51 (Senchuk et [Mn] 2, mitochondrial 6 TAAGC AGTC al., 2018)

147

3.3 RESULTS

3.3.1 E. coli strains expressing Nkp-1 are able to colonise and persist in the gastrointestinal lumen of C. elegans

Colonisation assays were performed with GFP labelled bacterial strains to determine if E. coli strains expressing Nkp-1 (i.e. E. coli HP1 and HG8) can colonise and persist within the C. elegans gastrointestinal system. Quantification of GFP-labelled fluorescence after 24 hours demonstrated a significantly higher degree of colonisation for strains HP1::GFP and HG8::GFP compared to the non-toxic BD24::GFP control strain

(One-way ANOVA followed by Tukey’s test; F (2, 6) = 42.57; p = 0.0044 and p = 0.0002 respectively, Figures 3.2 and 3.3, Supplementary Materials; Table S3.2). The relative fluorescence increased for these strains over the course of the experiment but remained undetected for controls (Figure 3.2 and Figure 3.3) suggesting E. coli expressing Nkp-1 are able to colonise and persist in the nematode gut. To more accurately quantify the number of HP1::GFP and HG8::GFP cells colonising the nematode gastrointestinal system, the nematode lysate was cultivated on the LB10 agar. After 6 hours of exposure to the bacterial strains, no significant differences were observed between the number of HP1::GFP and HG8::GFP bacterial colonies compared to BD24::GFP (One-way

ANOVA followed by Tukey’s test; F (2, 6) = 2.586; p = 0.7434 and p = 0.1432 respectively, Figure 3.4, Supplementary Materials; Table S3.3). However, the number of CFUs (colony formation units) of HP1::GFP and HG8::GFP started to increase after 24 hours. The highest number of HP1::GFP and HG8::GFP CFUs were counted after 96 hours of exposure with bacterial load higher in nematodes treated with E. coli HG8::GFP (One- way ANOVA followed by Tukey’s test; F (2, 6) = 44.09; HP1::GFP vs HG8::GFP; p = 0.0090; Figure 3.4, Supplementary Materials; Table S3.3). In nematodes exposed to HP1::GFP and HG8::GFP bacterial load increased 12-fold and 15-fold respectively at 96 hours of exposure compared to the initial CFU at 6-hours’ time point. In contrast, the number of BD24::GFP strain increased only 2-fold after 96 hours. This finding is in agreement with the previous microscopy observation which demonstrated the increasing abundance of HP1::GFP and HG8::GFP bacteria in C. elegans gastrointestinal system compared to the negative control BD24::GFP strain (Figures 3.2, 3.3).

148

24 hours 48 hours 72 hours 96 hours

DIC

HG8::GFP

FITC

DIC

HP1::GFP

FITC

DIC

::GFP

D24 B

FITC

Figure 3.2 E. coli HP1::GFP and HG8::GFP bacterial colonisation within C. elegans gastrointestinal system. The E. coli BD24::GFP strain was used as the negative control. Nematode images were acquired under the 10x magnification (for HP1::GFP and BD24::GFP bacterial colonisation) and 20x magnification (for HG8::GFP bacterial colonisation) using the Olympus BLX-61 equipped with the DIC and FITC filter. Scale bars indicates 100 µm (for images of HP1::GFP and BD24::GFP-treated nematodes) and 50 µm (for images of HG8::GFP-treated nematodes). Bacterial colonisation is indicated by green fluorescence emitted by the GFP-tagged bacterial cells.

149

5 0

e

d o t

a 40 m

e HP1::GFP

n

/

e HG8::GFP

c 30 n

e BD24::GFP

c

s e

r 20

o

u l

f

e

v i

t 10 a

l

e R 0 0 24 48 72 96 Time (hours) Figure 3.3 Relative fluorescence representing the colonisation of E. coli strains HP1::GFP, HG8::GFP and the negative control BD24::GFP within C. elegans body. Images were acquired using the fluorescence microscope equipped with FITC filter. The relative fluorescence was quantified using ImageJ software. An increase of nematode colonisation was shown by E. coli clones HP1::GFP (p = 0.0044) and HG8::GFP (p = 0.0002) after 24 hours of assay. After 96 hours of bacterial exposure, nematodes were heavily colonised by HP1::GFP (p = 0.0017) and HG8::GFP (p = 0.0008) compared to BD24::GFP. Data is presented as the mean of relative fluorescence from triplicate samples ± standard error.

8×103

6×103

e

d o

t a

m HP1::GFP

e 4×103

n

/ HG8::GFP

U

F BD24::GFP C 2×103

0 0 24 48 72 96 Time (hours) Figure 3.4 Bacterial colonisation assay of E. coli HP1::GFP, HG8::GFP and BD24::GFP against C. elegans. Number of viable bacteria in C. elegans gastrointestinal system is shown as colony formation unit (CFU) /nematode. No significant difference was observed on the CFU numbers of HP1::GFP (p = 0.7434) and HG8::GFP (p = 0.1432) compared to BD24::GFP strains after 6 hours of exposure. The HP1::GFP and HG8::GFP number of cells began to increase after 24 hours and showed the highest peak after 96 hours (HP1::GFP vs HG8::GFP; p = 0.0090). Data is presented as the mean of colony formation unit per nematode (CFU/nematode) from triplicates of assay ± standard error.

150

3.3.2 Exposure to E. coli strains HP1::GFP and HG8::GFP resulted in morphological changes in C. elegans

To assess any morphological changes in C. elegans due to E. coli HP1::GFP and HG8::GFP exposure, the nematodes were visualised under the fluorescence microscope. Adult nematodes treated with the negative control E. coli BD24::GFP showed a normal body length (~ 1.0 mm), a healthy pharynx, an intact internal organ and eggs without any evidence of internal hatching and a normal anal region (Figure 3.5 A, B, C, D). Whilst a small accumulation of BD24::GFP cells was observed at the C. elegans pharynx lumen (indicated by faint green fluorescence before the grinder bulb), no morphological changes were observed within this region (Figure 3.5B).

In contrast, C. elegans exposed to E. coli strains HP1::GFP and HG8::GFP showed evidence of physical changes including; smaller body size, pharynx distortion, internal hatching and internal organ damage (Figure 3.5 E, F, G). Unlike the other morphological changes, the deformed anal region (dar) was only observed on C. elegans exposed to E. coli HP1::GFP (Figure 3.5H). Bacterial colonisation indicated by the green fluorescence was also visualised at the pharynx and C. elegans’ gastrointestinal lumen particularly at the anterior (right after the terminal bulb) and at the posterior region (close to the anal region).

151

A E

B F Pharynx distortion

Terminal bulb / Internal hatching grinder

C G G

Internal organ damage

D H

Dar

Figure 3.5 Examples of C. elegans body structure after exposure to the Nkp-1 expressing clones (E. coli HP1::GFP and HG8::GFP) and the negative control strain BD24::GFP. Nematodes treated with BD24::GFP showed a normal adult body size (A), a normal pharynx (B) an intact internal organ and eggs (C) and a normal anal region (D). However, exposure to HP1::GFP and HG8::GFP resulted in several morphological changes including smaller body size indicating growth retardation (E), pharynx distortion and internal hatching (F) and internal organ damage (G). The deformed anal region (dar) (H) was only demonstrated by nematodes upon E. coli HP1::GFP exposure. Green fluorescence indicates bacterial colonisation. Images were captured using the DIC and FITC filters under 10x (A and E) or 40x (B-D, F-H) magnifications and subsequently merged. Scale bars indicate size of 100 µm (A and E) or 20 µm (B-D, F-H).

152

C. elegans exposed to the Nkp-1 expressing bacteria showed different level of physical injuries over the time course of study. After 24 hours of E. coli strains HP1::GFP and HG8::GFP exposure, C. elegans growth was retarded indicated by body length reduction compared to the normal adult nematodes fed with E. coli BD24::GFP (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 179.4; p = 0.0441 and p = 0.0007 respectively, Figure 3.6A, Supplementary Materials; Table S3.4). Nematodes exposed to those two toxic clones remained small after 96 hours and the size reduction was more pronounced on nematodes exposed to E. coli HG8::GFP compared to the HP1::GFP strain (Two-way

ANOVA followed by Tukey’s test; F (2, 6) = 179.4; p = 0.0016, Figure 3.6A, Supplementary Materials; Table S3.4).

There was no difference in the proportion of nematodes having pharynx distortion after 24, 48 and 72 hours of exposure to E. coli HP1::GFP and HG8::GFP (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 14.73; p > 0.999, p = 0.5384 and p > 0.999 respectively, Figure 3.6B, Supplementary Materials; Table S4). However, after 96 hours, an increase of pharynx distortion was observed on C. elegans exposed to E. coli HP1::GFP compared to HG8::GFP strain (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 14.73; p = 0.0105, Figure 3.6B, Supplementary Materials; Table S3.4).

Moreover, there was no difference in the proportion of internal hatching observed in nematodes exposed to either E. coli HP1::GFP or HG8::GFP after 24 hours (Two-way

ANOVA followed by Tukey’s test; F (2, 6) = 17.36; p = 0.1297, Figure 3.6C, Supplementary Materials; Table S3.4). However, after 96 hours, C. elegans exposed to E. coli HP1::GFP exhibited more internal hatching compared to nematodes treated with

HG8::GFP (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 17.36; p = 0.0088, Figure 3.6C, Supplementary Materials; Table S3.4).

A higher proportion of nematodes with internal organ damage was also observed for C. elegans exposed to E. coli HG8::GFP compared to HP1::GFP after 24 hours (Two-way

ANOVA followed by Tukey’s test; F (2, 6) = 442.1; p = 0.0338, Figure 3.6D, Supplementary Materials; Table S3.4). The proportion of nematodes with internal organ

153 damage constantly increased until all of the nematodes exposed to E. coli HG8::GFP were damaged internally after 72 hours (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 442.1; p = 0.0033, Figure 3.6D , Supplementary Materials; Table S3.4).

Unlike the other morphological changes, the deformed anal region (dar) phenotype was only detected in C. elegans treated with E. coli HP1::GFP. Half of the HP1::GFP exposed- nematodes displayed the dar after 24 hours (Figure 3.6E). The proportion of C. elegans showing dar constantly increased until 93.33 % of the nematodes demonstrated the phenotype after 96 hours (Two-way ANOVA followed by Tukey’s test; F (2, 6) = 58.84; p = 0.0092 when compared to HG8::GFP, Figure 3.6E, Supplementary Materials; Table S3.4).

Morphological changes observed on C. elegans as a result of exposure to E. coli strains HP1::GFP, HG8::GFP and non-toxic strain BD24::GFP are summarised in Table 3.3

154

A B

length (µm)

distortion (%)

Average nematodeAverage Nematodes with Nematodes pharynx

Time Time

C D

hatching (%)

organ damage organ(%)damage

Nematodes with Nematodes internal Nematodes with Nematodes internal

Time Time

E 100

)

%

(

r

a

D

h t i 50

w

e

d

o t

a

m

e N

with Nematodes dar (%) 0 24 h 48 h 72 h 96 h Time Time Figure 3.6 Proportion of C. elegans with signs of morphological changes after exposure to different GFP-tagged E. coli clones. Black, grey and blue bars indicate the proportion of morphological changes in C. elegans as a result of exposure to E. coli strains HP1::GFP, HG8::GFP and the non-toxic control BD24::GFP strains respectively. Images of C. elegans were acquired using the DIC filter and the physical condition assessed for at least 30 worms per treatment. Morphological conditions include; (A) average nematode length, (B) pharynx distortion, (C) internal hatching, (D) internal organ damage and (E) deformed anal region (dar). Graphical summaries represent the average percentage of assayed nematodes from triplicate samples ± standard error (n = 3). A p value < 0.05 was considered as statistically significant. Significance levels are indicated on figures as follows:*p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001, ns p < 0.05. 155

3.3.3 Total protein fraction of E. coli strains expressing Nkp-1 resulted in morphological changes on C. elegans

The morphological changes to C. elegans after exposure to the total soluble and insoluble protein fractions derived from E. coli clones HP1, HG8 and the negative control clone BD24 was assessed microscopically. After 72 hours, C. elegans exposed to the total soluble and insoluble protein fractions from clone BD24 displayed a normal body length (~ 1.0 mm), a healthy pharynx, no vacuole formation and no internal hatching (Figure 3.7 A, B, C, D).

In contrast, C. elegans exposed to either the soluble and/or insoluble protein of E. coli strains HP1 and HG8 after 72 hours showed several morphological changes i.e. body length reduction (indicating growth retardation), internal organ damage, pharynx distortion, vacuole formation and internal hatching (Figure 3.7 E, F, G, H). Detail comparisons of morphological changes observed on E. coli clones HG8, HP1 and the control E. coli BD24 were shown in Supplementary Materials [Figures S3.1 (average nematode size), S3.2 (internal organ damage), S3.3 (pharynx distortion), S3.4 (vacuole formation) and S3.5 (internal hatching)].

156

A E

Internal organ damage and reduced size

B F Normal pharynx Pharynx distortion

C G

Vacuole formation

No vacuole

D H

Internal hatching

Intact body and eggs

Figure 3.7 Examples of physical appearances observed on C. elegans treated with the total insoluble protein fractions of E. coli HP1, HG8 and the negative control BD24 strains. Nematodes treated with the BD24 insoluble protein extracts showed normal body structure (A-D). In contrast, C. elegans exposed to the total insoluble protein fractions from strains HP1 and/or HG8 exhibited several morphological changes i.e. size reduction and internal organ damage (E), pharynx distortion (F), vacuole formation (G) and internal hatching (H). Nematode images were acquired using the DIC filter under 10x (A and E) or 40x (B-D, F-H) magnifications. Scale bars indicate 100 µm (A and E) or 20 µm (B-D and F-H).

157

Differences in C. elegans morphological changes could be seen between exposure to either the soluble or insoluble protein fractions of E. coli HP1 and HG8. Specifically, while exposure to HP1 and HG8 soluble proteins did not result in growth retardation on C. elegans, the insoluble proteins from those two clones did negatively impact growth

(Two-way ANOVA followed by Tukey’s test; F (1, 12) = 10.47; p = 0.0059 and p = 0.0002 respectively, Figure 3.8A, Supplementary Materials; Table S3.5).

The insoluble protein treatment from HG8 also caused an increase in internal organ damage against the assayed nematodes compared to the other protein treatments (Two- way ANOVA followed by Tukey’s test; F (1, 12) = 364.5; p < 0.0001, Figure 3.8B, Supplementary Materials; Table S3.5).

The proportion of nematodes with pharynx distortion was significantly higher for those exposed to the soluble or insoluble protein fraction of HP1 compared to those exposed to either protein extract from HG8 (Two-way ANOVA followed by Tukey’s test; F (1, 12) = 0.2500; p = 0.0262 and p = 0.0056 respectively, Figure 3.8C, Supplementary Materials; Table S3.5).

An increase in vacuole formation was also observed for nematodes treated with the soluble protein of HP1 compared to the soluble protein of HG8 (Two-way ANOVA followed by Tukey’s test; F (1, 12) = 25.00; p = 0.0006, Figure 3.8D, Supplementary Materials; Table S3.5).

In addition, exposure to both protein fractions from HP1 resulted a higher proportion of internal hatching in C. elegans compared to exposure of either protein fraction from HG8

(Two-way ANOVA followed by Tukey’s test; F (1, 12) = 3.125, p = 0.0083 and p < 0.0001, Figure 3.8E, Supplementary Materials; Table S3.5). All of the morphological changes observed on C. elegans as a result of exposure to the soluble and insoluble protein fractions of E. coli strains HP1, HG8 and non-toxic control BD24 are summarised in Table 3.3

158

A B

organ damage organ(%)damage

Nematodes with Nematodes internal Average nematode length (µm) nematodeAverage length

C D

vacuole vacuole

(%)

formation

distortion (%)

Nematodes with Nematodes Nematodes with Nematodes pharynx

E

hatching (%)

with Nematodes internal

Figure 3.8 Proportion of C. elegans with morphological changes after exposure to the protein fractions of strains HP1 and HG8. Several morphological changes i.e. reduced body size based on the average nematode length (A), internal organ damage (B), pharynx distortion (C) vacuole formation (D) and internal hatching (E) were observed for nematodes exposed to protein fractions of HP1 and HG8. In contrast, those morphological changes were not observed on nematodes exposed to the negative control BD24 protein fractions. Results shown are representative data from the mean percentage of nematode proportion with morphological changes from triplicate samples ± standard error (n = 3). A p value < 0.05 was considered as statistically significant. Soluble protein: black, insoluble protein: grey. Significance levels are indicated on figures as follows: *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001, ns p < 0.05. 159

Table 3.3 Summary of morphological changes for C. elegans resulting from exposure to different E. coli bacterial cells, their soluble or insoluble protein fractions. The detected and non-detected morphological changes in the assayed nematodes are indicated by “+” and “-” symbols respectively.

Morphological changes in C. elegans Bacteria and Reduced body Vacuole protein fractions Internal hatching Pharynx distortion Internal damage Dar formation length formation E. coli HP1 Cells + + + + - + Soluble protein - + + + + - Insoluble protein + + + + + -

E. coli HG8 Cells + + + + - - Soluble protein - + + + + - Insoluble protein + + - + - -

E. coli BD24 Cells ------Soluble protein ------Insoluble protein ------

160

3.3.4 E. coli HP1 and HG8 bacterial strains and their protein extracts do not show any evidence of hydrolytic enzyme activity

In order to investigate whether the nematode killing activity by E. coli HP1 and HG8 strains were resulted from hydrolytic enzyme activity, the toxic clones and their protein extracts were screened for evidence of such activity. The BD24 overnight culture and its protein fraction were used as the negative control. A clearance zone was not observed surrounding the HP1, HG8, BD24 bacterial culture and their protein extracts (Table 3.4). In contrast, a clearance zone was observed surrounding the well containing P. aeruginosa overnight culture indicating positive secretion of protease, gelatinase/collagenase, lipase and chitinase enzymes which resulted in substrate degradation (Table 3.4).

161

Table 3.4 Assay on potential hydrolysis enzymatic activity of E. coli strains HP1 and HG8 bacterial culture and their protein extracts against the different substrates. Experiments were run in triplicate. Positive enzymatic reaction for each triplicate is denoted as ‘+++’ whilst ‘- - -’ indicates a negative result. NT denotes not tested.

Strain Sample Enzymatic activity Lipase Chitinase Protease Gelatinase / Collagenase

Nkp-1 Cell culture ------Protein Soluble ------extract Insoluble ------

HG8 Cell culture ------Protein Soluble ------extract Insoluble ------BD24 Cell culture ------Protein Soluble ------extract Insoluble ------P. aeruginosa Cell culture + + + + + + + + + + + + Protein Soluble NT NT NT NT extract Insoluble NT NT NT NT

3.3.5 Exposure to E. coli strains expressing Nkp-1 results in loss of C. elegans cell membrane integrity

The severe physical impairments observed in C. elegans treated with E. coli clones HP1 and HG8 bacterial culture could be attributed to damages at the cellular level. Pharyngeal and intestinal cells of nematodes exposed to the HP1 and HG8 strains showed evidence of loss of cell membrane integrity indicated by the uptake of propidium iodide (Figure 3.9 A-F). In total, 81.33% and 96.67% of nematodes treated with HP1 and HG8 strains respectively showed loss of cell membrane integrity, whereas none of the nematodes exposed to control strain BD24 showed evidence of membrane integrity loss (Figure 3.10). Further statistical analyses are provided in the Supplementary Material; Table S3.6).

162

Cy3 DIC Merged

A B C Merge HG8

D E F

HP1

G H I

BD24

J K L Positive control

M N O

Negative control

Figure 3.9 Propidium iodide (PI) staining on C. elegans exposed to E. coli strains HP1, HG8 and BD24 strains. PI staining on nematodes exposed to excessive heating and nematodes fed on E. coli OP50 were used as the positive and negative controls respectively. Images of nematodes were acquired under 40 x magnification using the Cy3 filter (A, D, G, J, M) and DIC filters (B, E, H, K, N). The captured images were then merged to give a precise visualisation of PI staining within C. elegans body (C, F, I, L, O). Scale bars indicate 20 µm.

163

**** 100

e ***

n

a

r

b

m

)

e

%

m

(

f

y

o

t

i

s

r

s

g

o l

e 50

t

h

n

t

i

i

s

l

w

l

e

e

c

d

o

t

a

Nematodeswith ofloss

m

e

membrane integritycells (%) N 0 HP1 HG8 BD24

Figure 3.10 Proportion of C. elegans with loss of cell membrane integrity after 48 hours of exposure to E. coli HP1 and HG8 bacteria. No loss of cell membrane integrity was observed on C. elegans treated with the negative control BD24. Results are representative data of mean percentage of nematodes showing loss of cell membrane cells integrity from triplicate assays ± standard error (n = 3). A p value < 0.05 was considered as statistically significant. *** indicates p = 0.0001 while **** indicates p < 0.0001

164

3.3.6 C. elegans has an innate and associative learned avoidance behaviour against E. coli strains HP1 and HG8

To investigate the innate and associative learned avoidance behaviour of C. elegans towards Nkp-1, naïve and trained nematodes were allowed to choose between the test E. coli strains expressing Nkp-1 (i.e. HP1 or HG8) and control strains that harbour the empty vector (i.e. BD24 or A1A respectively). The preference of naïve nematodes for the non- toxic control E. coli strains suggests an innate avoidance of C. elegans to Nkp-1. However, the CI value of trained nematodes (i.e. HP1= –0.49 ± 0.02, HG8 = - 0.78 ± 0.01, Figure 3.11) for Nkp-1 producing strains was significantly lower compared to naïve animals to HP1 (CI value = - 0.20 ± 0.07) (Unpaired t-test, F (2, 2) = 11.03; p = 0.0182, Figure 3.11A, Supplementary Materials; Table S3.7) or HG8 (CI value: – 0.48 ± 0.02)

(Unpaired t-test, F (2, 2) = 6.685; p = 0.0005, Figure 3.11C, Supplementary Materials; Table S3.7). Furthermore, this preference resulted in a learning index close to 0.3 (Figure 3.11 B, D) indicating that in addition to an innate avoidance, C. elegans has an associative learned avoidance behaviour against both E. coli HP1 and HG8 strains.

165

A 0.9 B 0.9 N2 naive 0.6 N2N2-HP1Nkp- trained1 traine d 0.6

x 0.3

x 0.3

e

e d

d

n

i

n

i

g e

0.0 n 0.0

i

c

i

n

o

r a

h e C -0.3

-0.3 L

-0.6 * -0.6 -0.9 -0.9

C 0.9 D 0.9 N2 naive 0.6 N2- HG8 trained 0.6

x 0.3

x 0.3

e

e

d

d

n

i

n

i

g

e 0.0 0.0

n

c

i

i

n

o

r

h

a e C -0.3 -0.3

L

-0.6 -0.6

-0.9 *** -0.9

Figure 3.11 Assessment for C. elegans innate and associative learned avoidance behaviour against the toxic E. coli strains HP1 and HG8 and the non-toxic E. coli BD24 and A1A strains. Choice index showing the food (bacteria) preference by the naïve and trained nematodes against HP1 versus BD24 and HG8 versus A1A are shown in (A) and (C) respectively. The learning index indicating the associative learned avoidance behaviour against the HP1 and HG8 are represented in (B) and (D) respectively. Results shown are the choice index or learning index ± the standard error calculated from triplicate assays. A p value < 0.05 was considered as statistically significant. * indicates p = 0.0182 while *** indicates p = 0.0005

166

3.3.7 Exposure to E. coli HP1 resulted in differential expression of innate immunity genes in C. elegans

qPCR was performed to investigate differential expression in selected C. elegans genes in response to the toxicity of E. coli HP1 both during the early stage of bacterial exposure (after 6 hours) and after 24 hours (before the 50% of nematode mortality). The qPCR efficiency (E) for each of the primer pairs were close to 100% and all of the correlation coefficients of R square were close to 1.0 (Supplementary Materials, Table S3.1). The presence of a single peak in the melting curves generated from the dissociation assay following the qPCR cycles (Supplementary Materials, Figure S3.6) confirmed gene amplification specificity. The expression ratio of eight genes; daf-2 and daf-16 (ILR signalling pathway), ksr-1, lin-45 and mek-2 (ERK-MAPK pathway), pmk-1, nsy-1 and sek-1 (p38 MAPK pathway) were determined based on the gene expression in nematodes exposed to E. coli HP1 relative to the gene expression in nematodes exposed to the non- toxic E. coli BD24 strain (see section 3.2.10). While an increase of daf-2 was observed after 6 hours of exposure to E. coli HP1 relative to the non-toxic E. coli BD24 (as indicated by the positive expression ratio), at 24 hours this trend was reversed with daf-2 expression lower in the E. coli HP1 compared to E. coli BD24 (Two-way ANOVA followed by Sidak’s test; F (1, 4) = 2.777; p = 0.0451, Figure 3.12A, Supplementary Materials; Table S3.8). For daf-16 there was no difference in gene expression between the two time points with both showing higher levels in nematodes exposed to E. coli HP1 relative to the non-toxic E. coli BD24 (Two-way ANOVA followed by Sidak’s test; F (1,

4) = 2.777; p = 0.4857, Figure 3.12A, Supplementary Materials; Table S3.8). For the genes involved in the p38-MAPK pathway, only sek-1 showed an increase after 6 hours of exposure to E. coli HP1 relative to E. coli BD24. Although a slight decrease in the sek-1 expression ratio was observed after 24 hours, the difference was not statistically significant (Two-way ANOVA followed by Sidak’s test; F (2, 6) = 1.381; p = 0.1173, Figure 3.12B, Supplementary Materials; Table S3.8). For the genes involved in the ERK- MAPK pathway, only ksr-1 demonstrated a small increase of expression while the other genes (lin-45 and mek-2) showed a reduced expression (indicated by an expression ratio below 1.0) following exposure to E. coli HP1 (Figure 3.12C).

167

A 2.0 n.s *

1.5

o

i t

a

r

n o

i 1.0

s

s e

r

p

x E 0.5

0.0 daf-2 daf-16

B 2.0 n.s n.s

1.5

o i

t n.s

a

r

n o

i 1.0

s

s e

r

p

x E 0.5

0.0 pmk-1 nsy-1 sek-1

C 2.0

1.5 n.s n.s n.s

o

i t

a

r

n o

i 1.0

s

s e

r

p

x E 0.5

0.0 ksr-1 lin-45 mek-2

Figure 3.12 C. elegans genes expression of different immunity cascades including the (A) daf-2/daf-16 ILR signalling, (B) p38-MAPK pathway and (C) ERK-MAPK pathway. The qPCR results are presented as the gene expression ratio in C. elegans exposed to E. coli HP1 relative to the gene expression of nematodes exposed to the non- toxic E. coli BD24 strain. A dotted line representing the gene expression ratio (1.0) at time 0 (before the experiment) is included for reference. Each data point represents the mean gene expression ± standard error from three independent biological replicates. Black and grey bars indicate genes expression of C. elegans after 6 and 24 hours respectively of exposure to E. coli HP1 relative to E. coli BD24. A p value < 0.05 (indicated by *) was considered as statistically significant while n.s. indicates not significant. 168

3.3.8 Exposure to E. coli HP1 downregulates the sod-3 and upregulates the lys-8 gene expression in C. elegans

The expression ratio of the downstream genes (bli-3, sod-3, lys-7 and lys-8) which are important for C. elegans protective response against bacterial toxicity were investigated through qPCR. An increase of sod-3 gene expression was observed after 6 hours of nematode exposure to E. coli HP1 strain relative to the non-toxic E. coli BD24. However, the relative expression of the sod-3 gene declined in these nematodes after 24 hours of treatment, as indicated by a significant reduction in the gene expression ratio (Two-way

ANOVA followed by Sidak’s test; F (3, 8) = 14.64; p = 0.0142, Figure 3.13, Supplementary Materials; Table S3.8). While higher levels of lys-8 gene expression were observed for nematodes exposed to the E. coli HP1 strain relative to the non-toxic BD24 strain at both time points, the increase was more pronounced after 24 hours (Two-way ANOVA followed by Sidak’s test; F (3, 8) = 14.64; p = 0.0052, Figure 3.13, Supplementary Materials; Table S3.8).

2.5 ** *

2.0 o

i

t

a r

1.5 n

o n.s

i n.s

s

s e r 1.0

p

x E 0.5

0.0 bli-3 sod-3 lys-7 lys-8

Figure 3.13 The expression level of several downstream genes that are important for C. elegans protective response against bacterial toxicity. The qPCR results are presented in form of gene expression ratio in nematodes exposed to E. coli HP1 relative to the gene expression in nematodes exposed to the non-toxic control strain E. coli BD24 after 6 and 24 hours of bacterial exposure. A dotted line representing the expression ratio at time 0 hour before the experiment (expression ratio:1.0) is included for reference. Each data point represents the mean gene expression ratio ± standard error from three independent biological replicates. Black and grey bars indicate genes expression of C. elegans after 6 and 24 hours of exposure to E. coli HP1 relative to E. coli BD24. A p value < 0.05 was considered as statistically significant while n.s. denotes not significant. * indicates p = 0.0142 and ** indicates p = 0.0052

169

3.4 DISCUSSION

A novel Nematode killing protein-1 (Nkp-1) produced by P. tunicata D2 was characterised in Chapter 2. In the current chapter, the potential mode of action (MOA) of Nkp-1 expressing clones (E. coli strains HP1 and HG8) against C. elegans was investigated. The key discoveries found in this chapter are further discussed below.

3.4.1 Nkp-1 expressing E. coli clones kill C. elegans via a step by step mode of action (MOA)

The ability of the Nkp-1 expressing clones to colonise the nematode was inspected and the degree of morphological changes caused by the colonising bacteria and their protein extracts was evaluated. These morphological changes (except for the deformed anal region (dar)) to exposed C. elegans led to the hypothesis that the Nkp-1 expressing clones were causing cellular level (necrosis) damage to C. elegans. A step-by-step Nkp-1 MOA against C. elegans is proposed as indicated in Figure 3.14 and further described below.

3.4.1.1 Stage 1: Ingestion and digestion of Nkp-1 expressing bacteria by C. elegans

I have found that only exposure to Nkp-1 expressing bacteria (E. coli HP1 and HG8) or their protein extracts is toxic to C. elegans, with cell-free supernatants having no impact on nematode survival. These observations raise the question of how C. elegans becomes exposed to the cytoplasmic protein Nkp-1? One mechanism maybe related to nematode ingestion. C. elegans’ pharynx is the main entrance route for food bacteria. When the pharyngeal muscles contract, bacteria are ingested, concentrated, pulverized (using a “grinder” located in the terminal bulb) and passed into the intestine through the pharyngeal-intestinal valve for degradation and nutrient absorption (Albertson and Thompson, 1976; Avery and Shtonda, 2003). However, upon ingestion of toxic bacteria, including E. coli strains HP1 and HG8, this grinding activity may also result in the liberation of toxic metabolites, such as Nkp-1(Figure 3.14A). While future work is 170 required to demonstrate the release of Nkp-1 by the nematode pharyngeal grinder, similar mechanisms have previously been demonstrated for the release of Cry toxin from B. thuringiensis (Griffitts et al., 2005; Zhang et al., 2016). In addition, nematode species lacking the pharyngeal grinder bulb i.e. Pristionchus pacificus are resistance to B. thuringiensis producing the Cry toxin (Wei et al., 2003), further suggesting that pharyngeal grinding in C. elegans may enable the release of Nkp-1 from E. coli strains HP1 and HG8.

3.4.1.2 Stage 2: Nkp-1 triggers cellular level (necrosis) damage in C. elegans

Upon bacterial ingestion and Nkp-1 liberation, the toxic protein may function as an undetermined degradative enzyme targeting C. elegans body components (i.e. protein, lipid, gelatine, collagen and chitin) or directly trigger cellular damage (necrosis). However, as there was no evidence of lipase, protease, chitinase or gelatinase/collagenase activity (Table 3.4), Nkp-1 killing mechanism via degradative enzymatic activity is unlikely. Rather, microscopic observation of nematodes exposed to the Nkp-1 expressing clones suggested that Nkp-1 may function as a pore-forming toxin (PFT) causing necrotic cell death (necrosis), as intestinal cells were permeable to propidium iodide (PI) staining (Figures 3.9 and 3.10). (Los et al., 2011; Timmons et al., 2013; Dementiev et al., 2016).

Two different models; Cry5B-like MOA (based on Cry toxin by B. thuringiensis) and aerolysin toxin-like MOA (based on aerolysin toxin by Aeromonas sp.) (Figure 3.14B) (Griffitts et al., 2005; Szczesny et al., 2011) are proposed to describe the Nkp-1 mechanism in causing necrosis against C. elegans cells. In both models, the Nkp-1 protein is solubilised in the nematode intestine prior to it binding to a yet undetermined glycoconjugate receptor of intestinal cells for example the invertebrate-specific glycolipid in Model A (Griffitts et al., 2005) or the glycosyl phosphatidyl inositol (GPI)- anchored protein in Model B (Cowell et al., 1997; Szczesny et al., 2011; Shi et al., 2020) (Figure 3.14B). To activate toxicity, aerolysin needs to be cleaved at the C-termini by the host gut protease (Iacovache et al., 2008; Szczesny et al., 2011). However, whether or not the Nkp-1 requires proteolytic activation is still unknown. Next, as occurs for both

171

Cry5B and aerolysin toxins, the Nkp-1 is predicted to oligomerize generating a ring-like structure inserted into the nematode intestinal cells hence resulting in pore formation (Figure 3.14B). For β-PFT toxins harbouring an anti-parallel β-barrel (as displayed by the Nkp-1 3D protein model, see section 2.3.11, Chapter 2) toxin oligomerisation is a crucial step prior to the oligomers insertion step into the targeted host cell (Iacovache et al., 2008). There is evidence that the cadherin-like receptor in C. elegans facilitates the Cry5B toxin to oligomerize and form pores (Peng et al., 2018). However, whether the cadherin- like receptor is required for Nkp-1 oligomerization and pore formation warrants further investigation. This can be determined by performing nematode killing assay of Nkp-1 expressing strains or their protein extracts against C. elegans mutants lacking cadherin expression (Peng et al., 2018). Following the insertion of Nkp-1 oligomers, the plasma membrane of nematode intestinal cell is proposed to be punctured, resulting in vacuole formation due to osmotic imbalance between the cytoplasm and the extracellular environment followed by necrosis and physical damage (Griffitts et al., 2005; Zhang et al., 2016; Peng et al., 2018) (Figure 3.14A).

3.4.1.3 Stage 3: Nkp-1 expressing bacteria cause a range of physical damages in C. elegans

Exposure of C. elegans to Nkp-1 results in severe growth retardation (Figure 3.14A). Consistent with this finding, toxic proteins Pp-ANP1a and Pp-ANP2a produced by Pseudomonas protegens and by Aspergillus flavus and Aspergillus parasitics also cause growth reduction in C. elegans (Feng et al., 2017; Wei et al., 2017). Similarly, exposure to E. coli clones expressing the B. thuringiensis Cry21Fa1 or Cry21Ha1 toxins results in reduced C. elegans body size (Iatsenko et al., 2014). Nematode growth retardation upon exposure to these toxins could be a result of reduced nutrient absorption efficiency caused by the perturbed intestinal cells and damaged microvilli (Marroquin et al., 2000).

An increase of pharynx distortion was also observed on C. elegans exposed to the Nkp-1 expressing clones and their protein extracts (Figure 3.14A). Pharynx distortion has been associated with “rigor mortis” and increasing cytoplasmic calcium (Ca2+) which are the

172 early signs of dying muscle cells due to necrosis in C. elegans pharyngeal and gastrointestinal system (Galimov et al., 2018). Whilst pharynx distortion could be an indicator of ageing in nematodes, it also can occur as an immediate impact of bacterial evasion within the pharyngeal cells (particularly at the terminal bulb), resulting in nematode mortality (Zhao et al., 2017). Indeed, other studies have shown an increase in this phenotype as a result of uncontrolled bacterial proliferation (Haskins et al., 2008).

Vacuole formation and internal organ damage (indicated by excessive physical discoloration, shrinking intestine and degenerated gonadal system and germline cells) was also observed in C. elegans exposed to the Nkp-1 expressing clones and their protein extracts (Figure 3.14A). Vacuole formation which may represent as the swelling necrotic- like cell death (Xu et al., 2001) is strongly linked to necrosis in C. elegans (Nikoletopoulou and Tavernarakis, 2014). Similar phenotypes have been observed for nematodes exposed to microbial pathogens (Nikoletopoulou and Tavernarakis, 2014; Durai and Balamurugan, 2017; Somasiri et al., 2020) and other toxins, including pore forming toxins (PFT) such as Cry toxin produced by B. thuringiensis (Huffman et al., 2004; Brito et al., 2019) and cytolysin produced by V. cholerae (Cinar et al., 2010; Durai and Balamurugan, 2017).

3.4.1.4 Stage 4: Enhanced Nkp-1 expressing bacterial colonisation

The intestine not only acts as a site for nutrient absorption but also provides important protective immunity against harmful microorganisms (Irazoqui et al., 2010; Pukkila- Worley and Ausubel, 2012). Several well characterised immunity cascades for example p38 MAPK, FSHR-1, ZIP-2, Wnt/Hox, UPR (unfolded protein response) and autophagy (Bischof et al., 2008; Irazoqui et al., 2008; Jia et al., 2009; Powell et al., 2009; Shivers et al., 2009; Estes et al., 2010) are also activated in C. elegans intestine and subsequently modulate genes encoding antimicrobial effector molecules such as antimicrobial peptides, caenopores, lysozymes, lectins and reactive oxygen species (ROS) (Engelmann and Pujol, 2010). Given the importance of C. elegans intestine in immunity, severe intestinal

173 damage as a result of exposure to Nkp-1 may amplify the nematode susceptibility to bacterial proliferation and deteriorate its fitness. Indeed, both direct microscopic observation and bacterial CFU counts indicated that E. coli cells expressing Nkp-1 are able to proliferate in the intestine of C. elegans (Figures 3.2, 3.3 and 3.4). This observation could be linked to the MOA of bacteria producing protein toxins i.e. indirectly hijacking the host immune system and its antimicrobial effectors hence improving the bacterial colonisation of the host (Los et al., 2013).

The end result of this proposed MOA for Nkp-1 described above is C. elegans mortality. Interestingly, rapid nematode death associated with high proportion of internal organ damage (see Figures 3.6D and 3.8B) was predominantly observed in C. elegans exposed to E. coli HG8 and its insoluble protein extract compared to the HP1 strain. This could be linked to the higher Nkp-1 expression in HG8 cells compared to the HP1 strain (see the protein separation result using SDS PAGE, section 2.3.11, Chapter 2) as a result of the different expression system and/or the function of the additional genes encoded on the HG8 fosmid (see sections 2.4.1 and 2.4.2, Chapter 2).

174

5 2 (hours) 1

4 3

1 2 3 1 2 3 4

Figure 3.14 Schematic diagram representing the proposed mode of action (MOA) of Nkp-1 expressing bacteria (E. coli HP1 and HG8) against C. elegans. Above proposed step by step (numbers 1 to 5; circled in red) nematode killing mechanism of Nkp-1 expressing bacteria in C. elegans body and the time frame (top right) of each Nkp-1 intoxication phase (starting from 0 hour to 96 hours of bacterial exposure). Below proposed models A and B of Nkp-1 MOA in causing necrotic cell death in C. elegans gastrointestinal system (stages of MOA are numbered and circled in green), see text for more details. Figure was designed using Biorender at https://biorender.com/

175

3.4.2 C. elegans response against Nkp-1 expressing bacteria

C. elegans showed both an innate and an associative learned avoidance behaviour against the Nkp-1 expressing clones suggesting avoidance is the nematode’s initial protective strategy against the bacterial toxicity. In addition, exposure to E. coli HP1 resulted in differential expression of innate immunity genes in the nematodes. The overall C. elegans’ responses against the Nkp-1 expressing strains are shown in Figure 3.15 and further discussed as below.

3.4.2.1 C. elegans avoids Nkp-1 expressing clones and shows associative learning behaviour

If Nkp-1 is to be applied as a new antinematode drug or biopesticide in the future, the behavioural response of the targeted parasites towards the active compound should be considered. This is particularly true for the development of bio-nematicide which needs to be ingested in order to be active (Betz et al., 2000; Senthil-Nathan, 2015) or for the utilisation of repellent antinematode agent that minimises parasitic nematodes invasion (Nandi et al., 2015). In this study I found that while naïve C. elegans displayed an innate preference for non-toxic clones over those expressing Nkp-1, this behaviour was more pronounced for nematodes that had previously been exposed to Nkp-1 (i.e. trained individuals, see Figure 3.11). These observations suggest that C. elegans employs both an innate and an associative learning avoidance behaviour as a primary strategy to escape from toxic Nkp-1 expressing strains. In nature, C. elegans can engage in an innate avoidance behaviour as a primary protection mechanism against the toxic producing microbes while preferring the non-toxic bacteria as the food source. The fact that C. elegans shows avoidance behaviour against the toxic microbes has been demonstrated in previous studies with bacterial strains B. thuringiensis, C. burnetii and V. parahaemolyticus (Hasshoff et al., 2007; Sellegounder et al., 2011; Luo et al., 2013; Battisti et al., 2017). Avoidance behaviour can also be learnt after the ingestion of some of the toxic bacteria, hence inducing the associative learning avoidance behaviour in trained animals (Zhang et al., 2005; Meisel and Kim, 2014). The observed innate and associative learning avoidance behaviours of C. elegans against Nkp-1 in this study might 176 indicate the Nkp-1 potentials as a dual function-bio-nematicide either as a parasitic nematodes’ repellent or as a nematicidal agent which can kill the parasites upon Nkp-1 ingestion. The dual mechanisms of parasitic nematode biocontrol were also shown by Pseudomonas chlororaphis due to the extracellular production of pyrrolnitrin and hydrogen cyanide (Nandi et al., 2015). In addition, for the development of new industrial biopesticide, one possibility is to clone the Nkp-1 encoding genetic materials into the plant host (i.e. crops) to diminish parasitic nematode infestation. This method has been applied against several genetically engineered plants expressing the Cry toxin from B. thuringiensis (Betz et al., 2000; Li et al., 2007) or transgenic crops that produce synthetic nematode repellent peptides and maize cystatin that impedes nematode digestive cysteine proteinases (Roderick et al., 2012).

3.4.2.2 C. elegans internal hatching and dar formation

Internal egg hatching was also observed in C. elegans individuals exposed to the Nkp-1 expressing strains or their corresponding protein extracts (Figure 3.15). This phenotype has been observed in C. elegans challenged with Vibrio parahaemolyticus and when sequentially exposed to Staphylococcus aureus and Proteus mirabilis (Sellegounder et al., 2011; JebaMercy and Balamurugan, 2012). Internal hatching allows the eggs to hatch within the parental nematodes and acts to protect progeny against starvation and bacterial toxicity (Chen and Caswell-Chen, 2004; Mosser et al., 2011; Souza et al., 2019). However, this phenotype adversely diminished the survival of the parent nematode by damaging the gonads hence resulting in matricidal death (Mosser et al., 2011). Therefore, it is possible that internal egg hatching occurs as part of a general stress response to Nkp- 1 and could be used by C. elegans as a mechanism of maintaining its population.

The deformed anal region (dar) phenotype or swelling of the post anal was observed on C. elegans exposed to E. coli HP1 (Figure 3.15). Previous studies have suggested that the dar phenotype acts as an immune response against bacteria as mutants lacking this phenotype experienced high rates of infection (Mendoza De Gives et al., 1999; Hodgkin et al., 2000; Jain et al., 2009). Dar formation has also been observed in C. elegans exposed 177 to a number of other microbes including Microbacterium nematophilum (Hodgkin et al., 2000), Corynebacterium diphtheriae, Corynebacterium ulcerans and Corynebacterium glutamicum (Antunes et al., 2016), Corynebacterium striatum (Souza et al., 2019), Saccharomyces cerevisiae (Jain et al., 2009) and Coxiella burnetii (Battisti et al., 2017). Unlike M. nematophilum which was abundantly found adhering to the nematode rectum (Nicholas and Hodgkin, 2004), E. coli HP1 colonises the C. elegans intestinal lumen particularly at the anterior region. Similarly, C. elegans exposed to S. cerevisiae and C. burnetii also demonstrated intestinal lumen colonisation but without cell accumulation at the nematode rectum (Jain et al., 2009; Battisti et al., 2017).

It has been previously suggested, that the dar phenotype is a mechanism employed by C. elegans to ease bacterial removal from the anal opening through the epithelial rectal swelling mechanism (Hodgkin et al., 2000; Jain et al., 2009; Anderson et al., 2019). In addition, nematodes with dar also show an avoidance behaviour against the toxic bacteria which may indirectly facilitate the clearance of toxic bacterial infections from the intestine (Anderson et al., 2019). Interestingly, the dar phenotype was not seen in C. elegans exposed to E. coli HG8. One possible explanation for this observation is that the HP1 bacterial colonisation and its toxic effect is milder compared to the HG8 strains. This observation could be linked to the higher abundance of HG8 bacterial cells in C. elegans compared to the HP1 strain within 24 hours of bacterial exposure (see Figures 3.2, 3.3 and 3.4). Consequently, within 24 hours, while 50% of HP1-exposed nematodes showed the dar phenotype, almost half of the HG8-exposed nematodes were severely damaged leading to their rapid death.

3.4.2.3 The daf-2/daf-16 ILR signalling and the downstream genes sod-3 and lys-8 maybe essential for C. elegans immune defence against the Nkp-1 expressing strains

As a protective response to the adverse effect of colonisation by toxic bacteria, C. elegans is capable of a coordinated regulation of a number of immune defence pathways (Pukkila- Worley and Ausubel, 2012; Kim, 2015; Wani et al., 2020). In this study, qPCR analysis revealed a relative increase in daf-2 and daf-16 gene expression (involved in ILR- 178 signalling pathway) after 6 hours exposure to E. coli HP1 compared to the control. This finding suggests that the ILR signalling pathway is either triggered during the early stages of E. coli HP1 exposure (Figure 3.15) or there is/are another additional pathway(s) regulating DAF-16 such as the germline and the JNK pathways that are also responsible in regulating C. elegans life span (Mukhopadhyay et al., 2006). After 24 hours, the daf-2 expression ratio was significantly lower which may represent part of C. elegans strategy to overcome E. coli HP1 colonisation. In previous studies, loss of the insulin receptor, daf-2, caused C. elegans to become more resistant to the Cry5B toxin produced by B. thuringiensis and other bacterial strains including P. aeruginosa, S. typhimurium and C. burnetii (Evans et al., 2008; Jia et al., 2009; Chen et al., 2010; Battisti et al., 2017). Under stress conditions, for example during bacterial colonisation, DAF-2 activity is reduced resulting in DAF-16 activation and the upregulation of a variety of downstream genes encoding for antimicrobials thus providing protection against the toxic microbes (Wani et al., 2020). In this study, expression of the lysozyme encoding gene lys-8 was significantly higher in nematodes exposed to E. coli HP1 compared to the control (Figure 3.15). Lysozymes are small molecules that can digest peptidoglycan component of the bacterial cell walls acting as antimicrobial. (Mallo et al., 2002). In previous studies, lys- 8 expression was upregulated upon exposure to Serratia marcescens and M. nematophilum (Mallo et al., 2002; O’Rourke et al., 2006) modulated by the daf-2/daf-16 and DBL-1/TGF-β signalling pathways (Murphy et al., 2003; Ewbank, 2006). Upon exposure to the detrimental bacteria, lys-8 was highly expressed at the nematode terminal pharyngeal bulb, pharyngeal gland cells and intestine (Mallo et al., 2002; Alper et al., 2007; Schulenburg and Boehnisch, 2008). Interestingly, colonisation of the C. elegans pharynx and intestine was observed for E. coli HP1 which may have resulted in the observed increase in lys-8 gene expression in the host nematode.

In contrast to lys-8, the sod-3 gene encoding superoxide dismutase which functions in reactive oxygen species (ROS) detoxification was significantly reduced after 24 hours compared to 6 hours of E. coli HP1 exposure (Figure 3.15). This finding is interesting because sod-3 has previously been shown to be upregulated in daf-2 mutants which was suggested to confer resistance against the oxidative stress in aging nematode and during microbial attack (Honda and Honda, 1999; Bai et al., 2014). However, the sod-3 function is argued to be more involved in stress response signalling rather than detoxifying the 179

ROS in C. elegans (see review by (Wang et al., 2018)). Although sod-3 has been shown to be upregulated during oxidative stress, its contribution to the overall SOD protein expression in C. elegans remains very small (Doonan et al., 2008). C. elegans possesses five superoxide dismutases (sod-1 – sod-5) of which sod-1 represent as the highest SOD activity contributor (Doonan et al., 2008). Furthermore, a mutation in sod-3 did not result in reduced C. elegans lifespan or increased bacterial accumulation after exposure to E. coli or S. typhimurium compared to the wild type (Portal-Celhay et al., 2012). Taken together, sod-3 may function only during the early stage of E. coli HP1 exposure and then become less relevant once other sod genes are upregulated. Further research using sod- defective mutants and analysis of the expression of other sod genes is needed to elucidate the relative importance of SOD proteins for C. elegans immunity against the E. coli HP1 strain.

3.4.2.4 sek-1 maybe important for the dar phenotype in C. elegans exposed to E. coli HP1

This study shows that genes involved in the ERK-MAPK pathway were not upregulated in C. elegans exposed to the E. coli HP1 clone. In contrast, the ERK-MAPK pathway was found to be upregulated upon exposure to the dar bacterial inducer; M. nematophilum (Nicholas and Hodgkin, 2004). However, an increase in the expression of sek-1, encoding for a mitogen-activated protein kinase activator protein from the p38-MAPK pathway, was observed in C. elegans when exposed to the E. coli HP1 clone, suggesting it may play a role in the dar phenotype observed here. There is evidence for the differential expression of sek-1 across different parts of C. elegans body, with high levels of expression in the excretory canal, uterine-vulva cells, neurones and rectal epithelial cells (Tanaka-Hino et al., 2002). This may lend support to the functional role of SEK-1 as an inducer of the dar phenotype of C. elegans epithelial cells observed in this study (Figure 3.15). However, the direct or indirect interaction of sek-1 with the downstream genes or the target rectal epithelial cells is still unknown. Nevertheless, in support of this hypothesis, sek-1 gene expression was shown to be important in C. elegans defence against the dar bacterial inducer; C. burnetii (Battisti et al., 2017). A mutation in sek-1 resulted in a diminished dar phenotype hence affecting the nematode’s ability to remove 180 the colonising bacteria (Battisti et al., 2017). Further nematode killing and bacterial colonisation assays using the sek-1 defective mutant nematode will be required to confirm such a role for SEK-1.

Figure 3.15 Schematic diagram representing the proposed C. elegans response to the toxicity of Nkp-1 expressing clones and their protein extracts. As the primary protective strategy, C. elegans uses its innate and associative learning avoidance behaviour to escape from the Nkp-1 expressing bacterial cells. The nematode also employs its innate immunity pathways specifically the ILR-signalling and p38-MAPK involving the daf- 2/daf-16 and sek-1 genes respectively. The ILR-signalling pathway upregulates the lysozyme (lys-8) gene expression which potentially enables bacterial cell wall degradation. In contrast, the superoxide dismutase (sod-3) is reduced. The upregulation of sek-1 gene expression may be linked to the increase of the dar phenotype in C. elegans exposed to E. coli HP1 (indicated by the dotted black arrow) presumably to ease the removal of invading bacteria through epithelial rectal swelling mechanism. In order to maintain the survival of new progenies in an unfavourable condition (i.e. when the toxic bacteria are present) the nematode eggs hatch within the parent’s gonadal system. The up- and down-regulation of each responsible genes are indicated by the red arrows. Figure was designed using Biorender at https://biorender.com/

181

3.5 CONCLUSION

In this study, the Nkp-1 mode of action (MOA) against C. elegans and the nematode response against the Nkp-1 toxicity have been extensively investigated. Based on the two models (Cry toxin-like MOA and aerolysin toxin-like MOA), Nkp-1 appears to act as a novel pore forming toxin (PFT) which can disrupt the targeted host membrane cell, leading to the loss of plasma membrane integrity and necrotic cell death. Injuries at cellular level (necrosis) cause detrimental physical damages including internal organ damage, vacuole formation, pharynx distortion and growth retardation. Severe damages observed on C. elegans particularly at the intestinal region may reduce the nematode immunity, hence enhancing the subsequent colonisation by the Nkp-1 expressing bacteria resulting in a shortened nematode lifespan.

I have found that C. elegans employs its innate and learned avoidance behaviours to escape from the toxic Nkp-1 expressing E. coli clones. Furthermore, the ILR-signalling pathway involving the daf-2 and daf-16 genes appears to be important to modulate the downstream genes encoding for lysozyme and superoxide dismutase. In addition, the upregulation of sek-1 from the p38-MAPK pathway maybe important for the formation of dar (deformed anal region) in C. elegans, presumably to facilitate the removal of the invading bacteria via epithelial rectal swelling process. In conclusion, this chapter has provided greater insight to the MOA of Nkp-1 against C. elegans and the strategy C. elegans uses in response to this newly characterised toxin. These findings may facilitate the future developments of Nkp-1 as a novel antinematode drug or biopesticide.

182

3.6 REFERENCES

Albertson, D. G., & Thompson, J. (1976). The pharynx of Caenorhabditis elegans. Philosophical Transactions of the Royal Society of London. B, Biological Sciences, 275(938), 299-325. Alegado, R. A., Campbell, M. C., Chen, W. C., Slutz, S. S., & Tan, M. W. (2003). Characterization of mediators of microbial virulence and innate immunity using the Caenorhabditis elegans host–pathogen model. Cellular Microbiology, 5(7), 435-444. Alper, S., McBride, S. J., Lackford, B., Freedman, J. H., & Schwartz, D. A. (2007). Specificity and complexity of the Caenorhabditis elegans innate immune response. Molecular and Cellular Biology, 27(15), 5544. Anderson, A., Chew, Y. L., Schafer, W., & McMullan, R. (2019). Identification of a conserved, orphan G protein-coupled receptor required for efficient pathogen clearance in Caenorhabditis elegans. Infection and Immunity, 87(4), e00034- 00019. Anderson, A., & McMullan, R. (2018). Neuronal and non-neuronal signals regulate Caernorhabditis elegans avoidance of contaminated food. Philosophical Transactions of the Royal Society B Biological Science, 373(1751), 20170255. Antunes, C. A., Clark, L., Wanuske, M. T., Hacker, E., Ott, L., Simpson-Louredo, L., de Luna, M., Hirata, R., Jr., Mattos-Guaraldi, A. L., Hodgkin, J., & Burkovski, A. (2016). Caenorhabditis elegans star formation and negative chemotaxis induced by infection with corynebacteria. Microbiology, 162(1), 84-93. Atkinson, S., Goldstone, R. J., Joshua, G. W., Chang, C.-Y., Patrick, H. L., Cámara, M., Wren, B. W., & Williams, P. (2011). Biofilm development on Caenorhabditis elegans by Yersinia is facilitated by quorum sensing-dependent repression of type III secretion. PLoS Pathogens, 7(1), e1001250. Avery, L., & Shtonda, B. B. (2003). Food transport in the C. elegans pharynx. The Journal of Experimental Biology, 206, 2441-2457. Bai, Y., Zhi, D., Li, C., Liu, D., Zhang, J., Tian, J., Wang, X., Ren, H., & Li, H. (2014). Infection and immune response in the nematode Caenorhabditis elegans elicited by the phytopathogen Xanthomonas. Journal of Microbiology and Biotechnology, 24(9), 1269-1279. Balan, S. S., Nethaji, R., Sankar, S., & Jayalakshmi, S. (2012). Production of gelatinase enzyme from Bacillus spp isolated from the sediment sample of Porto Novo coastal sites. Asian Pacific Journal of Tropical Biomedicine, 2(3), S1811-S1816. Ballestriero, F., Daim, M., Penesyan, A., Nappi, J., Schleheck, D., Bazzicalupo, P., Di Schiavi, E., & Egan, S. (2014). Antinematode activity of Violacein and the role of the insulin/IGF-1 pathway in controlling violacein sensitivity in Caenorhabditis elegans. PloS One, 9(10), e109201-e109201. Ballestriero, F., Nappi, J., Zampi, G., Bazzicalupo, P., Di Schiavi, E., & Egan, S. (2016). Caenorhabditis elegans employs innate and learned aversion in response to bacterial toxic metabolites tambjamine and violacein. Scientific Reports, 6, 29284-29284.

183

Ballestriero, F., Thomas, T., Burke, C., Egan, S., & Kjelleberg, S. (2010). Identification of compounds with bioactivity against the nematode Caenorhabditis elegans by a screen based on the functional genomics of the marine bacterium Pseudoalteromonas tunicata D2. Applied and Environmental Microbiology, 76(17), 5710-5717. Battisti, J. M., Watson, L. A., Naung, M. T., Drobish, A. M., Voronina, E., & Minnick, M. F. (2017). Analysis of the Caenorhabditis elegans innate immune response to Coxiella burnetii. Innate Immunity, 23(2), 111-127. Betz, F. S., Hammond, B. G., & Fuchs, R. L. (2000). Safety and advantages of Bacillus thuringiensis-protected plants to control insect pests. Regulatory Toxicology and Pharmacology, 32(2), 156-173. Bischof, L. J., Kao, C. Y., Los, F. C., Gonzalez, M. R., Shen, Z., Briggs, S. P., van der Goot, F. G., & Aroian, R. V. (2008). Activation of the unfolded protein response is required for defenses against bacterial pore-forming toxin in vivo. PLoS Pathogens, 4(10), e1000176. Brassinga, A. K. C., & Sifri, C. D. (2019). The Caenorhabditis elegans model of Legionella infection. In C. Buchrieser & H. Hilbi (Eds.), Legionella: Methods and Protocols (pp. 371-397). New York, NY: Springer New York. Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics, 77(1), 71-94. Brito, C., Cabanes, D., Sarmento Mesquita, F., & Sousa, S. (2019). Mechanisms protecting host cells against bacterial pore-forming toxins. Cellular and Molecular Life Sciences, 76(7), 1319-1339. Chen, C.-S., Bellier, A., Kao, C.-Y., Yang, Y.-L., Chen, H.-D., Los, F. C. O., & Aroian, R. V. (2010). WWP-1 is a novel modulator of the DAF-2 insulin-like signaling network involved in pore-forming toxin cellular defenses in Caenorhabditis elegans. PloS One, 5(3), e9494. Chen, J., & Caswell-Chen, E. P. (2004). Facultative vivipary is a life-history trait in Caenorhabditis elegans. Journal of Nematology, 36(2), 107-113. Chen, L., Jiang, H., Cheng, Q., Chen, J., Wu, G., Kumar, A., Sun, M., & Liu, Z. (2015). Enhanced nematicidal potential of the chitinase pachi from Pseudomonas aeruginosa in association with Cry21Aa. Scientific Reports, 5, 14395. Cinar, H. N., Kothary, M., Datta, A. R., Tall, B. D., Sprando, R., Bilecen, K., Yildiz, F., & McCardell, B. (2010). Vibrio cholerae hemolysin is required for lethality, developmental delay, and intestinal vacuolation in Caenorhabditis elegans. PloS One, 5(7), e11558. Cowell, S., Aschauer, W., Gruber, H. J., Nelson, K. L., & Buckley, J. T. (1997). The erythrocyte receptor for the channel‐forming toxin aerolysin is a novel glycosylphosphatidylinositol‐anchored protein. Molecular Microbiology, 25(2), 343-350. Darby, C., Cosma, C. L., Thomas, J. H., & Manoil, C. (1999). Lethal paralysis of Caenorhabditis elegans by Pseudomonas aeruginosa. Proceedings of the National Academy of Sciences of the USA, 96(26), 15202. Dementiev, A., Board, J., Sitaram, A., Hey, T., Kelker, M. S., Xu, X., Hu, Y., Vidal- Quist, C., Chikwana, V., Griffin, S., McCaskill, D., Wang, N. X., Hung, S.-C., Chan, M. K., Lee, M. M., Hughes, J., Wegener, A., Aroian, R. V., Narva, K. E., & Berry, C. (2016). The pesticidal Cry6Aa toxin from Bacillus thuringiensis is structurally similar to HlyE-family alpha pore-forming toxins. BMC Biology, 14(1), 71.

184

Doonan, R., McElwee, J. J., Matthijssens, F., Walker, G. A., Houthoofd, K., Back, P., Matscheski, A., Vanfleteren, J. R., & Gems, D. (2008). Against the oxidative damage theory of aging: superoxide dismutases protect against oxidative stress but have little or no effect on life span in Caenorhabditis elegans. Genes & Development, 22(23), 3236-3241. Durai, S., & Balamurugan, K. (2017). Vibrio: Caenorhabditis elegans as a laboratory model for Vibrio infections. In: Liu, D. (Eds.), Laboratory Models for Foodborne Infections (pp. 413). CRC Press, Australia. Engelmann, I., & Pujol, N. (2010). Innate immunity in C. elegans. In Invertebrate Immunity (pp. 105-121): Springer. Ermolaeva, M. A., & Schumacher, B. (2014). Insights from the worm: the C. elegans model for innate immunity. Seminars in Immunology, 26(4), 303-309. Estes, K. A., Dunbar, T. L., Powell, J. R., Ausubel, F. M., & Troemel, E. R. (2010). bZIP transcription factor zip-2 mediates an early response to Pseudomonas aeruginosa infection in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the USA, 107(5), 2153-2158. Evans, E. A., Chen, W. C., & Tan, M. W. (2008). The DAF-2 insulin-like signaling pathway independently regulates aging and immunity in C. elegans. Aging Cell, 7(6), 879-893. Ewbank, J. (2006). Signaling in the immune response. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_signalingimmuneresponse/signalingi mmuneresponse.html. Feng, W.-H., Xue, K. S., Tang, L., Williams, P. L., & Wang, J.-S. (2017). Aflatoxin B(1)- induced developmental and DNA damage in Caenorhabditis elegans. Toxins, 9(1), 9. Galdino, A. C. M., Branquinha, M. H., Santos, A. L., & Viganor, L. (2017). Pseudomonas aeruginosa and its arsenal of proteases: weapons to battle the host. In Pathophysiological Aspects of Proteases (pp. 381-397): Springer. Galimov, E. R., Pryor, R. E., Poole, S. E., Benedetto, A., Pincus, Z., & Gems, D. (2018). Coupling of rigor mortis and intestinal necrosis during C. elegans organismal death. Cell Reports, 22(10), 2730-2741. Garsin, D. A., Sifri, C. D., Mylonakis, E., Qin, X., Singh, K. V., Murray, B. E., Calderwood, S. B., & Ausubel, F. M. (2001). A simple model host for identifying Gram-positive virulence factors. Proceedings of the National Academy of Sciences of the USA, 98(19), 10892-10897. Griffitts, J. S., Haslam, S. M., Yang, T., Garczynski, S. F., Mulloy, B., Morris, H., Cremer, P. S., Dell, A., Adang, M. J., & Aroian, R. V. (2005). Glycolipids as receptors for Bacillus thuringiensis crystal toxin. Science, 307(5711), 922-925. Guzman, L. M., Belin, D., Carson, M. J., & Beckwith, J. (1995). Tight regulation, modulation, and high-level expression by vectors containing the arabinose PBAD promoter. Journal of Bacteriology, 177(14), 4121-4130. Haskins, K. A., Russell, J. F., Gaddis, N., Dressman, H. K., & Aballay, A. (2008). Unfolded protein response genes regulated by CED-1 are required for Caenorhabditis elegans innate immunity. Developmental Cell, 15(1), 87-97. Hasshoff, M., Böhnisch, C., Tonn, D., Hasert, B., & Schulenburg, H. (2007). The role of Caenorhabditis elegans insulin-like signaling in the behavioral avoidance of pathogenic Bacillus thuringiensis. The FASEB Journal, 21(8), 1801-1812.

185

Hodgkin, J., Kuwabara, P. E., & Corneliussen, B. (2000). A novel bacterial pathogen, Microbacterium nematophilum, induces morphological change in the nematode C. elegans. Current Biology, 10(24), 1615-1618. Honda, Y., & Honda, S. (1999). The daf-2 gene network for longevity regulates oxidative stress resistance and Mn-superoxide dismutase gene expression in Caenorhabditis elegans. The FASEB Journal, 13(11), 1385-1393. Hoogewijs, D., Houthoofd, K., Matthijssens, F., Vandesompele, J., & Vanfleteren, J. R. (2008). Selection and validation of a set of reliable reference genes for quantitative sod gene expression analysis in C. elegans. BMC Molecular Biology, 9(1), 9. Hu, J., Cai, W., Wang, C., Du, X., Lin, J., & Cai, J. (2018). Purification and characterization of alkaline lipase production by Pseudomonas aeruginosa HFE733 and application for biodegradation in food wastewater treatment. Biotechnology & Biotechnological Equipment, 32(3), 583-590. Huffman, D. L., Bischof, L. J., Griffitts, J. S., & Aroian, R. V. (2004). Pore worms: using Caenorhabditis elegans to study how bacterial toxins interact with their target host. International Journal of Medical Microbiology, 293(7-8), 599-607. Iacovache, I., van der Goot, F. G., & Pernot, L. (2008). Pore formation: An ancient yet complex form of attack. Biochimica et Biophysica Acta (BBA) - Biomembranes, 1778(7), 1611-1623. Iatsenko, I., Boichenko, I., & Sommer, R. J. (2014). Bacillus thuringiensis DB27 produces two novel protoxins, Cry21Fa1 and Cry21Ha1, which act synergistically against nematodes. Applied and Environmental Microbiology, 80(10), 3266-3275. Irazoqui, J. E., Ng, A., Xavier, R. J., & Ausubel, F. M. (2008). Role for β-catenin and HOX transcription factors in Caenorhabditis elegans and mammalian host epithelial-pathogen interactions. Proceedings of the National Academy of Sciences of the USA, 105(45), 17469-17474. Irazoqui, J. E., Urbach, J. M., & Ausubel, F. M. (2010). Evolution of host innate defence: insights from Caenorhabditis elegans and primitive invertebrates. Nature Reviews Immunology, 10(1), 47-58. Jain, C., Yun, M., Politz, S. M., & Rao, R. P. (2009). A pathogenesis assay using Saccharomyces cerevisiae and Caenorhabditis elegans reveals novel roles for yeast AP-1, Yap1, and host dual oxidase BLI-3 in fungal pathogenesis. Eukaryotic Cell, 8(8), 1218-1227. JebaMercy, G., & Balamurugan, K. (2012). Effects of sequential infections of Caenorhabditis elegans with Staphylococcus aureus and Proteus mirabilis. Microbiology and Immunology, 56(12), 825-835. Jia, K., Thomas, C., Akbar, M., Sun, Q., Adams-Huet, B., Gilpin, C., & Levine, B. (2009). Autophagy genes protect against Salmonella typhimurium infection and mediate insulin signaling-regulated pathogen resistance. Proceedings of the National Academy of Sciences of the USA, 106(34), 14564-14569. Karlen, Y., McNair, A., Perseguers, S., Mazza, C., & Mermod, N. (2007). Statistical significance of quantitative PCR. BMC Bioinformatics, 8, 131-131. Kim, D. H. & Ewbank, J. J. (2015). Signaling in the innate immune response. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at https://www.ncbi.nlm.nih.gov/books/NBK19673/. Kourtis, N., Nikoletopoulou, V., & Tavernarakis, N. (2012). Small heat-shock proteins protect from heat-stroke-associated neurodegeneration. Nature, 490, 213.

186

Kurm, V., van der Putten, W. H., Weidner, S., Geisen, S., Snoek, B. L., Bakx, T., & Hol, W. H. G. (2019). Competition and predation as possible causes of bacterial rarity. Environmental Microbiology, 21(4), 1356-1368. Lamine, B., Lamine, B., & Bouziane, A. (2012). Optimisation of the chitinase production by Serratia marcescens DSM 30121T and biological control of locusts. Journal of Biotechnology and Biomaterials, 2(3), 3-7. Li, W., Wang, D., & Wang, D. (2018). Regulation of the response of Caenorhabditis elegans to simulated microgravity by p38 Mitogen-Activated Protein Kinase signaling. Scientific Reports, 8(1), 857. Li, X. Q., Wei, J. Z., Tan, A., & Aroian, R. V. (2007). Resistance to root-knot nematode in tomato roots expressing a nematicidal Bacillus thuringiensis crystal protein. Plant Biotechnology Journal, 5(4), 455-464. Los, F. C., Kao, C. Y., Smitham, J., McDonald, K. L., Ha, C., Peixoto, C. A., & Aroian, R. V. (2011). RAB-5- and RAB-11-dependent vesicle-trafficking pathways are required for plasma membrane repair after attack by bacterial pore-forming toxin. Cell Host & Microbe, 9(2), 147-157. Los, F. C. O., Randis, T. M., Aroian, R. V., & Ratner, A. J. (2013). Role of pore-forming toxins in bacterial infectious diseases. Microbiology and Molecular Biology Reviews, 77(2), 173. Luo, H., Xiong, J., Zhou, Q., Xia, L., & Yu, Z. (2013). The effects of Bacillus thuringiensis Cry6A on the survival, growth, reproduction, locomotion, and behavioral response of Caenorhabditis elegans. Applied Microbiology and Biotechnology, 97(23), 10135-10142. Mallo, G. V., Kurz, C. L., Couillault, C., Pujol, N., Granjeaud, S., Kohara, Y., & Ewbank, J. J. (2002). Inducible antibacterial defense system in C. elegans. Current Biology, 12(14), 1209-1214. Marroquin, L. D., Elyassnia, D., Griffitts, J. S., Feitelson, J. S., & Aroian, R. V. (2000). Bacillus thuringiensis (Bt) toxin susceptibility and isolation of resistance mutants in the nematode Caenorhabditis elegans. Genetics, 155(4), 1693-1699. Matz, C., & Kjelleberg, S. (2005). Off the hook – how bacteria survive protozoan grazing. Trends in Microbiology, 13(7), 302-307. Meisel, J. D., & Kim, D. H. (2014). Behavioral avoidance of pathogenic bacteria by Caenorhabditis elegans. Trends in Immunology, 35(10), 465-470. Mendoza De Gives, P. M., Davies, K. G., Clark, S. J., & Behnke, J. M. (1999). Predatory behaviour of trapping fungi against srf mutants of Caenorhabditis elegans and different plant and animal parasitic nematodes. Parasitology, 119 (1), 95-104. Mosser, T., Matic, I., & Leroy, M. (2011). Bacterium-induced internal egg hatching frequency is predictive of life span in Caenorhabditis elegans populations. Applied and Environmental Microbiology., 77(22), 8189-8192. Mukhopadhyay, A., Oh, S. W., & Tissenbaum, H. A. (2006). Worming pathways to and from DAF-16/FOXO. Experimental Gerontology, 41(10), 928-934. Murphy, C. T., McCarroll, S. A., Bargmann, C. I., Fraser, A., Kamath, R. S., Ahringer, J., Li, H., & Kenyon, C. (2003). Genes that act downstream of DAF-16 to influence the lifespan of Caenorhabditis elegans. Nature, 424(6946), 277-283. Nandi, M., Selin, C., Brassinga, A. K. C., Belmonte, M. F., Fernando, W. G. D., Loewen, P. C., & de Kievit, T. R. (2015). Pyrrolnitrin and hydrogen cyanide production by Pseudomonas chlororaphis strain PA23 exhibits nematicidal and repellent activity against Caenorhabditis elegans. PloS One, 10(4), e0123184.

187

Navarro, L., Alto, N. M., & Dixon, J. E. (2005). Functions of the Yersinia effector proteins in inhibiting host immune responses. Current Opinion in Microbiology, 8(1), 21-27. Nicholas, H. R., & Hodgkin, J. (2004). The ERK MAP Kinase cascade mediates tail swelling and a protective response to rectal infection in C. elegans. Current Biology, 14(14), 1256-1261. Nikoletopoulou, V., & Tavernarakis, N. (2014). Chapter Six - Necrotic cell death in Caenorhabditis elegans. In A. Ashkenazi, J. A. Wells, & J. Yuan (Eds.), Methods in Enzymology (Vol. 545, pp. 127-155): Academic Press. Niu, Q., Huang, X., Zhang, L., Xu, J., Yang, D., Wei, K., Niu, X., An, Z., Bennett, J. W., Zou, C., Yang, J., & Zhang, K.-Q. (2010). A Trojan horse mechanism of bacterial pathogenesis against nematodes. Proceedings of the National Academy of Sciences of the USA, 107(38), 16631. Nursyam, H., Prihanto, A., Warasari, N., Saadah, M., Masrifa, R., Nabila, N., Istiqfarin, N., & Siddiq, I. (2018). The isolation and identification of endophytic bacteria from mangrove (Sonneratia alba) that produces gelatinase. Paper presented at the IOP Conference Series: Earth and Environmental Science. O’Rourke, D., Baban, D., Demidova, M., Mott, R., & Hodgkin, J. (2006). Genomic clusters, putative pathogen recognition molecules, and antimicrobial genes are induced by infection of C. elegans with M. nematophilum. Genome research, 16(8), 1005-1016. Penesyan, A., Ballestriero, F., Daim, M., Kjelleberg, S., Thomas, T., & Egan, S. (2012). Assessing the effectiveness of functional genetic screens for the identification of bioactive metabolites. Marine Drugs, 11(1), 40-49. Peng, D., Wan, D., Cheng, C., Ye, X., & Sun, M. (2018). Nematode-specific cadherin CDH-8 acts as a receptor for Cry5B toxin in Caenorhabditis elegans. Applied Microbiology and Biotechnology, 102(8), 3663-3673. Pernthaler, J. (2005). Predation on prokaryotes in the water column and its ecological implications. Nature Reviews Microbiology, 3(7), 537-546. Pfaffl, M. W. (2001). A new mathematical model for relative quantification in real-time RT–PCR. Nucleic Acids Research, 29(9), e45-e45. Portal-Celhay, C., Bradley, E. R., & Blaser, M. J. (2012). Control of intestinal bacterial proliferation in regulation of lifespan in Caenorhabditis elegans. BMC Microbiology, 12(1), 49. Powell, J. R., Kim, D. H., & Ausubel, F. M. (2009). The G protein-coupled receptor FSHR-1 is required for the Caenorhabditis elegans innate immune response. Proceedings of the National Academy of Sciences of the USA, 106(8), 2782-2787. Pukkila-Worley, R., & Ausubel, F. M. (2012). Immune defense mechanisms in the Caenorhabditis elegans intestinal epithelium. Current Opinion in Immunology, 24(1), 3-9. Roderick, H., Tripathi, L., Babirye, A., Wang, D., Tripathi, J., Urwin, P., & Atkinson, H. J. (2012). Generation of transgenic plantain (Musa spp.) with resistance to plant pathogenic nematodes. Molecular Plant Pathology, 13(8), 842-851. Ruijter, J., Ramakers, C., Hoogaars, W., Karlen, Y., Bakker, O., Van den Hoff, M., & Moorman, A. (2009). Amplification efficiency: linking baseline and bias in the analysis of quantitative PCR data. Nucleic Acids Research, 37(6), e45-e45. Samad, M. Y. A., Razak, C. N. A., Salleh, A. B., Zin Wan Yunus, W. M., Ampon, K., & Basri, M. (1989). A plate assay for primary screening of lipase activity. Journal of Microbiological Methods, 9(1), 51-56.

188

Sarker, K. N., Das Mohapatra, P. K., & Dutta, S. (2019). Use of low cost natural resources for enhanced chitinase production and optimization using CCD and RSM: a new initiative for bio-control of plant pathogen. Indian Phytopathology, 72(2), 281- 300. Schulenburg, H., & Boehnisch, C. (2008). Diversification and adaptive sequence evolution of Caenorhabditis lysozymes (Nematoda: Rhabditidae). BMC Evolutionary Biology, 8(1), 114. Seidman, C. E., Struhl, K., Sheen, J., & Jessen, T. (1997). Introduction of plasmid DNA into cells. Current Protocols in Molecular Biology, 37(1), 1.8. 1-1.8. 10. Sellegounder, D., Pandian, S., & Balamurugan, K. (2011). Changes in Caenorhabditis elegans exposed to Vibrio parahaemolyticus. Journal of Microbiology and Biotechnology, 21, 1026-1035. Sem, X., & Rhen, M. (2012). Pathogenicity of Salmonella enterica in Caenorhabditis elegans relies on disseminated oxidative stress in the infected host. PloS One, 7(9), e45417-e45417. Senchuk, M. M., Dues, D. J., Schaar, C. E., Johnson, B. K., Madaj, Z. B., Bowman, M. J., Winn, M. E., & Van Raamsdonk, J. M. (2018). Activation of DAF-16/FOXO by reactive oxygen species contributes to longevity in long-lived mitochondrial mutants in Caenorhabditis elegans. PLOS Genetics, 14(3), e1007268. Senthil-Nathan, S. (2015). A review of biopesticides and their mode of action against insect pests. In Environmental Sustainability (pp. 49-63): Springer. Shi, J., Peng, D., Zhang, F., Ruan, L., & Sun, M. (2020). The Caenorhabditis elegans CUB-like-domain containing protein RBT-1 functions as a receptor for Bacillus thuringiensis Cry6Aa toxin. PLoS Pathogens, 16(5), e1008501. Shivers, R. P., Kooistra, T., Chu, S. W., Pagano, D. J., & Kim, D. H. (2009). Tissue- specific activities of an immune signaling module regulate physiological responses to pathogenic and nutritional bacteria in C. elegans. Cell Host & Microbe, 6(4), 321-330. Somasiri, P., Behm, C. A., Adamski, M., Wen, J., & Verma, N. K. (2020). Transcriptional response of Caenorhabditis elegans when exposed to Shigella flexneri. Genomics, 112(1), 774-781. Souza, C. d., Simpson-Louredo, L., Mota, H. F., Faria, Y. V., Cabral, F. d. O., Colodette, S. d. S., Canellas, M. E. F. C., Cucinelli, A. d. E. S., Luna, M. d. G. d., Santos, C. d. S., Moreira, L. d. O., & Mattos-Guaraldi, A. L. (2019). Virulence potential of Corynebacterium striatum towards Caenorhabditis elegans. Antonie van Leeuwenhoek, 112(9), 1331-1340. Stiernagle, T. (2006). Maintenance of C. elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_strainmaintain/strainmaintain.html. Stretton, S., Techkarnjanaruk, S., McLennan, A. M., & Goodman, A. E. (1998). Use of green fluorescent protein to tag and investigate gene expression in marine bacteria. Applied and Environmental Microbiology, 64(7), 2554-2559. Sun, S., Noorian, P., & McDougald, D. (2018). Dual role of mechanisms involved in resistance to predation by protozoa and virulence to humans. Frontiers in Microbiology, 9, 1017-1017. Szczesny, P., Iacovache, I., Muszewska, A., Ginalski, K., van der Goot, F. G., & Grynberg, M. (2011). Extending the aerolysin family: from bacteria to . PloS One, 6(6), e20349.

189

Tanaka-Hino, M., Sagasti, A., Hisamoto, N., Kawasaki, M., Nakano, S., Ninomiya-Tsuji, J., Bargmann, C. I., & Matsumoto, K. (2002). SEK-1 MAPKK mediates Ca2+ signaling to determine neuronal asymmetric development in Caenorhabditis elegans. EMBO Reports, 3(1), 56-62. Timmons, A. K., Meehan, T. L., Gartmond, T. D., & McCall, K. (2013). Use of necrotic markers in the Drosophila ovary. In Necrosis (pp. 215-228): Springer. Wang, B., Wang, H., Xiong, J., Zhou, Q., Wu, H., Xia, L., Li, L., & Yu, Z. (2017). A proteomic analysis provides novel insights into the stress responses of Caenorhabditis elegans towards nematicidal Cry6A Toxin from Bacillus thuringiensis. Scientific Reports, 7(1), 14170. Wang, Y., Branicky, R., Noë, A., & Hekimi, S. (2018). Superoxide dismutases: dual roles in controlling ROS damage and regulating ROS signaling. Journal of Cell Biology, 217(6), 1915-1928. Wani, K. A., Goswamy, D., & Irazoqui, J. E. (2020). Nervous system control of intestinal host defense in C. elegans. Current Opinion in Neurobiology, 62, 1-9. Waschkowitz, T., Rockstroh, S., & Daniel, R. (2009). Isolation and characterization of metalloproteases with a novel domain structure by construction and screening of metagenomic libraries. Applied and Environmental Microbiology, 75(8), 2506- 2516. Wei, J.-Z., Hale, K., Carta, L., Platzer, E., Wong, C., Fang, S.-C., & Aroian, R. V. (2003). Bacillus thuringiensis crystal proteins that target nematodes. Proceedings of the National Academy of Sciences of the USA, 100(5), 2760-2765. Wei, J.-Z., Siehl, D. L., Hou, Z., Rosen, B., Oral, J., Taylor, C. G., & Wu, G. (2017). An enterotoxin-like binary protein from Pseudomonas protegens with potent nematicidal activity. Applied and Environmental Microbiology, 00942-00917. Xu, K., Tavernarakis, N., & Driscoll, M. (2001). Necrotic cell death in C. elegans requires the function of calreticulin and regulators of Ca2+ release from the endoplasmic reticulum. Neuron, 31(6), 957-971. Zhang, F., Peng, D., Cheng, C., Zhou, W., Ju, S., Wan, D., Yu, Z., Shi, J., Deng, Y., & Wang, F. (2016). Bacillus thuringiensis crystal protein Cry6Aa triggers Caenorhabditis elegans necrosis pathway mediated by aspartic protease (ASP-1). PLoS Pathogens, 12(1), e1005389. Zhang, Y., Chen, D., Smith, M. A., Zhang, B., & Pan, X. (2012). Selection of reliable reference genes in Caenorhabditis elegans for analysis of nanotoxicity. PloS One, 7(3), e31849. Zhang, Y., Lu, H., & Bargmann, C. I. (2005). Pathogenic bacteria induce aversive olfactory learning in Caenorhabditis elegans. Nature, 438(7065), 179-184. Zhao, Y., Gilliat, A. F., Ziehm, M., Turmaine, M., Wang, H., Ezcurra, M., Yang, C., Phillips, G., McBay, D., & Zhang, W. B. (2017). Two forms of death in ageing Caenorhabditis elegans. Nature Communications, 8, 15458. Zou, C. G., Tu, Q., Niu, J., Ji, X. L., & Zhang, K. Q. (2013). The DAF-16/FOXO transcription factor functions as a regulator of epidermal innate immunity. PLoS Pathogens, 9(10), e1003660.

190

SUPPLEMENTARY MATERIALS

Table S3.1 qPCR efficiency and R2 value of C. elegans genes tested in this study

qPCR efficiency R2 value of Gene Equation (E) standard curve Reference gene

cdc-42 95.92% 0.9973 y = -3.420 x + 26.60 p38-MAPK pathway

pmk-1 111.10% 0.9848 y = -3.083 x + 29.16 sek-1 93.93% 0.9886 y = -3.476 x + 30.35 nsy-1 93.73% 0.9951 y = -3.482 x + 27.65 ILR siganalling pathway

daf-2 88.71% 0.9607 y = -3.63 x + 29.17 daf-16 87.05% 0.9906 y = -3.68 x + 27.84 ERK MAPK pathway ksr-1 87.74% 0.9839 y = -3.656 x + 30.50 lin-45 90.40% 0.9911 y = -3.576 x + 28.25 mek-2 82.23% 0.9882 y = -3.837 x + 32.22 Downstream genes

lys-7 84.50% 0.9956 y = -3.77 x + 27.15 lys-8 86.09% 0.9958 y = -3.71 x + 24.82 bli-3 104.30% 0.9838 y = -3.22 x + 30.28 sod-3 95.12% 0.9826 y = -3.44 x + 32.32

191

Table S3.2 Relative fluorescence quantification indicating E. coli HP1::GFP, HG8::GFP and BD24::GFP colonisation in C. elegans gastrointestinal system

One-way ANOVA Relative fluorescence quantification after 24 hours df = 2 F = 42.57 p = 0.0003 Relative fluorescence quantification after 48 hours df = 2 F = 20.36 p = 0.0021 Relative fluorescence quantification after 72 hours df = 2 F = 11.11 p = 0.0096 Relative fluorescence quantification after 96 hours df = 2 F = 31.48 p = 0.0007 Tukey's test Test strain df q p-value

Relative fluorescence quantification after 24 hours HP1::GFP vs. HG8::GFP 6 5.508 0.0188 HP1::GFP vs. BD24::GFP 6 7.49 0.0044 HG8::GFP vs. BD24::GFP 6 13 0.0002 Relative fluorescence quantification after 48 hours HP1::GFP vs. HG8::GFP 6 0.2102 0.9879 HP1::GFP vs. BD24::GFP 6 7.708 0.0038 HG8::GFP vs. BD24::GFP 6 7.918 0.0033 Relative fluorescence quantification after 72 hours HP1::GFP vs. HG8::GFP 6 1.005 0.7666 HP1::GFP vs. BD24::GFP 6 5.206 0.024 HG8::GFP vs. BD24::GFP 6 6.21 0.0109 Relative fluorescence quantification after 96 hours HP1::GFP vs. HG8::GFP 6 1.204 0.6874 HP1::GFP vs. BD24::GFP 6 9.059 0.0017 HG8::GFP vs. BD24::GFP 6 10.26 0.0008

192

Table S3.3 Bacterial colonisation assay of E. coli HP1::GFP, HG8::GFP and BD24::GFP colonisation in C. elegans gastrointestinal system

One-way ANOVA Bacterial colonisation after 6 hours df = 2 F = 2.586 p = 0.1549 Bacterial colonisation after 12 hours df = 2 F = 12.54 p = 0.0072 Bacterial colonisation after 24 hours df = 2 F = 19.87 p = 0.0023 Bacterial colonisation after 48 hours df = 2 F = 11.67 p = 0.0086 Bacterial colonisation after 72 hours df = 2 F = 21.15 p = 0.0019 Bacterial colonisation after 96 hours df = 2 F = 44.09 p = 0.0003 Tukey's test Test strain df q p-value

Bacterial colonisation after 6 hours HP1::GFP vs. HG8::GFP 6 2.096 0.3628 HP1::GFP vs. BD24::GFP 6 1.064 0.7434 HG8::GFP vs. BD24::GFP 6 3.16 0.1432 Bacterial colonisation after 12 hours HP1::GFP vs. HG8::GFP 6 4.455 0.0452 HP1::GFP vs. BD24::GFP 6 2.54 0.249 HG8::GFP vs. BD24::GFP 6 6.995 0.0062 Bacterial colonisation after 24 hours HP1::GFP vs. HG8::GFP 6 6.773 0.0072 HP1::GFP vs. BD24::GFP 6 1.635 0.5184 HG8::GFP vs. BD24::GFP 6 8.408 0.0025 Bacterial colonisation after 48 hours HP1::GFP vs. HG8::GFP 6 4.472 0.0446 HP1::GFP vs. BD24::GFP 6 2.236 0.3231 HG8::GFP vs. BD24::GFP 6 6.708 0.0076 Bacterial colonisation after 72 hours HP1::GFP vs. HG8::GFP 6 7.059 0.0059 HP1::GFP vs. BD24::GFP 6 1.577 0.5402 HG8::GFP vs. BD24::GFP 6 8.636 0.0021 Bacterial colonisation after 96 hours HP1::GFP vs. HG8::GFP 6 6.471 0.009 HP1::GFP vs. BD24::GFP 6 6.807 0.0071 HG8::GFP vs. BD24::GFP 6 13.28 0.0002

193

Table S3.4 Morphological changes on C. elegans resulted from E. coli HP1::GFP and HG8::GFP bacterial colonisation

Two-way ANOVA Average nematode length df = 2 F = 179.4 p < 0.0001 Nematode with pharynx distortion df = 2 F = 14.73 p = 0.0048 Nematode with internal hatching df = 2 F = 17.36 p = 0.0032 Nematode with internal organ damage df = 2 F = 442.1 p < 0.0001 Nematode with dar df = 2 F = 58.84 p = 0.0001 Tukey's pairwise comparison test among treatments Test strain df q p-value

Average nematode length 24 hours comparison

HP1::GFP vs. HG8::GFP 2.437 0.2159 0.9873 HP1::GFP vs. BD24::GFP 2.468 7.214 0.0441 HG8::GFP vs. BD24::GFP 3.995 16.41 0.0007 48 hours comparison HP1::GFP vs. HG8::GFP 2.721 22.05 0.002 HP1::GFP vs. BD24::GFP 3.922 22.2 0.0003 HG8::GFP vs. BD24::GFP 2.934 34.6 0.0004 72 hours comparison HP1::GFP vs. HG8::GFP 2.116 17.99 0.0091 HP1::GFP vs. BD24::GFP 3.993 9.883 0.0049 HG8::GFP vs. BD24::GFP 2.126 32.81 0.0021 96 hours comparison HP1::GFP vs. HG8::GFP 3.881 13.84 0.0016 HP1::GFP vs. BD24::GFP 3.992 7.566 0.0129 HG8::GFP vs. BD24::GFP 3.816 21.64 0.0003 Nematode with pharynx distortion 24 hours comparison HP1::GFP vs. HG8::GFP 3.469 0.000 > 0.9999 HP1::GFP vs. BD24::GFP 2.000 5.345 0.1133 HG8::GFP vs. BD24::GFP 2.000 3.536 0.2259 48 hours comparison HP1::GFP vs. HG8::GFP 3.928 1.627 0.5384 HP1::GFP vs. BD24::GFP 2.000 4.950 0.1297 HG8::GFP vs. BD24::GFP 2.000 6.481 0.08 72 hours comparison HP1::GFP vs. HG8::GFP 3.124 0.000 > 0.9999 HP1::GFP vs. BD24::GFP 2.000 7.071 0.0681 HG8::GFP vs. BD24::GFP 2.000 3.922 0.1917 96 hours comparison HP1::GFP vs. HG8::GFP 4.000 8.000 0.0105 HP1::GFP vs. BD24::GFP 2.000 22.630 0.007 HG8::GFP vs. BD24::GFP 2.000 11.310 0.0277 194

Nematode with internal hatching 24 hours comparison HP1::GFP vs. HG8::GFP 2.000 4.950 0.1297 HP1::GFP vs. BD24::GFP +infinity <0.0001 HG8::GFP vs. BD24::GFP 2.000 7.071 0.0681 48 hours comparison HP1::GFP vs. HG8::GFP 3.920 0.000 > 0.9999 HP1::GFP vs. BD24::GFP 2.000 2.449 0.3779 HG8::GFP vs. BD24::GFP 2.000 2.828 0.3131 72 hours comparison HP1::GFP vs. HG8::GFP 2.941 1.897 0.4694 HP1::GFP vs. BD24::GFP 2.000 3.536 0.2259 HG8::GFP vs. BD24::GFP 2.000 11.310 0.0277 96 hours comparison HP1::GFP vs. HG8::GFP 3.920 8.552 0.0088 HP1::GFP vs. BD24::GFP 2.000 14.700 0.0166 HG8::GFP vs. BD24::GFP 2.000 1.414 0.6452 Nematode with internal organ damage 24 hours comparison HP1::GFP vs. HG8::GFP 2.941 6.957 0.0338 HP1::GFP vs. BD24::GFP 2.000 2.828 0.3131 HG8::GFP vs. BD24::GFP 2.000 9.192 0.0414 48 hours comparison HP1::GFP vs. HG8::GFP 2.941 16.440 0.0031 HP1::GFP vs. BD24::GFP 2.000 2.828 0.3131 HG8::GFP vs. BD24::GFP 2.000 19.800 0.0092 72 hours comparison HP1::GFP vs. HG8::GFP 2.000 36.770 0.0019 HP1::GFP vs. BD24::GFP 2.000 5.657 0.1024 HG8::GFP vs. BD24::GFP +infinity < 0.0001 96 hours comparison HP1::GFP vs. HG8::GFP 2.000 31.110 0.0033 HP1::GFP vs. BD24::GFP 2.000 11.310 0.0277 HG8::GFP vs. BD24::GFP +infinity < 0.0001 Nematode with dar 24 hours comparison HP1::GFP vs. HG8::GFP 2.000 6.124 0.0887 HP1::GFP vs. BD24::GFP 2.000 6.124 0.0887 HG8::GFP vs. BD24::GFP 48 hours comparison HP1::GFP vs. HG8::GFP 2.000 11.310 0.0277 HP1::GFP vs. BD24::GFP 2.000 11.310 0.0277 HG8::GFP vs. BD24::GFP

195

72 hours comparison HP1::GFP vs. HG8::GFP 2.000 9.192 0.0414 HP1::GFP vs. BD24::GFP 2.000 9.192 0.0414 HG8::GFP vs. BD24::GFP 96 hours comparison HP1::GFP vs. HG8::GFP 2.000 19.800 0.0092 HP1::GFP vs. BD24::GFP 2.000 19.800 0.0092 HG8::GFP vs. BD24::GFP

Table S3.5 Morphological changes on C. elegans resulted from E. coli HP1 and HG8 total protein fraction exposure

Two-way ANOVA Average nematode length df = 1 F = 10.47 p = 0.0071 Nematode with internal organ damage df = 1 F = 364.5 p < 0.0001 Nematode with pharynx distortion df = 1 F = 0.2500 p = 0.6261 Nematode with vacuole df = 1 F = 25.00 p = 0.0003 Nematode with internal hatching df = 1 F = 3.125 p = 0.1025 Tukey's pairwise comparison test among treatments Test strain df q p-value Average nematode length

Soluble protein HP1 vs. HG8 12 0.1436 0.9943 HP1 vs. BD24 12 1.186 0.6872 HG8 vs. BD24 12 1.042 0.7469 Insoluble protein HP1 vs. HG8 12 2.766 0.1658 HP1 vs. BD24 12 5.465 0.0059 HG8 vs. BD24 12 8.232 0.0002 Nematode with internal organ damage Soluble protein HP1 vs. HG8 12 3.464 0.0731 HP1 vs. BD24 12 10.39 <0.0001 HG8 vs. BD24 12 6.928 0.001 Insoluble protein HP1 vs. HG8 12 39.84 <0.0001 HP1 vs. BD24 12 12.12 <0.0001 HG8 vs. BD24 12 51.96 <0.0001 Nematode with pharynx distortion Soluble protein HP1 vs. HG8 12 4.287 0.0262 HP1 vs. BD24 12 5.511 0.0056 196

HG8 vs. BD24 12 1.225 0.6709 Insoluble protein HP1 vs. HG8 12 5.511 0.0056 HP1 vs. BD24 12 5.511 0.0056 HG8 vs. BD24 12 0 >0.9999 Nematode with vacuole Soluble protein HP1 vs. HG8 12 7.348 0.0006 HP1 vs. BD24 12 13.88 <0.0001 HG8 vs. BD24 12 6.532 0.0016 Insoluble protein HP1 vs. HG8 12 8.165 0.0002 HP1 vs. BD24 12 8.165 0.0002 HG8 vs. BD24 12 0 >0.9999 Nematode with internal hatching Soluble protein HP1 vs. HG8 12 5.196 0.0083 HP1 vs. BD24 12 13.86 <0.0001 HG8 vs. BD24 12 8.66 0.0001 Insoluble protein HP1 vs. HG8 12 12.99 <0.0001 HP1 vs. BD24 12 15.59 <0.0001 HG8 vs. BD24 12 2.598 0.1995

Table S3.6 Proportion of C. elegans with membrane loss of integrity (necrosis) resulted from E. coli HP1 and HG8 bacterial exposure

One-way ANOVA Proportion of nematode with necrosis resulted from E. coli HP1 and HG8 bacterial df = 2 F = 86.38 p < 0.0001 exposure Tukey's test Test strain df q p-value

HP1 vs. HG8 6 2.915 0.1786 HP1 vs. BD24 6 14.44 0.0001 HG8 vs. BD24 6 17.36 <0.0001

197

Table S3.7 Innate and associative learned avoidance behaviour of C. elegans against E. coli HP1 and HG8

Unpaired t-test (two-tailed) Test strain F df t p-value

HP1 vs BD24 naïve vs trained nematodes 11.03 4 3.856 0.0182

HG8 vs A1A naïve vs trained nematodes 6.685 4 10.23 0.0005

Table S3.8 Differential expression of innate immunity genes in C. elegans resulted from E. coli HP1 exposure

Two-way ANOVA ILR signalling pathway df = 1 F = 2.777 p = 0.1709 p38-MAPK pathway df = 2 F = 1.381 p = 0.3212 ERK MAPK pathway df = 2 F = 1.340 p = 0.3303 Downstream genes df = 3 F = 14.64 p = 0.0013 Sidak's test 6 hours vs 24 hours df t p-value

ILR siganalling pathway daf-2 4 3.596 0.0451 daf-16 4 1.24 0.4857

p38-MAPK pathway pmk-1 6 0.5393 0.9403 nsy-1 6 0.5883 0.9247 sek-1 6 2.599 0.1173

ERK MAPK pathway ksr-1 6 1.279 0.5752 lin-45 6 1.205 0.6168 mek-2 6 0.7622 0.8551

Downstream genes bli-3 8 1.044 0.7949 sod-3 8 4.074 0.0142 lys-7 8 1.859 0.3442 lys-8 8 4.838 0.0052

198

Soluble protein Insoluble protein

A B

HG8

C D

HP1

E F

BD24

Figure S3.1 Example of size reduction observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones. Size reduction was detected on nematodes assayed with insoluble proteins of E. coli HG8 (B) and E. coli HP1 (D). No size reduction was observed on nematodes exposed to the soluble protein of E. coli HG8 (A) and E. coli HP1 (C) and the soluble and insoluble proteins of the non-toxic control E. coli BD24 (E, F). Nematode images were acquired using the DIC filter under 10x magnification. Scale bars indicate 100 µm.

199

Soluble protein Insoluble protein

A B

HG8

C D

HP1

E F

BD24

Figure S3.2 Example of internal organ damage observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones. Internal organ (intestine and/or gonad) damage was detected on nematodes assayed with soluble and insoluble protein fractions of E. coli HG8 (A, B) and E. coli HP1 (C, D). In contrast, no internal organ damage was observed on nematodes exposed to the soluble and insoluble proteins of the non-toxic control E. coli BD24. Nematode images were acquired using the DIC filter under 40x magnification. Scale bars indicate 20 µm.

200

Soluble protein Insoluble protein

A B

HG8

C D

HP1

E F

BD24

Figure S3.3 Example of pharynx distortion observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones. Pharynx distortion was detected on nematodes assayed with soluble protein of E. coli HG8 (A) and the soluble and insoluble proteins of E. coli HP1 (C, D). A normal pharynx was observed on nematodes exposed to the insoluble protein of HG8 and the soluble and insoluble proteins of the non-toxic control E. coli BD24. Nematode images were acquired using the DIC filter under 40x magnification. Scale bars indicate 20 µm.

201

Soluble protein Insoluble protein

A B

HG8

C D

HP1

E F

BD24

Figure S3.4 Example of vacuole formation observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones. Vacuole formations (indicated by white arrow) were detected on nematodes assayed with soluble protein fraction of E. coli HG8 (A) and soluble and insoluble proteins of E. coli HP1 (C, D). In contrast, no vacuole formation was observed on nematodes exposed to the insoluble protein of E. coli HG8 (B) and soluble and insoluble proteins of the non-toxic control E. coli BD24 (E, F). Nematode images were acquired using the DIC filter under 40x magnification. Scale bars indicate 20 µm.

202

Soluble protein Insoluble protein

A B

HG8

C D

HP1

E F

BD24

Figure S3.5 Example of internal hatching observed on C. elegans upon exposure to soluble and insoluble protein fractions of different E. coli clones. Internal hatching was detected on nematodes assayed with soluble and insoluble protein fractions of E. coli HG8 (A, B) and E. coli HP1 (C, D). In contrast, intact eggs and no internal hatching was observed on nematodes exposed to the soluble and insoluble proteins of the non-toxic control E. coli BD24. Nematode images were acquired using the DIC filter under 10x (B) or 40x (A, C-F) magnifications. Scale bars indicate 100 µm (B) or 20 µm (A, C-F).

203

Figure S3.6 Melt curve of the amplified qPCR product of C. elegans innate immunity genes. Single peak observed on each graph indicates a specific qPCR amplification of the gene of interest. Blue and black lines observed on each graph indicate the melt curves of qPCR negative controls NRT (reaction tubes containing the raw RNA) or NTC (cDNA was substituted with the molecular grade water) respectively.

204

Figure S3.6 (Continue) Melt curve of the amplified qPCR product of C. elegans innate immunity genes. Single peak observed on each graph indicates a specific qPCR amplification of the gene of interest. Blue and black lines observed on each graph indicate the melt curves of qPCR negative controls NRT (reaction tubes containing the raw RNA) or NTC (cDNA was substituted with the molecular grade water) respectively.

205

CHAPTER 4

The presence of a defined gut microbiota in Caenorhabditis elegans alleviates toxicity of the antinematode protein Nkp-1

206

4.1 INTRODUCTION

Members of the gut microbiota of animals are crucial for host health and fitness not only because of their essential role in digestion (Martinez-Guryn et al., 2018) but also because they support host growth, tissue development, metabolism and behaviour (Sampson and Mazmanian, 2015; Tilg and Moschen, 2015; Thursby and Juge, 2017; Cani, 2018). Furthermore, gut microbiota can provide protection against environmental stressors including toxic pollutants (Claus et al., 2016), excessive temperatures (Raza et al., 2020) and colonisation by pathogenic microorganisms (Rolhion and Chassaing, 2016). Protection from pathogens can be achieved via multiple mechanisms including (i) growth inhibition, either through competition for available nutrients or via the secretion of antagonistic compounds (Kamada et al., 2012; Garcia-Gutierrez et al., 2019), (ii) the host intestinal immune system and protective inflammatory response, through the production of antimicrobial peptides and pattern recognition receptor signalling (Rakoff-Nahoum et al., 2004; Blander et al., 2017) and (iii) secretion of short-chain fatty acids by the gut microbiota that enhances mucus production by the host intestinal cells thereby conserving the integrity of the gut barrier (Willemsen et al., 2003; Silva et al., 2020). These important gut microbiota functions are also key to the success of bacterivorous organisms, including the model nematode C. elegans (Samuel et al., 2016; Zimmermann et al., 2020).

In nature, C. elegans is in close contact with highly diverse soil microorganisms. However, the nematode harbours a unique gut microbiota (Samuel et al., 2016; Jiang and Wang, 2018). Recent studies have revealed that the C. elegans microbiota resulting from the soil environment is composed predominantly of bacteria belonging to the families Enterobacteriaceae, Xanthomonadaceae (genus Stenotrophomonas), Pseudomonadaceae (genus Pseudomonas), Burkholderiaceae (genus Achromobacter), Sphingobacteriaceae, Acetobacteraceae and the genera Sphingomonas (family Sphingomonadaceae) and Ochrobactrum (family Brucellaceae) (Berg et al., 2016; Dirksen et al., 2016; Samuel et al., 2016). In-vitro studies have observed this complex C. elegans gut microbiota can protect the nematode against microbial infection either by boosting the host innate immunity or via the production of a cyclic lipopeptide that inhibits the growth of pathogens such as Bacillus thuringiensis (Montalvo-Katz et al., 2013; Kissoyan et al., 2019). Moreover, specific gut microbes are known to directly 207 support C. elegans growth through the production of micronutrients (e.g. vitamin B12) and by improving fatty acid, amino acid and energy metabolism of the nematodes (Yang et al., 2019; Zimmermann et al., 2020). Interestingly, some of these beneficial gut microbes including members of the genera Achromobacter sp. and Ochrobactrum, are also associated with free-living or entomopathogenic nematodes i.e. Acrobeloides maximus (Baquiran et al., 2013) Oscheius chongmingensis Tumian (Fu and Liu, 2019) and Rhabditis sp. (Deepa et al., 2015). These observations suggest an important evolutionary relationship between nematodes and specific microbial taxa which is widespread rather than exclusively associated with a particular nematode strain (Baquiran et al., 2013).

Despite the growing recognition of the importance of gut microbiota, the establishment of C. elegans as a model organism in the 1970 has resulted in the predominant use of laboratory-reared monoxenic C. elegans N2 populations (Brenner, 1974; Stiernagle, 2006; Samuel et al., 2016). Although the value of the C. elegans N2 strain cannot be underestimated (Sulston, 2019) (see also Chapter 1, section 1.4), the role of a more complex C. elegans gut microbiota should not be disregarded (Zimmermann et al., 2020). Indeed, performing toxicity assays against nematodes harbouring “natural” gut microbiota will presumably give more informative results, particularly for antinematode drug development. The Pseudoalteromonas tunicata nematode killing protein Nkp-1 was characterised in Chapter 2 and its mode of action as a necrosis causing pore forming toxin was proposed in Chapter 3. However, all of these experiments were performed in monoxenic C. elegans N2 which feed only on single strain E. coli OP50, therefore, investigating the effect of gut microbiota on the N2 nematode survival upon exposure to the Nkp-1 will provide a better understanding of Nkp-1 toxic activity.

This research project aimed to firstly establish a native gut microbiota in C. elegans N2- laboratory cultures and subsequently determine what impact this microbiota has on the sensitivity of C. elegans to Nkp-1. I hypothesise that the presence of a more diverse gut microbiota will enhance the nematode survival by reducing toxic bacterial colonisation

208 and morphological changes related to physical damages to the host upon exposure to the Nkp-1 expressing bacteria; E. coli HP1 and HG8.

4.2 MATERIALS AND METHODS

4.2.1 Bacterial strains and culture conditions

All bacterial strains and vectors used in this study are listed in Table 4.1. Unless otherwise stated, all of the bacterial strains were routinely grown in Luria Bertani broth (LB10, Appendix II) at 37°C (200 rpm) and Nematode Growth Media (NGM, appendix II) (Brenner, 1974). Solid media was prepared with the addition of 1.5% (w/v) of agar (Oxoid, Australia). All bacterial strains were kept in 30% (v/v) glycerol at -80°C. Where required, L-(+)-Arabinose (0.2% w/v) and antibiotics such as chloramphenicol (12.5 µg/mL), ampicillin (50 µg/mL) and kanamycin (50 µg/mL) were added into the media. In every experiment which involved Caenorhabditis elegans exposure to the E. coli HP1 or HP1::GFP, HG8 or HG8::GFP and the non-toxic control BD24 or BD24::GFP bacterial lawn, assay plates were initially prepared by inoculating 30 µL of the bacterial overnight culture onto the LB10 agar with antibiotics and induced with L-(+)-Arabinose for 96 hours at 25°C.

209

Table 4.1 Bacterial strains and vectors used in this study

Reference or Strain or vector Relevant characteristic or genotype source Strains

E. coli EPI300-T1R F-mcrA ∆ (mrrhsdRMSmcrBC) Epicentre ɸ80dlacZ∆M15∆lacX74 recA1 endA1 araD139 ∆(ara, leu) 7697 galU galK λ- rpsL nupG trfA tonA dhfr EPI300 F– λ– mcrA Δ(mrr-hsdRMS-mcrBC) Epicentre Φ80dlacZΔM15 Δ(lac)X74 recA1 endA1 araD139 Δ(ara, leu)7697 galU galK rpsL (StrR) nupG' trfA dhfr OP50 Uracil auxotroph (Brenner, 1974) HP1 EPI300 transformed with pBAD24hp1 This study BD24 EPI300 transformed with empty pBAD24 This study vector HG8 EPI300-T1R transformed with (Ballestriero et al., pCC1FOS™ :: ZP_ 01132230.1 to ZP_ 2010) 01132246.1 HP1::GFP E. coli HP1 transformed with p519ngfp This study plasmid HG8::GFP E. coli HG8 strain transformed with This study p519ngfp plasmid BD24::GFP E. coli BD24 strain transformed with This study p519ngfp plasmid

Vectors

pCC1FOS™ :: ZP_ Fosmid backbone for genomic library of (Burke et al., 2007; 01132230.1 to ZP_ Pseudoalteromonas tunicata D2 carrying Ballestriero et al., 01132246.1 a wild type D2 insert (13.8 kb) expressing 2010) putative antinematode activity, Cmr

pBAD24 F-, Δ(argF-lac)169, (Guzman et al., φ80dlacZ58(M15), glnX44(AS), λ- 1995) , rfbC1, gyrA96(NalR), recA1, endA1, spo T1, thiE1, hsdR17, pBAD24 pBAD24hp1 P. tunicata D2 wild type gene hp1 (NCBI This study Accession: ZP_01132246.1) cloned downstream the pBAD promoter, Ampr p519ngfp High-copy-number plasmid with (Stretton et al., constitutive GFP expression; Kmr 1998)

a Inducible expression with the presence of L-(+)-Arabinose 0.2% (w/v)

210

4.2.2 Maintenance and manipulation of C. elegans

Unless otherwise indicated, C. elegans N2 strain Bristol (Brenner, 1974) were maintained on NGM plates seeded with E. coli OP50. Nematode synchronisation was performed according to the standard bleaching procedure (Stiernagle, 2006) with modifications as described by Ballestriero et al. (2010) in section 2.2.2 (Chapter 2).

4.2.3. Enriched soil preparation

Garden soil was sampled from the School of Biological, Earth and Environmental Science (BEES), University of New South Wales (UNSW), Sydney, Australia glasshouse in September 2019. To support C. elegans development with sufficient microbial growth, the soil was supplemented with over-ripened and diced fruits (banana, apple and orange) at a ratio of 2:1 (w/w) (200 g soil : 100 g fruit) and the fruits were allowed to decompose in the soil at 25°C for 14 days as previously described (Berg et al., 2016) (Figure 4.1A). To ensure the enriched soil was devoid of native nematodes, soil samples (5g in 50 mL glass beaker) were autoclaved at 121°C for 15 minutes and allowed to cool at ambient room temperature. To allow for the native soil microbiota to be restored, prior to sterilisation, microorganisms were harvested from the enriched soil by homogenizing the soil in M9 buffer (1:3 w/v) (10 g soil : 30 mL M9 buffer) followed by centrifugation at 1800 rpm for 3 minutes (Figure 4.1A). The resulting soil supernatant containing native microorganisms was then added to the autoclaved soil samples at a ratio of 1:5 (w/v) (1 mL soil supernatant : 5 g soil) and samples were incubated for 24 hours at 25°C to allow the soil microbes to re-establish (Figure 4.1A).

211

4.2.4. Establishment of C. elegans N2 with an undefined gut microbiota from enriched soil

C. elegans N2 with an undefined gut microbiota (N2_UGM) were established from the enriched soil samples according to the procedure of Berg et al. (2016) with some modifications (Figure 4.1). Briefly, synchronised L1 stage C. elegans larvae were obtained as described in section 2.2.2 (Chapter 2) and placed on fresh NGM plates without E. coli OP50 at 25°C. The L1 larvae were washed using M9 buffer (Appendix II) and introduced into the enriched soils (800 to 900 worms/ beaker) (Figure 4.1A). C. elegans N2_UGM were raised in the soil at 25°C for three consecutive days, followed by nematode harvesting using Baermann funnels lined with two layers of muslin cloth (Figure 4.1B). The soil samples containing the N2_UGM nematodes were submerged in M9 buffer for approximately two hours, allowing the nematodes to escape from the soil and accumulated at the clipped tube (Figure 4.1B). C. elegans N2_UGM were harvested by opening the clip and pooled in a 15 mL macrocentrifuge tube. Nematodes were washed six times with M9 buffer to remove the cuticle associating bacteria and surface sterilised by placing them onto fresh NGM agar supplemented with gentamicin (100 µg/mL) for at least one hour (Figure 4.1B). Next, C. elegans N2_UGM were washed using M9 buffer and subsamples of nematodes subjected to either the nematode killing assay described in section 4.2.7 or sampling for the cultivation and 16S rRNA gene amplicon sequencing of the gut microbiota described in section 4.2.5 and 4.2.10 respectively (Figure 4.1B). Removal of surface bacteria was assessed by inoculating the final M9 buffer wash onto the LB10 agar plates and incubating the plates for 72 hours at 25°C, a lack of bacterial growth on agar plates was considered an indicator of successful nematode surface sterilisation.

212

A B

*

Figure 4.1 Schematic diagram representing the preparation of enriched soils and establishment of C. elegans undefined gut microbiota. (A) Soil enrichment was performed through supplementation of diced fruits and were decomposed in the soil. Native nematodes of the soil were cured by autoclaving while the soil microorganisms were restored by mixing with the soil microbial supernatant. (B) C. elegans N2 with undefined gut microbiota (N2_UGM) were harvested using the Baermann funnel and surface sterilised. The N2_UGM nematodes were subjected for further analysis including nematode killing assay and isolation of gut microbiota. *The 16S rRNA gene amplicon sequencing analysis was performed on the soil samples and the N2_UGM gut bacterial communities. Figure was designed using BioRender at https://biorender.com/ 213

4.2.5 Isolation and identification of the cultivable C. elegans N2_UGM gut microbiota

Bacteria were isolated from the gut of N2_UGM nematodes following the protocol of (Berg et al., 2016) with some modifications. Briefly, ~ 200 C. elegans N2_UGM (obtained from the section 4.2.4) were transferred into a 1.5 mL microcentrifuge tube containing 200 µL of Phosphate Buffer Saline (PBS, Appendix II) and mechanically disrupted using sterilized pestle. Nematode lysate was serially diluted using PBS and spread plated onto the LB10 agar following incubation at 25°C for 48 to 72 hours. An overnight culture was grown from the single bacterial colonies in LB10 broth at 25°C and 200 rpm of shaking condition. For bacterial identification, the full length of 16S rRNA gene was amplified through PCR using the primer pair of 27 Forward (5’- AGAGTTTGATCCTGGCTCAG-3’) and 1492 Reverse (5’- GGTTACCTTGTTACGACTT-3’) (Miller et al., 2013; Chen et al., 2015). PCR reactions containing 12.5 µL EconoTaq ® PLUS 2X Master Mix (Lucigen), 10 µM of 27 Forward and 1492 Reverse primers (0.2 µM final concentration of each), 1 µL of bacterial genomic DNA as the template and Ambion® nuclease-free water up to 25 µL of reaction volume were prepared. Next, the PCR mixtures were exposed to the initial denaturation at 94°C for 2 minutes followed by 30 cycles of denaturation at 94°C for 30 seconds, annealing at 50°C for 30 seconds and an extension at 72°C for 2 minutes. A final extension step was performed at 72°C for 5 minutes and samples were hold at 4°C. The specific PCR amplification was confirmed via agarose gel electrophoresis 1% (w/v) incorporated with GelRed ® Nucleic Acid Gel Stain (Biotium) 0.01% (v/v) while the DNA sequencing and nucleotide alignment using BLAST (Altschul et al., 1997) were performed as described in section 2.2.5 (Chapter 2).

214

4.2.6 Establishment of C. elegans N2 with a defined gut microbiota

C. elegans N2 associated with a defined gut microbiota (N2_DGM) were established following a protocol of (Berg et al., 2019) with some modifications. In brief, bacterial isolates obtained in section 4.2.5 were grown for 24 hours in LB10 broth at 25°C and 200 rpm of shaking condition. Twenty microlitres of each bacterial culture (OD600nm = 2.0) were inoculated onto the surface of NGM agar plates following incubation at 25°C for 3 days (Figure 4.2). Sixty to seventy C. elegans N2 eggs (obtained as described in section 2.2.2, Chapter 2) were inoculated between the bacterial colonies on the NGM plates (see Figure 4.2). The hatched N2 nematode L1 larvae were raised on the NGM agar plates containing the bacterial colonies for three consecutive days in 25°C to develop to young adult C. elegans N2_DGM (Figure 4.2). Young monoxenic adult C. elegans N2 developed on E. coli OP50 using the same procedure were used as the control. Surface sterilisation of C. elegans N2_DGM and confirmation of bacterial removal from the nematode surface were performed as described previously in section 4.2.4. C. elegans N2_DGM were used for nematode killing assay as described in section 4.2.7 or microscopy analysis and DNA extraction for 16S rRNA gene amplicon sequencing as described in section 4.2.8 and section 4.2.10 respectively.

Gut bacterial isolates

Figure 4.2 Schematic diagram representing the establishment of C. elegans N2_DGM with a defined gut microbiota. Individual bacterial isolates obtained from C. elegans N2_UGM were inoculated onto the NGM agar plate. Sixty to seventy nematode eggs were inoculated between the bacterial colonies (indicated by the blue ‘x’) and were incubated for three days at 25°C. Figure was designed using BioRender at https://biorender.com/

215

4.2.7 Nematode killing assay

The nematode killing activity of Nkp-1 expressing strains E. coli HP1 and HG8 were tested against the following C. elegans populations; N2 (monoxenic strain feeding solely on E. coli OP50), N2_UGM and N2_DGM (30-40 nematodes/plate) in triplicate. The non-toxic E. coli BD24 strain was used as the negative control. Plate preparation and the nematode killing assays were performed as described previously in section 2.2.9 (Chapter 2). Experiments were performed in triplicate and monitored until the maximum lifespan of the nematodes exposed to the control strain (approximately 17 days). Statistical analysis on nematode survival was performed using the log-rank (Mantel-Cox) method (Mantel, 1966; Harrington, 2005) with the GraphPad Prism software 8.3.0 (GraphPad Software, La Jolla, CA, USA). A One-way ANOVA followed by Tukey’s pairwise comparison test was used to determine the difference of C. elegans survival in the nematode killing assay. Results are presented as the means ± standard error from triplicate samples. A p-value < 0.05 was considered to be significant.

4.2.8 E. coli HP1::GFP and HG8::GFP bacterial colonisation assay and microscopy imaging

C. elegans N2_UGM and N2_DGM were investigated for morphological changes after the establishment of an undefined or a defined gut microbiota respectively using the Fluorescence Microscope Olympus BX61 under the differential interference contrast (DIC) filter (see section 3.2.7, Chapter 3). After 24, 48 and 72 hours of exposure to E. coli HP1::GFP, HG8::GFP and the non-toxic control E. coli BD24::GFP, slides containing live nematodes [30 nematodes per treatment (10 nematodes / replicate)] were prepared and visualised for any morphological changes i.e. pharynx distortion, internal hatching, internal organ (gonad or intestine) damage and dar (deformed anal region, see section 3.1, Chapter 3 for a detail description) and also for bacterial colonisation using the DIC and Fluorescein Isothiocyanate (FITC) filters respectively (see section 3.2.7, Chapter 3). The observed morphological changes were recorded while the bacterial colonisation was evaluated using ImageJ (https://imagej.nih.gov/ij/). The GraphPad

216

Prism software 8.3.0 (GraphPad Software, La Jolla, CA, USA) was used to analyse the differences of bacteria fluorescence intensity using a One-way ANOVA followed by Sidak’s test and a Two-way ANOVA followed by Tukey’s tests was used to analyse the proportion of C. elegans showing morphological changes due to bacterial exposure. Results are presented as the means ± standard error from triplicate samples A p-value < 0.05 was considered to be significant

4.2.9 16S rRNA gene amplicon sequencing analysis of enriched soil and C. elegans samples

4.2.9.1 DNA isolation

Microbial DNA was extracted from five replicates of the enriched soil samples (hereafter denoted as ‘SOIL’) (obtained from section 4.2.3), C. elegans N2_UGM (hereafter denoted as ‘N2_UGM’) (obtained from section 4.2.4) and C. elegans N2_DGM (obtained from section 4.2.6) (hereafter denoted as ‘N2_DGM’). Total genomic DNA was extracted using the DNeasy Powersoil Kit (Qiagen, Germany) according to the manufacturer’s instructions. The DNA concentration was determined using the Qubit ® 3.0 fluorometer (Thermo Fisher Scientific, Australia) and DNA purity assessed based on the 260/280 and 260/230 ratio using the Nanodrop (Thermo Fisher Scientific, Australia). The amplified PCR products from the microbial genomic samples were determined for specific amplification via agarose gel electrophoresis 1% (w/v) incorporated with GelRed ® Nucleic Acid Gel Stain (Biotium) 0.01% (v/v).

4.2.9.2 16S rRNA gene amplification and sequencing

Bacterial communities from the SOIL (n=5), N2_UGM (n=5) and N2_DGM (n=5) samples were investigated by sequencing the 16S rRNA gene amplicons. The primers; 341 Forward (5’ TCGTCGGCAGCGTCAGATGTGTATAAGAGACAGCCTACGGGNGGCWGCAG 217

3’) and 758 Reverse (5’ GTCTCGTGGGCTCGGAGATGTGTATAAGAGACAGGACTACHVGGGTATCTA ATCC 3’) (Klindworth et al., 2013) were used to amplify the V3-V4 hypervariable regions of the 16S rRNA gene which corresponds to approximately 500 base pairs (bp). The Illumina adapter overhang was also included in the primer oligonucleotide sequences. PCR reaction for each of the samples was prepared in a 50 µL reaction volume comprised of Econotaq® PLUS GREEN 2X Master Mix (Lucigen) (25 µL), Ambion® nuclease-free water (15 μL), the primer pair 341F and 785R (2.5 μL of each; 10 μM of initial concentration) and DNA template (5 μL). The 16S rRNA gene amplification of the genomic samples was run under the cycling condition of initial denaturation at 94°C for 2 minutes followed by 35 cycles of denaturation at 94°C for 30 seconds, annealing at 55°C for 30 seconds and extension at 72°C for 40 seconds. A final extension step was applied at 72°C for 7 minutes. PCR amplicons were visualised and quantified using agarose gel electrophoresis 1% (w/v) incorporated with GelRed ® Nucleic Acid Gel Stain (Biotium, USA) 0.01% (v/v). Paired-end sequencing (2 x 300 bp) of the resulting 16S rRNA gene amplicons was performed at the Ramaciotti Centre for Genomics, University of New South Wales (UNSW), Sydney on an Illumina MiSeq platform as per the MiSeq System User Guide (Illumina, 2013).

4.2.10 16S rRNA gene sequencing analysis

Sequence data for the SOIL, N2_UGM and N2_DGM were processed according to Ozkan et al. (2019) and Wilkes Walburn et al. (2019). In brief, the 16S rRNA gene sequence reads were initially quality‐filtered and trimmed using TRIMMOMATIC version 0.36 (Bolger et al., 2014). USEARCH version 11.0.667 (Edgar, 2010) was used for further processing as described by Wemheuer and Wemheuer (2017) to merge and quality‐filter sequencing reads, removing reads with < 250 or > 550 nucleotides, in addition to reads with more than one ambiguous base or an expected error of more than 1 (Edgar and Flyvbjerg, 2015). Filtered sequences were denoised and clustered into unique sequences (zero-distance operational taxonomic units; zOTUs) using the UNOISE algorithm (Edgar, 2016) implemented in USEARCH. zOTU represent unique bacterial entities and roughly are equivalent to species or strains. Chimeric sequences were removed with

218

UCHIME (Edgar et al., 2011) de novo during zOTU clustering and subsequently also removed with a reference-based comparison against the GTDB v89 database (https://ftdb.ecogenomic.org). zOTUs were then taxonomically classified (i.e. assigned a likely taxonomic name) by BLASTN (Camacho et al., 2009) against the GTDB database. Sequences classified as eukaryotes (organelles) or producing no Blast hit were removed, as were zOTUs appearing with only one sequence in the entire dataset. Finally, processed sequences were mapped on zOTU sequences to calculate the count distribution and counts of each zOTU in every sample. Only zOTUs occurring in more than two samples were considered for further statistical analysis.

4.2.11 Microbial community analysis

Rarefaction curves were generated using the rarecurve function in vegan (Oksanen et al., 2018) and used to determine if a complete representation of the sample’s microbiome had been achieved given the sequencing effort. Prior to the following analysis, the number of sequences was standardised for each sample to account for different sequencing depths by randomly subsampling each sample to the lowest number of sequences counts obtained for any given sample. Bacterial alpha-diversities (i.e. zOTU richness and Shannon's diversity) were calculated in R (version 3.5.3) in the microbial samples using the rrarefy function in the vegan package for community ecology analysis (Oksanen et al., 2017). Results were analysed using an Unpaired t-test in GraphPad Prism 8.3.0 (GraphPad Software, La Jolla, CA, USA) to determine the significant differences between the different groups. A p value < 0.05 was considered to be significant.

For distance-based analysis of the bacterial communities, zOTU table was imported into the R package vegan (Oksanen et al., 2018) to compare the community structure (i.e. relative abundance data) for the SOIL and N2_UGM samples. Bray-Curtis dissimilarity distances were calculated using square-root transformed zOTU abundances and the resulting similarity matrix was visualized using non-metric, multi-dimensional scaling (nMDS). Permutational multivariate analysis of variance (PERMANOVA) (Anderson, 2001) with 999 random permutations was used to test the effect of “samples” on microbial communities in the SOIL and N2_UGM microbial samples. To check if any observed 219 statistical difference was caused by differences in data dispersion, a test for homogeneity of multivariate dispersions (PERMDISP) was performed (Anderson, 2001). A p-value < 0.05 was considered to be significant. To determine which microbial zOTUs were significantly different between the SOIL and N2_UGM microbial samples, a one-factor design was used to adjust the data to a Multivariate Generalized Linear Model (MGLM) using the multivariate package Mvabund (Wang et al., 2012). Each zOTU was treated as a variable fitted to a separate Generalized Linear Model (GLM) using a negative binomial distribution. An adjusted p-value of < 0.05 was considered to be significant. All analyses and statistics were performed under R version 3.6.3 (Team, 2019). The sequence data has been submitted to the BioSample NCBI database under accession numbers SAMN15411619 to SAMN15411633.

4.2.11.1 Detection of sequences corresponding to individual gut bacteria and the differentially abundant bacterial communities in the N2_DGM gut microbiota

To determine the presence of each of the bacterial isolate sequences (obtained from section 4.3.5), near full length 16S rRNA gene sequences of the bacterial isolates were searched against the representative sequences of zOTUs in the N2_DGM samples using BLAST 2.2.30+ (Altschul et al., 1990). The relative proportion of cultured bacterial isolates (grouped to the genus level) was then compared to the relative abundance of each genera detected in N2_DGM samples using a Two-way ANOVA test in GraphPad Prism 8.3.0 (GraphPad Software, La Jolla, CA, USA) followed by Sidak’s pairwise comparisons with a p-value < 0.05 considered to be significant.

220

4.3 RESULTS

4.3.1 Establishment of an undefined gut microbiota (UGM) in C. elegans N2.

C. elegans N2 cultures were exposed to microbially enriched soil (see section 4.2.4) to establish a near natural gut microbiota. 16S rRNA gene amplicon sequencing analysis of both the enriched soil (SOIL) and resulting gut microbiomes (C. elegans N2_UGM) revealed the gut bacterial community associating C. elegans N2_UGM had fewer zOTUs (lower richness) and was less diverse compared to the enriched SOIL environment

[Unpaired t-test; F (4, 4) = 5.045; ***; p = 0.0002 (Shannon diversity) and F (4, 4) = 4.138; ****; p < 0.0001 (Richness), Figure 4.3A and B, Supplementary Materials; Tables S4.1, S4.2 and Figure S4.1].

Figure 4.3 Comparison of Shannon diversity and richness (zOTU counts) of bacterial communities between SOIL and C. elegans N2_UGM microbial samples (A) Shannon diversity (B) bacterial richness. The difference of Shannon’s diversity and richness between the SOIL and N2_UGM samples was statistically significant (p = 0.0002 and p < 0.0001 respectively). Graphical summaries represent the individual data from five replicates of each sample.

221

With respect to beta-diversity, a comparison of the of SOIL and N2_UGM bacterial communities using Bray-Curtis dissimilarity found a significant difference in bacterial community structure and composition (Supplementary materials; Table S4.3; PERMANOVA; p = 0.004, Figure S2) between SOIL and N2_UGM. No difference was observed among replicates from each sample (Supplementary materials; Table S4.3; PERMANOVA; p = 0.363, Figure S2). Across all of the SOIL and C. elegans N2_UGM microbial samples, the total zOTUs with relative abundance greater than 1% could be classified into 5 different classes (Supplementary Materials, Figure S4.3), 19 families (Supplementary Materials, Figure S4.4) and 20 genera (Figure 4.4).

The majority of sequences from SOIL and N2_UGM samples were assigned to either the class Gammaproteobacteria, , , Bacteroidia or (Supplementary Materials, Figure S4.3). Overall, the class Gammaproteobacteria were enriched in C. elegans N2_UGM (representing 35%) compared to SOIL (30.9%). Similarly, sequences belonging to Bacteroidia were enriched in the N2_UGM microbial communities (10.83%) compared to the SOIL samples (5.37%). At the genus level, sequences assigned to the genera Xanthomonas (family Xanthomonadaceae) (24.02%), Sphingobacterium (family Sphingobacteriaceae) (9.18%), Pararhizobium (family Rhizobiaceae) (5.19%), Lactobacillus (family Lactobacillaceae) (4.48%), Achromobacter (family ) (2.62%), Novosphingobium (family Sphingomonadaceae) (2.57%) and Dielma (family Erysipelotrichidae) (2.48%), were all enriched in C. elegans N2_UGM bacterial communities compared to SOIL [Xanthomonas (15.82%), Sphingobacterium (0.57%), Pararhizobium (0.14%), Lactobacillus (1.79%), Achromobacter (0.26%), Novosphingobium (0.04%) and Dielma (0.06%) (Figure 4.4).

222

Multivariate statistics of individual zOTUs revealed that sample type (SOIL or N2_UGM) had a significant effect on the bacterial community structure. A total of 324 zOTUs were significantly affected by the sample type (see Supplementary Materials; Table S4.4) of these, 72 zOTUs (22.2%) were enriched in the N2_UGM samples compared to the SOIL samples. This includes zOTUs from the class Clostridia (genera , Clostridium_AD, Clostridium_T, F0540, Epulopiscium_B, Anaerocolumna, Clostridium_M, Hespellia, Lachnoclostridium, MS4, D5 and Dethiosulfatibacter), the class Bacilli_A (genera Brevibacillus, Cohnella, DXL2, Gorillibacterium, Paenibacillus, Paenibacillus_G, Paenibacillus_S and UNC496MF), the class Alphaproteobacteria (genera Gluconacetobacter, Kaistia, Ochrobactrum, Ochrobactrum_A, Pararhizobium, Sinorhizobium, Novosphingobium, Sphingobium, Achromobacter and Comamonas), the class Bacteroidia (genera NS-102 and Sphingobacterium), the class Bacilli (genera Rummeliibacillus, Dielma and Haloplasma) and the class Actinobacteria (genus Rhodococcus) (see Supplementary Materials; Table S4.4).

223

(%)

Relative abundance

Figure 4.4 Taxonomic profiles of bacterial communities shown at the genus level of all SOIL and N2_UGM samples. Bacterial communities with a relative abundance less than 1% are grouped as ‘Other’

224

4.3.2 C. elegans with an undefined gut microbiota (N2_UGM) showed a reduced survival compared to the N2 monoxenic nematodes

To investigate the effect of an undefined gut microbiota (UGM) on C. elegans survival, the N2 monoxenic and N2_UGM nematodes were assayed against the Nkp- 1 expressing clones; E. coli HP1 and HG8 with the non-toxic E. coli BD24 as the control. After 2 days of exposure, C. elegans N2_UGM survival was reduced compared to the N2 monoxenic nematodes (One-way ANOVA followed by Tukey’s test; F (5, 12) = 477.5; p = 0.0216 for comparison between N2_UGM vs HP1 and N2 vs HP1, Figure 4.5, Supplementary Materials; Table S4.5). However, survival of C. elegans N2_UGM was also reduced when exposed to the non-toxic E. coli BD24 compared to the N2 nematodes (One-way ANOVA followed by Tukey’s test; F (5, 12) = 477.5; p < 0.0001, Figure 4.5, Supplementary Materials; Table S4.5), suggesting that the presence of the UGM was deleterious to the host. A closer inspection of the C. elegans N2_UGM revealed several morphological changes e.g. damage to the tissue surrounding the pharynx region and degenerated intestine (see Supplementary Materials, Figures S4.5 and S4.6).

225

100 N2_UGM vs HP1

N2_UGM vs HG8 )

% N2_UGM vs BD24

(

l

a N2 vs HP1

v

i v

r N2 vs HG8 u

s 50

N2 vs BD24

e

d

o

t

a

m

e N

0 0 5 10 15 20 Day

Figure 4.5 C. elegans N2 monoxenic and N2_UGM survival against E. coli HP1, HG8 and the non-toxic control BD24 bacterial strains. Statistical test revealed significant differences among the treatments (Log-rank (Mantel-Cox) test; df = 5; Chi square = 40.83; p < 0.0001). A One-way ANOVA and Tukey’s pairwise comparison test reveals that the survival of C. elegans N2_UGM was reduced upon exposure to the toxic E. coli HP1 (p = 0.0216) and also to the non-toxic control BD24 strains (p < 0.0001) compared to the N2 nematodes after two days of assay. Each data point represents means of nematode survival ± standard error from triplicate samples.

226

4.3.3 Isolation of bacterial cultures from C. elegans N2 UGM and establishment of C. elegans N2 with a defined gut microbiota (DGM)

Due to the observation that C. elegans N2 with an undefined gut microbiota had a reduced lifespan irrespective of the treatment with Nkp-1 expressing clones, I decided to test my original hypothesis using a defined gut microbiota (DGM). C. elegans N2 with a defined gut microbiota (herein referred to as C. elegans N2_DGM) were prepared by exposing the nematode larvae to the 22 bacterial isolates cultivated from C. elegans N2_UGM individuals (Table 4.2). Following the establishment of a defined gut microbiota (DGM), the morphology of N2_DGM nematodes was confirmed as intact as indicated by regular pharynx, intestine, gonad and tail structures (see Figure 4.6).

16S rRNA gene amplicon sequencing of the resulting gut microbiota revealed that gut microbiota of C. elegans N2_DGM consisted of all of the original isolates that comprised of eight different genera (indicated in Table 4.2). Of that resulting DGM, members from the genera Achromobacter followed by Ochrobactrum and Stenotrophomonas were the most dominant with an average relative abundance of ~ 44%, 27% and 19% respectively (see Figure 4.7B). In contrast, bacterial abundance from Pseudomonas (3.92%), Citrobacter (2.27%) and Sphingobacterium (2.16%) were reduced in the C. elegans N2_DGM bacterial composition (Figure 4.7B) compared their average relative abundance in the starting communities (Figure 4.7A) [Pseudomonas (9.09%, p = 0.0257), Citrobacter (18.18%, p < 0.0001) and Sphingobacterium (13.64%, p < 0.0001)] (Two-way ANOVA followed by Sidak’s test, F (8, 36) = 63.67, Supplementary Materials, Table S4.6).

227

Pharynx Terminal pharyngeal bulb

Eggs

Vulva Intestine Tail

Anus Intestine

Vulva

Figure 4.6 Morphology of C. elegans N2 after the establishment of a defined gut microbiota (DGM). The N2_DGM individuals showed an intact (A) whole body structure (B) pharynx (C) intestine and (D) anal region. Nematode images were acquired using the Fluorescence Microscope (Olympus BLX-61) equipped with the DIC filter under the 10x (A) and 40x magnification (D). Scale bars indicate 100 µm (A) or 20 µm (B-D).

228

Table 4.2 List of isolated gut bacteria from C. elegans N2_UGM previously raised in the enriched SOIL that were detected in the N2_DGM nematodes.

Isolate Identity/Query Identity** Closest Identification / NCBI Accession Number Matched zOTUs* number cover (%) (%) Isolate 1 Sphingobacterium multivorum strain IAE9 16S ribosomal RNA gene, partial sequence 99.9/100 zOTU47, zOTU29, 99.9 (MK415039.1) zOTU22, zOTU13 Isolate 2 Sphingobacterium multivorum strain IAE9 16S ribosomal RNA gene, partial sequence 99.7/100 zOTU47, zOTU29, 99.7 (MK415039.1) zOTU22, zOTU13 Isolate 3 Bosea sp. dv3 16S ribosomal RNA gene, partial sequence (FJ774000.1) 100/100 zOTU46 100 Isolate 4 Pseudomonas oryzihabitans strain Os_Ep_PSA_16 16S ribosomal RNA gene, partial 100/100 zOTU40 100 sequence (MN932348.1) Isolate 5 Ochrobactrum anthropi strain A8 16S ribosomal RNA gene, partial sequence 100/100 zOTU3 100 (MN252068.1) Isolate 6 Achromobacter denitrificans strain PR1, complete genome (CP020917.1) 90.8/99 zOTU2 90.83 Isolate 7 Ochrobactrum lupini strain SH24_27f 16S ribosomal RNA gene, partial sequence 100/100 zOTU3 100 (MK453290.1) Isolate 8 Ochrobactrum intermedium strain M16-10-4 16S ribosomal RNA gene, partial 100/100 zOTU12 100 sequence (MN606156.1) Isolate 9 Ochrobactrum anthropi strain A8 16S ribosomal RNA gene, partial sequence 100/100 zOTU3 100 (MN252068.1) Isolate 10 Leuconostoc mesenteroides strain NM188-2 16S ribosomal RNA gene, partial 90.6/99 zOTU262 90.57 sequence (HM218781.1) Isolate 11 Pseudomonas sp. strain ADP partial 16S rRNA gene, strain DSM 11735 100/99 ZOTU37 100 (AM088478.1)

*Closest match to the N2_DGM zOTUs **Identity of each bacterial isolate to the N2_DGM zOTUs

229

Table 4.2 (Continue) List of isolated gut bacteria from C. elegans N2_UGM previously raised in the enriched SOIL that were detected in the N2_DGM nematodes.

List of Identity/Query Matched Identity** Closest Identification / NCBI Accession Number isolates cover (%) zOTUs* (%) Isolate 12 Uncultured Stenotrophomonas sp. clone Sm175 16S ribosomal RNA gene, partial sequence 94.4/99 zOTU6 94.35 (MH569118.1) Isolate 13 Citrobacter freundii strain EF57 16S ribosomal RNA gene, partial sequence (MN582978.1) 98.2/99 zOTU35 98.16 Isolate 14 Achromobacter denitrificans strain A15 16S ribosomal RNA gene, partial sequence 99.5/100 zOTU2 99.48 (MN252073.1) Isolate 15 Achromobacter sp. strain RABA6 16S ribosomal RNA gene, partial sequence (MN022537.1) 99.8/100 zOTU2 99.82 Isolate 16 Citrobacter freundii strain OG052 16S ribosomal RNA gene, partial sequence 100/100 zOTU39 100 (CMK353941.1) Isolate 17 Achromobacter sp. strain JM-1 16S ribosomal RNA gene, partial sequence (MN220504.1) 100/100 zOTU2 100 Isolate 18 Citrobacter sp. strain ZX-57 16S ribosomal RNA gene, partial sequence (MF148484.1) 99.9/100 zOTU39 99.89 Isolate 19 Ochrobactrum anthropi strain TY171-20 16S ribosomal RNA gene, partial sequence 99.9/100 zOTU3 99.9 (MT083950.1) Isolate 20 Sphingobacterium multivorum strain ALS-5 16S ribosomal RNA gene, partial sequence 99.8/100 zOTU21 99.79 (KJ638992.1) Isolate 21 Ochrobactrum anthropi strain TY171-20 16S ribosomal RNA gene, partial sequence 100/100 zOTU3 100 (MT083950.1) Isolate 22 Citrobacter sp. strain ZX-57 16S ribosomal RNA gene, partial sequence (MF148484.1) 100/100 zOTU39 100

*Closest match to the N2_DGM zOTUs **Identity of each bacterial isolate to the N2_DGM zOTUs

230

A A B B

Figure 4.7 The average of bacterial relative abundances associated with (A) the starting communities and (B) the resulting gut microbiota of C. elegans N2_DGM at the genus level. Bacteria belonging to the genera Achromobacter, Ochrobactrum and Stenotrophomonas showed the highest relative abundance in the gut microbial communities of C. elegans N2_DGM. While the abundance of bacteria from the genera Achromobacter and Stenotrophomonas were increased in the N2_DGM nematodes (p < 0.0001), a reduced abundance was observed in Pseudomonas (p = 0.0257), Citrobacter (p < 0.0001) and Sphingobacterium (p < 0.0001) compared to the initial abundance of each genera in the starting communities. The differences of Ochrobactrum, Bosea and Leuconostoc were not significant. The sequences related to the genus Paramesorhizobium in the final N2_DGM communities is presumably resulting from contamination or a sequencing error as they did not match to any of the 22 original isolates.

231

4.3.4 A defined gut microbiota provides resistance to E. coli HP1 compared to monoxenic nematodes

To investigate the effect of a defined gut microbiota (DGM) on C. elegans survival, the N2_DGM and N2 monoxenic nematodes were assayed against the Nkp-1 expressing E. coli strains; HP1 and HG8 and the non-toxic control E. coli BD24. After two days, the survival of C. elegans N2_DGM was increased compared to the N2 monoxenic nematodes upon exposure to E. coli HP1 (One-way ANOVA followed by

Tukey’s test; F (5, 12) = 330.6; p < 0.0001, Figure 4.7, Supplementary Materials; Table S4.7). However, there was no statistical support for the increased survival of the N2_DGM nematode compared to the N2 monoxenic upon exposure to the HG8 bacteria (One-way ANOVA followed by Tukey’s test; F (5, 12) = 330.6; p = 0.5489, Figure 4.7, Supplementary Materials; Table S4.7). Furthermore, treatment against the non-toxic control E. coli BD24 did not affect C. elegans N2_DGM or N2 nematode survival (One-way ANOVA followed by Tukey’s test; F (5, 12) = 330.6; p > 0.9999 for comparison between N2_DGM vs BD24 and N2 vs BD24, Figure 4.7, Supplementary Materials; Table S4.7). These findings show firstly that unlike the undefined gut microbiota nematode, fitness is not affected by the presence of a defined gut microbiota. Secondly, that the presence of the defined gut microbiota can improve C. elegans survival when exposed to the toxic E. coli HP1 cultures.

232

100 N2_DGM vs HP1 N2_DGM vs HG8

)

% (

N2_DGM vs BD24

l

a v

i N2 vs HP1

v

r u

s 50

N2 vs HG8

e d

o

t N2 vs BD24 a

m e N

0 0 5 10 15 20

Day

Figure 4.8 Survival of C. elegans N2_DGM and N2 monoxenic when exposed to E. coli HP1, HG8 and the non-toxic control BD24 bacterial strains. Statistical test revealed significant differences among the treatments (Log-rank (Mantel-Cox) test; df = 5; Chi square = 58.73; p < 0.0001). A One-way ANOVA and Tukey’s pairwise comparison test reveals that the association of a defined gut microbiota successfully increased C. elegans N2_DGM survival compared to the N2 monoxenic nematodes after two days of HP1 bacterial exposure (p < 0.0001). While a slight increase on C. elegans N2_DGM survival was observed upon exposure to HG8, the difference compared to the N2 nematodes was not significant (p = 0.5489). Both C. elegans N2_DGM and N2 monoxenic also showed a normal survival rate when exposed to E. coli BD24 (p > 0.9999). Each data point represents means of nematode survival ± standard error from triplicate samples.

233

4.3.5 Reduction of Nkp-1 expressing E. coli colonisation in C. elegans with a defined gut microbiota compared to the N2 monoxenic nematodes

To determine if a defined gut microbiota in C. elegans is able to mitigate colonisation of the Nkp-1 expressing strains, C. elegans N2 monoxenic and N2_DGM were exposed to the GFP-tagged E. coli HP1 and HG8 and subjected to microscopic investigation. Colonisation of all bacterial strains increased over time but was consistently higher in C. elegans N2 monoxenic compared to the N2_DGM nematodes. However statistical support for this trend was only found for the increased HG8::GFP colonisation of N2 monoxenic nematodes compared to N2_DGM nematodes at 24 hours (p = 0.0003, One-way ANOVA followed by Sidak’s test; F (5,

12) = 43.82) and for the higher colonisation of E. coli HP1::GFP and HG8::GFP strains to C. elegans N2 compared with the N2_DGM nematodes after 72 hours (One-way

ANOVA followed by Sidak’s test; F (5, 12) = 20.36; p = 0.0188 and p = 0.0001 respectively, Figure 4.9, Supplementary Materials Table S4.8). These results indicate that association of a defined gut microbiota was able to successfully reduce E. coli HP1::GFP and HG8::GFP colonisation in C. elegans N2_DGM.

234

1.0

N2 vs HP1::GFP

e

d

o t

a 0.8 N2_DGM vs HP1::GFP

m e

n * N2 vs HG8::GFP

/

e

c 0.6 N2_DGM vs HG8::GFP

n

e c

s N2 vs BD24::GFP e r 0.4

o *** N2_DGM vs BD24::GFP

u

l

f

e

v i

t 0.2

a

l

e R 0.0 24 h 48 h 72 h Time

Figure 4.9 Relative fluorescence representing the colonisation of E. coli HP1::GFP, HG8::GFP and the non-toxic control E. coli BD24::GFP in C. elegans N2 monoxenic and N2_DGM. Nematode images were acquired using the Fluorescence Microscope Olympus BLX-61 equipped with the DIC and FITC filter. The relative fluorescence for each image was quantified using the ImageJ. There was no difference for HP1::GFP colonisation observed in both N2 monoxenic or N2_DGM nematodes after 24 hours of exposure (p = 0.2956). However, the increased colonisation of HG8::GFP in N2 compared to N2_DGM was significant (***; p = 0.0003). After 48 hours, the differences of HP1::GFP and HG8::GFP colonisation in C. elegans N2 and N2_DGM were not significant (p = 0.3222 and 0.1358 respectively). The highest HP1::GFP and HG8::GFP colonisation was shown in C. elegans N2 monoxenic compared to the N2_DGM nematodes after 72 hours of bacterial exposure (*; p = 0.0188 and ***; p = 0.0001 respectively). Data were presented as the mean of relative fluorescence from triplicate samples ± standard error.

235

4.3.6 Morphology of C. elegans N2_DGM with a defined gut microbiota was less affected by E. coli HP1::GFP and HG8::GFP compared to the N2 monoxenic nematodes

Fluorescence microscopy was used to observe the extent of morphological changes to C. elegans strains on exposure to the GFP-tagged E. coli clones expressing Nkp-1. Only strains exposed to toxic clones (i.e. HP1::GFP and HG8::GFP but not the control strain BD24::GFP) showed signs of morphological changes including pharynx distortion (Figure 4.10C,D), internal hatching (Figure 4.11C,D) and internal organ (intestine or gonad) damage (Figure 4.12C,D). Deformed anal region (dar) was only observed in E. coli HP1::GFP exposed animals (Figure 4.13C,D) consistent with previous results (see section 3.3.2, Chapter 3).

The presence of a defined gut microbiota significantly reduced the number of nematodes with pharynx distortion (Figure 4.10A, Table 4.3), internal hatching (Figure 4.11A, Table 4.3) and internal organ damage (Figure 4.12A) after exposure to the E. coli HP1::GFP clone. Dar formation was also reduced in C. elegans N2_DGM after 24 and 48 hours (Figure 4.13A, Table 4.3), however was significantly higher in these nematodes compared to C. elegans N2 after 72 hours exposure (Two-way

ANOVA followed by Tukey’s test; F (10, 24) = 28.72, p = 0.0482, Figure 4.13A, Table 4.3, Supplementary Materials, Table S4.9).

In contrast to the observations with N2_DGM individuals exposed to E. coli HP1::GFP, exposure of HG8::GFP to the defined gut microbiota had no effect in reducing pharynx distortion or internal hatching (Figures 4.10A and 4.11A respectively, Table 4.3). However, a reduction of internal organ damage was observed in the N2_DGM nematodes following 24 hours exposure to E. coli HG8::GFP (see Figure 4.12A, Table 4.3). Further statistical data are described in the Supplementary Materials, Table S4.9).

236

A **

) 50 %

( N2 vs HP1::GFP

n

o i

t N2_DGM vs HP1::GFP

r 40

o *

t s

i N2 vs HG8::GFP

d

x

n 30 N2_DGM vs HG8::GFP

y

r a

h N2 vs BD24::GFP

p

h 20

t N2_DGM vs BD24::GFP

i

w

s e

d 10

o

t

a

m e

N 0 24 h 48 h 72 h Time

B C D

Figure 4.10 Proportion of C. elegans N2 monoxenic and N2_DGM with pharynx distortion. (A) Average percentage of the assayed nematodes with pharynx distortion from triplicate samples ± standard error (n = 3). A higher percentage of pharynx distortion was shown by N2 monoxenic nematodes compared to N2_DGM after 24 hours (*; p = 0.0348) and 72 hours of HP1::GFP bacterial exposure (**; p = 0.0087). An example of normal pharynx was demonstrated by nematode when exposed to the control strain BD24::GFP (B). In contrast, exposure to the HP1::GFP and HG8::GFP resulted in pharynx distortion (indicated by the white arrow) in the N2 (C) and N2_DGM nematodes (D). The green fluorescence indicating the GFP-tagged bacterial colonisation was observed in the N2 monoxenic nematodes (C). Images were captured under 40x magnification using the DIC and FITC filters and subsequently merged. Scale bars indicate 20µm.

237

A

**

) 100

% N2 vs HP1::GFP

(

g n

i 80 N2_DGM vs HP1::GFP

h

c t

a N2 vs HG8::GFP

h

l

a 60 * N2_DGM vs HG8::GFP

n

r

e t

n N2 vs BD24::GFP

i

h *

t 40

i N2_DGM vs BD24::GFP

w

s

e d

o 20

t

a

m e

N 0 24 h 48 h 72 h Time

B C D

Figure 4.11 Proportion of C. elegans N2 monoxenic and N2_DGM with internal hatching. (A) Average percentage of the assayed nematodes with internal hatching from triplicate samples ± standard error (n = 3). A higher percentage of internal hatching was shown by N2 monoxenic nematodes compared to N2_DGM after 24, 48 and 72 hours of HP1::GFP exposure (*; p < 0.05). In contrast, the N2_DGM nematodes demonstrated a higher proportion of internal hatching compared to N2 monoxenic after 24 hours of exposure to HG8::GFP (**; p = 0.0063). An example of normal body with intact eggs was demonstrated by nematode when exposed to the control strain BD24::GFP (B). In contrast, exposure to the HP1::GFP and HG8::GFP resulted in internal hatching in the N2 monoxenic (C) and N2_DGM nematodes (D). The green fluorescence indicating GFP-tagged bacterial colonisation was observed on the N2 nematodes (C). Images were captured under 40x magnification using the DIC and FITC filters and subsequently merged. Scale bars indicate 20µm.

A 238

) %

( 150

e N2 vs HP1::GFP

g a

m N2_DGM vs HP1::GFP

a

d

n N2 vs HG8::GFP a

g 100

r * o

N2_DGM vs HG8::GFP

l

a n

r N2 vs BD24::GFP

e

t

n i N2_DGM vs BD24::GFP

h 50

t

i

w

s

e

d

o

t

a m

e 0

N 24 h 48 h 72 h Time

BB Eggs C D

Intestine

Gonad

Figure 4.12 Proportion of C. elegans N2 monoxenic and N2_DGM with internal organ (intestine or gonad) damage. (A) Average percentage of the assayed nematodes showing internal organ damage from triplicate samples ± standard error (n = 3). After 24 hours, a higher percentage of internal organ damage was shown by the N2 monoxenic compared to the N2_DGM nematodes when exposed to the HG8::GFP (*; p = 0.0289). For the following time points; 48 and 72 hours, all of the N2 monoxenic and N2_DGM nematodes showed internal organ damage after exposure to the HG8::GFP bacteria. After 24, 48 and 72 hours of exposure to HP1::GFP, only N2 monoxenic nematodes showed the internal organ damage phenotype (*; p < 0.05), but not the N2_DGM individuals. An example of normal body structure was demonstrated by nematode exposed to the strain BD24::GFP (B). In contrast, exposure to the HG8::GFP strain resulted in internal organ damage in N2 monoxenic (C) and N2_DGM nematodes (D). The green fluorescence indicating GFP-tagged bacterial colonisation was observed in the nematodes (C and D). Images were captured under 40x magnification using the DIC and FITC filters and subsequently merged. Scale bars indicate 20µm.

239

A

100 N2 vs HP1::GFP *

) 80 N2_DGM vs HP1::GFP

%

(

r N2 vs HG8::GFP

a D

60 N2_DGM vs HG8::GFP

h

t

i w

N2 vs BD24::GFP

s e

d 40

o N2_DGM vs BD24::GFP

t

a

m e

N 20

0 24 h 48 h 72 h Time

B C D

Figure 4.13 Proportion of C. elegans N2 monoxenic and N2_DGM with deformed anal region (dar). (A) Average percentage of the assayed nematodes showing dar from triplicate samples ± standard error (n = 3). A higher percentage of dar was shown by N2 monoxenic nematodes compared to N2_DGM after 48 hours of HP1::GFP exposure (*; p = 0.0489). However, after 72 hours, a higher proportion of dar was shown by N2_DGM compared to N2 monoxenic individuals (*; p = 0.0482). An example of normal anal region was demonstrated by nematode exposed to the strain BD24::GFP (B). In contrast, the dar phenotype (indicated by the white arrow) was demonstrated by N2 monoxenic (C) and N2_DGM nematodes (D) only upon exposure to the HP1::GFP strain. The green fluorescence indicating HP1::GFP bacterial colonisation was observed on the N2 monoxenic nematodes (C). Images were captured under 40x magnification using the DIC and FITC filters and subsequently merged. Scale bars indicate 20µm.

240

Table 4.3 Summary of morphological changes observed on C. elegans N2 monoxenic and N2_DGM exposed to the toxic Nkp-1 expressing strains (E. coli HP1::GFP and HG8::GFP) and the non-toxic control BD24::GFP. P-values indicate the difference of comparison between nematodes showing the lowest proportion of the tested phenotypes compared to the other assayed animals and were calculated using a Two-way ANOVA followed by Tukey’s pairwise comparison test. ‘ND’ indicates ‘Not Detected’ whilst ‘-’ shows no p-value calculated. A p-value < 0.05 is considered as statistically significant.

Bacterial exposure Morphological phenotype HP1::GFP HG8::GFP BD24::GFP 24 h 48 h 72 h 24 h 48 h 72 h 24 h 48 h 72 h Pharynx distortion Nematode showing the lowest N2_DGM N2_DGM N2_DGM N2_DGM N2_DGM N2 ND ND ND proportion of pharynx distortion p-value 0.0348 0.9251 0.0087 0.8094 0.971 0.8866 - - -

Internal hatching Nematode showing the lowest N2_DGM N2_DGM N2_DGM N2 N2 N2_DGM ND ND ND proportion of internal hatching p-value 0.0289 0.0208 0.0141 0.0063 0.7587 0.4196 - - -

Internal organ damage Nematode showing the lowest N2_DGM N2_DGM N2_DGM N2_DGM N2_DGM N2_DGM ND ND ND proportion of internal organ & N2 damage p-value 0.1919 0.5302 0.0224 0.0289 0.8866 - - - -

Deformed anal region (dar) Nematode showing the lowest N2_DGM N2_DGM N2 ND ND ND ND ND ND proportion of dar p-value 0.1914 0.0489 0.0482 ------

241

4.4 DISCUSSION

In Chapter 3, colonisation of Nkp-1 expressing E. coli strains caused necrosis and severe physical damages against C. elegans N2 monoxenic, leading to the animals’ mortality. Therefore, to investigate whether the presence of gut microbiota can enhance survival of C. elegans N2 when exposed to Nkp-1 toxicity, the nematodes were established with an undefined gut microbiota (UGM) originating from the soil environment. Since the presence of an UGM reduced the nematodes survival, C. elegans N2 with a defined gut microbiota (DGM) were established based on the isolation of individual bacteria from the N2_UGM nematodes (and confirmed using the 16S rRNA gene amplicon sequencing). The DGM did not affect C. elegans fitness and nematode survival was enhanced by the DGM establishment upon exposure to the Nkp-1 producing E. coli clones.

4.4.1 Establishment of potentially detrimental gut microbiota reduce C. elegans survival

Establishment of an undefined gut microbiota (UGM) from the enriched soil environment resulted in diminished C. elegans N2 survival. The reduced survival could be due to external physico-chemical factors for example suboptimal pH, osmolarity or chemical properties of the soil (Gbadegesin et al., 1993; Qi et al., 2018). In addition, the complexity of the soil microbial communities which consists of bacteria, fungi, protists and viruses and their by-products may also affect the nematodes fitness (Jiang and Wang, 2018). The importance of these biotic and abiotic factors is further supported by evidence that shows a naturally shortened C. elegans lifespan in the soil environment compared to the nematodes that were raised on agar media (Van Voorhies et al., 2005). While the relative influence of these factors is still to be determined, 16S rRNA gene amplicon sequencing analysis of the undefined gut microbiota (UGM) of C. elegans N2 identified a number of interesting bacterial taxa that were significantly different compared to the soil microbiome from which they originated. Members from the families Xanthomonadaceae, Rhizobiaceae Sphingobacteriaceae, Lactobacillaceae, Burkholderiaceae, Sphingomonadaceae and Erysipelotrichaceae were enriched in C. elegans N2_UGM

242 compared to the soil samples. Bacteria belonging to these families are commonly representing the gut microbiota of C. elegans either in those that were isolated from the wild (i.e. from naturally decaying fruits or snail vector) or from microcosms emulating natural habitats (Berg et al., 2016; Dirksen et al., 2016; Samuel et al., 2016). While the overlap of bacterial taxa found in C. elegans N2_UGM with the other studies (Berg et al., 2016; Dirksen et al., 2016; Samuel et al., 2016) may support the existence of a C. elegans core gut microbiota, the relatively high abundance of taxa previously shown to cause disease (e.g. Xanthomonas and Sphingobacterium) (Samuel et al., 2016) may also explain the reduced lifespan of N2_UGM nematodes.

Previous studies have shown that the plant pathogen Xanthomonas oryzae is toxic to C. elegans (Bai et al., 2014) and exposure to specific strains of Sphingobacterium sp. diminished brood size (Dirksen et al., 2016), reduced growth and activated internal nematode pathogen gene reporters for example those encoding for heat shock proteins (hsp); hsp-4::GFP or hsp-6::GFP (Samuel et al., 2016), suggesting at least some members of this taxa are pathogenic to C. elegans. Closer inspection of C. elegans N2_UGM individuals revealed evidence of a damaged and degraded pharynx (see Supplementary Materials, Figure S6). Interestingly a recent study has shown similar disease- symptoms in C. elegans exposed to Chryseobacterium nematophagum (also referred to as golden bacteria) (Page et al., 2019). Future isolation and analyses of the microbes associated with N2_UGM may reveal the presence of similar soil borne pathogens and shed further light on why these nematodes have a shortened lifespan.

4.4.2 The high relative abundance of potentially beneficial gut bacteria may improve C. elegans survival against the toxic Nkp-1 expressing clones

Unlike the undefined gut microbiota (UGM), the establishment of a defined gut microbiota (DGM) was not deleterious to C. elegans instead it enhanced nematode survival and diminished the ability of Nkp-1 expressing E. coli to colonise the gut. The 16S rRNA gene amplicon sequencing analysis of the gut microbiota of C. elegans N2_DGM revealed that members from the genera Achromobacter followed by Ochrobactrum and Stenotrophomonas comprised the majority (~ 90.36% of the relative 243 abundance) of the bacteria that successfully established in the C. elegans gut. Other studies have also reported Achromobacter, Ochrobactrum and Stenotrophomonas as the main members of the native C. elegans gut microbiota (Dirksen et al., 2016). Species within these three genera are also known to associate with entomopathogenic nematodes (EPN) including Rhabditis sp., and Heterorhabditis sp. (Babic et al., 2000; Deepa et al., 2015; Wollenberg et al., 2016); the free living soil nematode Acrobeloides maximus (Baquiran et al., 2013) and the pinewood nematode Bursaphlenchus xylophilus (Cheng et al., 2013). The dominance of Achromobacter, Ochrobactrum and Stenotrophomonas maybe in part due to their reported resistance to antimicrobial activities of other bacteria (Crossman et al., 2008; Furlan and Stehling, 2017) which could provide a selective advantage for the colonisation of the nematode’s gut system compared to the other susceptible microorganisms.

The widespread occurrence of members of the Achromobacter, Stenotrophomonas and Ochrobactrum across diverse nematode species suggests they provide the nematode host with an evolutionary advantage driven by beneficial traits (Baquiran et al., 2013). For example, Stenotrophomonas and Ochrobactrum improve C. elegans fitness under changing osmotic or heat stress condition (Dirksen et al., 2016), while Achromobacter sp. and Achromobacter denitrificans conferred host protection against the Staphylococcus aureus, Pseudomonas aeruginosa and Proteus mirabilis infection by producing bactericidal arginine derived cyclic dipeptides or secondary metabolite i.e. prodigiosin, (Deepa et al., 2015; Erandapurathukadumana Sreedharan et al., 2020). Members from the genus Ochrobactrum (and Pseudomonas) are the main suppliers of vitamin B12 (cobalamin) to C. elegans that enhances nematode fitness (Zimmermann et al., 2020) while some of the species (e.g. Ochrobactrum anthropi) can improve fatty acid, amino acid, folate and energy metabolism of animals (Yang et al., 2019). Moreover, the genus Stenotrophomonas possess vital metabolic pathways to catalyse the degradation of hazardous compounds, hence protecting the associated nematodes from toxicity (Cheng et al., 2013).

244

While further work is required to understand the mechanisms by which the DGM mitigates the toxicity of Nkp-1 producing E. coli, the broadly beneficial function of members of this community including Achromobacter sp. and Stenotrophomonas sp. may provide an explanation. For example, the possible antibacterial and antibiofilm capabilities of these strains (Amara et al., 2011; Christiaen et al., 2014; de Rossi et al., 2014; Deepa et al., 2015; Erandapurathukadumana Sreedharan et al., 2020) could directly prevent the colonisation and proliferation of Nkp-1 producing E. coli within the nematode gut. These strains may also have indirect mechanisms of protecting C. elegans from Nkp- 1 toxicity, for example through the production of essential nutrients. Members of genus Ochrobactrum can synthesise vitamin B12 and may supply it to the nematode (Zimmermann et al., 2020). Ensuring adequate B12 will mitigate the negative impacts of vitamin deficiency (Watson et al., 2016) such as increased susceptibility to environmental stress (Revtovich et al., 2019), including pathogens and in this case Nkp-1 producing E. coli.

While members of the potentially detrimental bacterial genera Sphingobacterium were present in the gut microbiota of N2_DGM nematodes, their relative abundance was low compared to that of N2_UGM nematodes. Thus, any adverse effect of these strains may have been mitigated by the beneficial gut microbiota (Samuel et al., 2016). It is also important to note that not all members of any given bacterial genera are likely to be harmful and it is possible that the individual strains of Sphingobacterium in N2_DGM nematodes are different to those in N2_UGM.

4.4.3 An increased proportion of potentially beneficial gut microbiota diminish the physical damage caused by the Nkp-1 expressing strains

Further inspection of C. elegans exposed to Nkp-1 producing E. coli showed that animals harbouring a DGM suffered from less physical injuries (i.e. pharynx distortion, internal organ damage or internal hatching) than animals with a monoxenic gut microbiota. In addition to the direct and indirect impacts of specific members of the DGM mentioned above, reduced damage in nematodes with a DGM may be due to priming of the innate 245 immunity pathways by the DGM, thus resulting in better preparedness upon exposure to the toxic Nkp-1 bacteria (Nyholm and Graf, 2012; Montalvo-Katz et al., 2013). This hypothesis is supported by previous work that has demonstrated enhanced p38 MAPK immunity pathway induced by the commensal bacterium Pseudomonas mendocina and resulting in increased C. elegans survival upon exposure to the pathogen Pseudomonas aeruginosa (Montalvo-Katz et al., 2013). Likewise, C. elegans exposed to the probiotic Lactobacillus acidophilus showed increased survival when challenged against pathogens Enterococcus faecalis and Staphylococcus aureus (Kim and Mylonakis, 2012). In these examples C. elegans recognises microorganisms through the microbe-associated molecular patterns (MAMPs) molecules on the cells surface (for example peptidoglycan or ) resulting in the primed activation of innate immunity system in the animals (Nyholm and Graf, 2012). However, in some cases, additional perturbation as a result of toxin production in harmful bacteria leads to the activation of the downstream microbial effectors to combat the infection (Pukkila-Worley and Ausubel, 2012; Cohen and Troemel, 2015).

The possibility that establishment of a DGM enhanced the innate immunity of C. elegans by the DGM may also explain the increased proportion of these nematodes with an observed dar phenotype relative to nematodes with a monoxenic gut microbiota (Figure 4.13). Previous studies have shown that dar formation can be induced by genes within the innate immune pathways p38-MAPK (Tanaka-Hino et al., 2002; Battisti et al., 2017) and ERK-MAPK (Nicholas and Hodgkin, 2004). While not tested here the likely increase in these immune pathways in nematodes with DGM could have indirectly resulted in an increase in dar formation in N2_DGM nematodes exposed to HP1 compared to the N2 monoxenic nematodes. Whilst the precise function of dar formation in C. elegans is unclear (Hodgkin et al., 2000), the rectal epithelial swelling is believed to assist the nematodes with efficient removal of potentially harmful bacteria (e.g. E. coli strains producing Nkp-1) from the nematode intestinal system (Hodgkin et al., 2000; Anderson et al., 2019).

246

4.5 CONCLUSION

This chapter found that the establishment of an UGM in C. elegans N2 could not support the nematodes survival. By contrast, establishment of individuals with a DGM successfully improved their survival, reduced colonisation by Nkp-1 producing bacteria and mitigated the resulting physical damages. Furthermore, dar formation, which is believed to be a mechanism used by the nematodes to remove the toxic Nkp-1 expressing strains, was increased in the N2 nematodes associated with the DGM.

The differences observed on nematode survival, Nkp-1 bacterial colonisation and the resulting physical damages are hypothesised to result from the difference in abundance of potentially detrimental compared to beneficial bacterial groups detected in the UGM compared to DGM. The 16S rRNA gene sequencing analysis confirmed that the main constituents of the UGM consisted of potentially detrimental bacteria i.e. Xanthomonas and Sphingobacterium. By contrast, Sphingobacterium was reduced in the DGM whilst the potentially beneficial bacteria belonging to Achromobacter, Ochrobactrum and Stenotrophomonas were highly abundant, thus possibly contribute to the improved nematode survival when exposed to toxic compounds for example Nkp-1. While it is important to consider how other external factors not tested here (e.g. environmental pH and microorganisms other than bacteria) could be affecting the nematode lifespan, my results highlight that the intestinal gut bacteria can mitigate the effects of a potent anti- nematode compound (i.e. Nkp-1). Therefore, future work should carefully consider the efficacy of potential drugs under both standard laboratory conditions with a monoxenic nematodes and under conditions that represent nematodes with a more diverse gut microbiota.

247

4.6 REFERENCES

Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology, 215(3), 403-410. Altschul, S. F., Madden, T. L., Schäffer, A. A., Zhang, J., Zhang, Z., Miller, W., & Lipman, D. J. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Research, 25(17), 3389-3402. Amara, N., Krom, B. P., Kaufmann, G. F., & Meijler, M. M. (2011). Macromolecular inhibition of quorum sensing: enzymes, antibodies, and beyond. Chemical Reviews, 111(1), 195-208. Anderson, A., Chew, Y. L., Schafer, W., & McMullan, R. (2019). Identification of a conserved, orphan G protein-coupled receptor required for efficient pathogen clearance in Caenorhabditis elegans. Infection and Immunity, 87(4), e00034- 00019. Anderson, M. J. (2001). A new method for non‐parametric multivariate analysis of variance. Austral Ecology, 26(1), 32-46. Babic, I., Fischer-Le Saux, M., Giraud, E., & Boemare, N. (2000). Occurrence of natural dixenic associations between the symbiont Photorhabdus luminescens and bacteria related to Ochrobactrum spp. in tropical entomopathogenic Heterorhabditis spp. (Nematoda, Rhabditida). Microbiology, 146 (3), 709-718. Bai, Y., Zhi, D., Li, C., Liu, D., Zhang, J., Tian, J., Wang, X., Ren, H., & Li, H. (2014). Infection and immune response in the nematode Caenorhabditis elegans elicited by the phytopathogen Xanthomonas. Journal of Microbiology and Biotechnology, 24(9), 1269-1279. Ballestriero, F., Thomas, T., Burke, C., Egan, S., & Kjelleberg, S. (2010). Identification of compounds with bioactivity against the nematode Caenorhabditis elegans by a screen based on the functional genomics of the marine bacterium Pseudoalteromonas tunicata D2. Applied and Environmental Microbiology, 76(17), 5710-5717. Baquiran, J.-P., Thater, B., Sedky, S., De Ley, P., Crowley, D., & Orwin, P. M. (2013). Culture-independent investigation of the microbiome associated with the nematode Acrobeloides maximus. PloS One, 8(7), e67425. Battisti, J. M., Watson, L. A., Naung, M. T., Drobish, A. M., Voronina, E., & Minnick, M. F. (2017). Analysis of the Caenorhabditis elegans innate immune response to Coxiella burnetii. Innate Immunity, 23(2), 111-127. Berg, M., Monnin, D., Cho, J., Nelson, L., Crits-Christoph, A., & Shapira, M. (2019). TGFβ/BMP immune signaling affects abundance and function of C. elegans gut commensals. Nature Communications, 10(1), 604. Berg, M., Stenuit, B., Ho, J., Wang, A., Parke, C., Knight, M., Alvarez-Cohen, L., & Shapira, M. (2016). Assembly of the Caenorhabditis elegans gut microbiota from diverse soil microbial environments. The ISME Journal, 10(8), 1998-2009. Blander, J. M., Longman, R. S., Iliev, I. D., Sonnenberg, G. F., & Artis, D. (2017). Regulation of inflammation by microbiota interactions with the host. Nature Immunology, 18(8), 851-860. Bolger, A. M., Lohse, M., & Usadel, B. (2014). Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics, 30(15), 2114-2120. Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics, 77(1), 71-94.

248

Burke, C., Thomas, T., Egan, S., & Kjelleberg, S. (2007). The use of functional genomics for the identification of a gene cluster encoding for the biosynthesis of an antifungal tambjamine in the marine bacterium Pseudoalteromonas tunicata. Environmental Microbiology, 9(3), 814-818. Camacho, C., Coulouris, G., Avagyan, V., Ma, N., Papadopoulos, J., Bealer, K., & Madden, T. L. (2009). BLAST+: architecture and applications. BMC Bioinformatics, 10(1), 421. Cani, P. D. (2018). Human gut microbiome: hopes, threats and promises. Gut, 67(9), 1716. Chen, Y.-L., Lee, C.-C., Lin, Y.-L., Yin, K.-M., Ho, C.-L., & Liu, T. (2015). Obtaining long 16S rDNA sequences using multiple primers and its application on dioxin- containing samples. BMC Bioinformatics, 16 (Suppl 18), S13. Cheng, X.-Y., Tian, X.-L., Wang, Y.-S., Lin, R.-M., Mao, Z.-C., Chen, N., & Xie, B.-Y. (2013). Metagenomic analysis of the pinewood nematode microbiome reveals a symbiotic relationship critical for xenobiotics degradation. Scientific Reports, 3(1), 1869. Christiaen, S. E. A., Matthijs, N., Zhang, X.-H., Nelis, H. J., Bossier, P., & Coenye, T. (2014). Bacteria that inhibit quorum sensing decrease biofilm formation and virulence in Pseudomonas aeruginosa PAO1. Pathogens and Disease, 70(3), 271- 279. Claus, S. P., Guillou, H., & Ellero-Simatos, S. (2016). The gut microbiota: a major player in the toxicity of environmental pollutants? Biofilms and Microbiomes, 2(1), 1- 11. Cohen, L. B., & Troemel, E. R. (2015). Microbial pathogenesis and host defense in the nematode C. elegans. Current Opinion in Microbiology, 23, 94-101. Crossman, L. C., Gould, V. C., Dow, J. M., Vernikos, G. S., Okazaki, A., Sebaihia, M., Saunders, D., Arrowsmith, C., Carver, T., & Peters, N. (2008). The complete genome, comparative and functional analysis of Stenotrophomonas maltophilia reveals an organism heavily shielded by drug resistance determinants. Genome Biology, 9(4), R74. de Rossi, B. P., García, C., Alcaraz, E., & Franco, M. (2014). Stenotrophomonas maltophilia interferes via the DSF-mediated quorum sensing system with Candida albicans filamentation and its planktonic and biofilm modes of growth. Revista Argentina de Microbiología, 46(4), 288-297. Deepa, I., Kumar, S. N., Sreerag, R. S., Nath, V. S., & Mohandas, C. (2015). Purification and synergistic antibacterial activity of arginine derived cyclic dipeptides, from Achromobacter sp. associated with a rhabditid entomopathogenic nematode against major clinically relevant biofilm forming wound bacteria. Frontiers in Microbiology, 6, 876. Dirksen, P., Marsh, S. A., Braker, I., Heitland, N., Wagner, S., Nakad, R., Mader, S., Petersen, C., Kowallik, V., Rosenstiel, P., Félix, M.-A., & Schulenburg, H. (2016). The native microbiome of the nematode Caenorhabditis elegans: gateway to a new host-microbiome model. BMC Biology, 14(1), 38. Edgar, R. C. (2010). Search and clustering orders of magnitude faster than BLAST. Bioinformatics, 26(19), 2460-2461. Edgar, R. C. (2016). UNOISE2: improved error-correction for Illumina 16S and ITS amplicon sequencing. bioRxiv, 081257. Edgar, R. C., & Flyvbjerg, H. (2015). Error filtering, pair assembly and error correction for next-generation sequencing reads. Bioinformatics, 31(21), 3476-3482.

249

Edgar, R. C., Haas, B. J., Clemente, J. C., Quince, C., & Knight, R. (2011). UCHIME improves sensitivity and speed of chimera detection. Bioinformatics, 27(16), 2194-2200. Erandapurathukadumana Sreedharan, H., Harilal, C. C., & Pradeep, S. (2020). Response surface optimization of prodigiosin production by phthalate degrading Achromobacter denitrificans SP1 and exploring its antibacterial activity. Preparative Biochemistry & Biotechnology, 50(6), 1-8. Fu, J.-R., & Liu, Q.-Z. (2019). Evaluation and entomopathogenicity of gut bacteria associated with dauer juveniles of Oscheius chongmingensis (Nematoda: Rhabditidae). Microbiology Open, 8(9), e00823-e00823. Furlan, J. P. R., & Stehling, E. G. (2017). High-level of resistance to β-lactam and presence of β-lactamases encoding genes in Ochrobactrum sp. and Achromobacter sp. isolated from soil. Journal of Global Antimicrobial Resistance, 11, 133-137. Garcia-Gutierrez, E., Mayer, M. J., Cotter, P. D., & Narbad, A. (2019). Gut microbiota as a source of novel antimicrobials. Gut Microbes, 10(1), 1-21. Gbadegesin, R. A., Adesiyan, S. O., & Khan, F. A. (1993). Effects of seasonal changes on the population levels of some plant-parasitic nematodes in the plantations of Pinus caribaea and P. oocarpa in the savanna areas of Nigeria. Forest Ecology and Management, 60(3), 346-353. Guzman, L. M., Belin, D., Carson, M. J., & Beckwith, J. (1995). Tight regulation, modulation, and high-level expression by vectors containing the arabinose PBAD promoter. Journal of Bacteriology, 177(14), 4121-4130. Harrington, D. (2005). Linear rank tests in survival analysis. In Encyclopedia of Biostatistics. John Wiley & Sons, Ltd. 10.1002/0470011815.b2a11047. Hodgkin, J., Kuwabara, P. E., & Corneliussen, B. (2000). A novel bacterial pathogen, Microbacterium nematophilum, induces morphological change in the nematode C. elegans. Current Biology, 10(24), 1615-1618. Illumina. (2013). MiSeq system user guide. In Illumina Inc. FU Berlin. Jiang, H., & Wang, D. (2018). The microbial zoo in the C. elegans intestine: bacteria, fungi and viruses. Viruses, 10(2), 85. Kamada, N., Kim, Y.-G., Sham, H. P., Vallance, B. A., Puente, J. L., Martens, E. C., & Núñez, G. (2012). Regulated virulence controls the ability of a pathogen to compete with the gut microbiota. Science, 336(6086), 1325. Kim, Y., & Mylonakis, E. (2012). Caenorhabditis elegans immune conditioning with the probiotic bacterium Lactobacillus acidophilus strain NCFM enhances Gram- positive immune responses. Infection and Immunity, 80(7), 2500-2508. Kissoyan, K. A. B., Drechsler, M., Stange, E.-L., Zimmermann, J., Kaleta, C., Bode, H. B., & Dierking, K. (2019). Natural C. elegans microbiota protects against infection via production of a cyclic lipopeptide of the viscosin group. Current Biology, 29(6), 1030-1037.e1035. Klindworth, A., Pruesse, E., Schweer, T., Peplies, J., Quast, C., Horn, M., & Glöckner, F. O. (2013). Evaluation of general 16S ribosomal RNA gene PCR primers for classical and next-generation sequencing-based diversity studies. Nucleic Acids Research, 41(1), e1. Mantel, N. (1966). Evaluation of survival data and two new rank order statistics arising in its consideration. Cancer Chemotherapy Reports, 50(3), 163-170.

250

Martinez-Guryn, K., Hubert, N., Frazier, K., Urlass, S., Musch, M. W., Ojeda, P., Pierre, J. F., Miyoshi, J., Sontag, T. J., Cham, C. M., Reardon, C. A., Leone, V., & Chang, E. B. (2018). Small intestine microbiota regulate host digestive and absorptive adaptive responses to dietary lipids. Cell Host & Microbe, 23(4), 458-469. Miller, C. S., Handley, K. M., Wrighton, K. C., Frischkorn, K. R., Thomas, B. C., & Banfield, J. F. (2013). Short-read assembly of full-length 16S amplicons reveals bacterial diversity in subsurface sediments. PloS One, 8(2), e56018. Montalvo-Katz, S., Huang, H., Appel, M. D., Berg, M., & Shapira, M. (2013). Association with soil bacteria enhances p38-dependent infection resistance in Caenorhabditis elegans. Infection and Immunity, 81(2), 514-520. Nicholas, H. R., & Hodgkin, J. (2004). The ERK MAP Kinase cascade mediates tail swelling and a protective response to rectal infection in C. elegans. Current Biology, 14(14), 1256-1261. Nyholm, S. V., & Graf, J. (2012). Knowing your friends: invertebrate innate immunity fosters beneficial bacterial symbioses. Nature Reviews Microbiology, 10(12), 815-827. Oksanen, J., Blanchet, F., Kindt, R., Legendre, P., Minchin, P., O’Hara, R., Simpson, G., Solymos, P., Stevens, M., & Wagner, H. (2018). Package ‘vegan’—Community Ecology Package. 2019. Available at: https://cran.r- project.org/web/packages/vegan/vegan.pdf Oksanen, J., Blanchet, F. G., Kindt, R., Legendre, P., Minchin, P., O’hara, R., Simpson, G., Solymos, P., Stevens, M., & Wagner, H. (2017). Package ‘vegan’. 2016. Community ecology package, R. package version, 2.4-1. Ozkan, J., Willcox, M., Wemheuer, B., Wilcsek, G., Coroneo, M., & Thomas, T. (2019). Biogeography of the human ocular microbiota. The Ocular Surface, 17(1), 111- 118. Page, A. P., Roberts, M., Félix, M.-A., Pickard, D., Page, A., & Weir, W. (2019). The golden death bacillus Chryseobacterium nematophagum is a novel matrix digesting pathogen of nematodes. BMC Biology, 17(1), 10. Pukkila-Worley, R., & Ausubel, F. M. (2012). Immune defense mechanisms in the Caenorhabditis elegans intestinal epithelium. Current Opinion in Immunology, 24(1), 3-9. Qi, D., Wieneke, X., Tao, J., Zhou, X., & Desilva, U. (2018). Soil pH is the primary factor correlating with soil microbiome in karst rocky desertification regions in the Wushan County, Chongqing, China. Frontiers in Microbiology, 9, 1027. Rakoff-Nahoum, S., Paglino, J., Eslami-Varzaneh, F., Edberg, S., & Medzhitov, R. (2004). Recognition of commensal microflora by Toll-Like receptors is required for intestinal homeostasis. Cell, 118(2), 229-241. Raza, M. F., Wang, Y., Cai, Z., Bai, S., Yao, Z., Awan, U. A., Zhang, Z., Zheng, W., & Zhang, H. (2020). Gut microbiota promotes host resistance to low-temperature stress by stimulating its arginine and proline metabolism pathway in adult Bactrocera dorsalis. PLoS Pathogens, 16(4), e1008441. Revtovich, A. V., Lee, R., & Kirienko, N. V. (2019). Interplay between mitochondria and diet mediates pathogen and stress resistance in Caenorhabditis elegans. PLOS Genetics, 15(3), e1008011-e1008011. Rolhion, N., & Chassaing, B. (2016). When pathogenic bacteria meet the intestinal microbiota. Philosophical Transactions of the Royal Society B, Biological Science, 371(1707), 20150504. Sampson, T. R., & Mazmanian, S. K. (2015). Control of brain development, function, and behavior by the microbiome. Cell Host Microbe, 17(5), 565-576. 251

Samuel, B. S., Rowedder, H., Braendle, C., Félix, M.-A., & Ruvkun, G. (2016). Caenorhabditis elegans responses to bacteria from its natural habitats. Proceedings of the National Academy of Sciences of the USA, 113(27), E3941. Silva, Y. P., Bernardi, A., & Frozza, R. L. (2020). The role of short-chain fatty acids from gut microbiota in gut-brain communication. Frontiers in Endocrinology, 11, 25- 25. Stiernagle, T. (2006). Maintenance of C. elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_strainmaintain/strainmaintain.html. Stretton, S., Techkarnjanaruk, S., McLennan, A. M., & Goodman, A. E. (1998). Use of green fluorescent protein to tag and investigate gene expression in marine bacteria. Applied and Environmental Microbiology, 64(7), 2554-2559. Sulston, J. (2019). The power of C. elegans: A tribute to Sydney Brenner. Developmental Cell, 49, 496-498 Tanaka-Hino, M., Sagasti, A., Hisamoto, N., Kawasaki, M., Nakano, S., Ninomiya-Tsuji, J., Bargmann, C. I., & Matsumoto, K. (2002). SEK-1 MAPKK mediates Ca2+ signaling to determine neuronal asymmetric development in Caenorhabditis elegans. EMBO reports, 3(1), 56-62. Team, R. C. (2019). R: a language and environment for statistical computing. R Foundation for Statistical Computing website. Available at http://www.R- project.org/. Thursby, E., & Juge, N. (2017). Introduction to the human gut microbiota. The Biochemical Journal, 474(11), 1823-1836. Tilg, H., & Moschen, A. R. (2015). Food, immunity, and the microbiome. Gastroenterology, 148(6), 1107-1119. Van Voorhies, W. A., Fuchs, J., & Thomas, S. (2005). The longevity of Caenorhabditis elegans in soil. Biology Letters, 1(2), 247-249. Wang, Y., Naumann, U., Wright, S. T., & Warton, D. I. (2012). Mvabund–an R package for model‐based analysis of multivariate abundance data. Methods in Ecology and Evolution, 3(3), 471-474. Watson, E., Olin-Sandoval, V., Hoy, M. J., Li, C. H., Louisse, T., Yao, V., Mori, A., Holdorf, A. D., Troyanskaya, O. G., Ralser, M., & Walhout, A. J. (2016). Metabolic network rewiring of propionate flux compensates vitamin B12 deficiency in C. elegans. eLife, 5, e17670. Wemheuer, B., & Wemheuer, F. (2017). Assessing bacterial and fungal diversity in the plant endosphere. In W. R. Streit & R. Daniel (Eds.), Metagenomics: Methods and Protocols (pp. 75-84). New York : Springer. Wilkes Walburn, J., Wemheuer, B., Thomas, T., Copeland, E., O'Connor, W., Booth, M., Fielder, S., & Egan, S. (2019). Diet and diet-associated bacteria shape early microbiome development in Yellowtail Kingfish (Seriola lalandi). Microbial Biotechnology, 12(2), 275-288. Willemsen, L., Koetsier, M., Van Deventer, S., & Van Tol, E. (2003). Short chain fatty acids stimulate epithelial mucin 2 expression through differential effects on prostaglandin E1 and E2 production by intestinal myofibroblasts. Gut, 52(10), 1442-1447. Wollenberg, A. C., Jagdish, T., Slough, G., Hoinville, M. E., & Wollenberg, M. S. (2016). Death becomes them: bacterial community dynamics and stilbene antibiotic production in cadavers of Galleria mellonella killed by Heterorhabditis and Photorhabdus spp. Applied and Environmental Microbiology, 82(19), 5824-5837. 252

Yang, W., Petersen, C., Pees, B., Zimmermann, J., Waschina, S., Dirksen, P., Rosenstiel, P., Tholey, A., Leippe, M., Dierking, K., Kaleta, C., & Schulenburg, H. (2019). The inducible response of the nematode Caenorhabditis elegans to members of its natural microbiota across development and adult life. Frontiers in Microbiology, 10, 1793-1793. Zimmermann, J., Obeng, N., Yang, W., Pees, B., Petersen, C., Waschina, S., Kissoyan, K. A., Aidley, J., Hoeppner, M. P., & Bunk, B. (2020). The functional repertoire contained within the native microbiota of the model nematode Caenorhabditis elegans. The ISME Journal, 14(1), 26-38.

253

SUPPLEMENTARY MATERIALS

Table S4.1 Assessment on the bacterial community structure in SOIL and nematodes samples; C. elegans N2_UGM and N2_DGM

Amount of sample 3 (SOIL, N2_UGM, N2_DGM) Replicate per sample 5 Total sample 15 Total sequences 1,436 733 Total zOTUs 879

Table S4.2 Comparison of diversity and richness of bacterial communities in SOIL and N2_UGM microbial DNA samples

Unpaired t-test (two-tailed) SOIL vs N2_UGM df t p-value

Shannon diversity 8 6.407 0.0002 Richness 8 14.41 < 0.0001

254

Table S4.3 PERMANOVA based on Bray-Curtis (BC) dissimilarity measure for square- root transformed abundances of bacterial communities in SOIL and C. elegans N2_UGM. The p-values were calculated using 999 permutations under a residual model. Bold indicates statistical significance (at alpha = 0.05). Abbreviations: SOIL (the enriched soil supplemented with rotten fruit produce), N2_UGM (C. elegans N2 with Undefined Gut Microbiota).

PERMANOVA

Df SumsOfSqs MeanSqs F.Model R2 Pr(>F) Samples 1 0.74474 0.74474 75.544 0.90745 0.004 ** Replicate 1 0.00886 0.00886 0.899 0.01079 0.367 Treatment: 1 0.00794 0.00794 0.806 0.00968 0.363 Replicate Residuals 6 0.05915 0.00986 0.07207 Total 9 0.82070 1.00000

PERMDISP Df Sum Sq Mean Sq F N Pr(>F) Perm Groups 1 0.0052335 0.0052335 27.398 999 0.001 *** Residuals 8 0.0015281 0.000191 0.899 0.01079 0.367

255

Table S4.4 Relative abundance of zOTUs found to be significantly affected by the sample types; SOIL or N2_UGM microbial samples (ANOVA with p adjusted < 0.05). Differential abundance analysis was performed using Mvabund

List of SOIL N2_UGM Class Family Genus Species zOTUs Zotu901 0.01 0 Acidobacteriae UBA7541 UBA7541 UBA7541_sp002478115 Zotu497 0.02 0 Actinobacteria Cellulomonadaceae Cellulomonas Cellulomonas timonensis Zotu740 0.02 0 Actinobacteria Micromonosporaceae Micromonospora Micromonospora sp003013775 Mycolicibacterium Zotu934 0.01 0 Actinobacteria Mycobacteriaceae Mycolicibacterium thermoresistibile Zotu271 0 0.06 Actinobacteria Mycobacteriaceae Rhodococcus Rhodococcus hoagii Zotu23 1.68 0.04 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu529 0.03 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu681 0.01 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu835 0.01 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu297 0.03 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu224 0.07 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu182 0.09 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu64 0.43 0.01 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu68 0.38 0.01 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu62 0.45 0.01 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu96 0.28 0.01 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu138 0.12 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu339 0.04 0 Bacteroidia Chitinophagaceae Chitinophaga Chitinophaga jiangningensis Zotu225 0 0.08 Bacteroidia Chitinophagaceae NS-102 NS-102_sp003075435 Zotu378 0.03 0 Bacteroidia Flavobacteriaceae Flavobacterium Flavobacterium sp001429295 Zotu77 0.01 0.15 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium multivorum

256

Zotu47 0.02 0.58 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium multivorum Zotu847 0 0.03 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium multivorum Zotu13 0.17 2.31 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium multivorum Zotu21 0.13 2.25 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium sp001952815 Zotu29 0.08 1.56 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium sp002000245 Zotu22 0.1 1.4 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium sp002000245 Zotu45 0.05 0.78 Bacteroidia Sphingobacteriaceae Sphingobacterium Sphingobacterium sp002000245 Zotu915 0.01 0 Anaerolineae UBA4823 UTCFX2 UTCFX2_sp002050125 Zotu632 0.01 0 Sericytochromatia UBA7694 GCA-2770975 GCA-2770975_sp002783405 Zotu742 0.02 0 Bacilli Bacillaceae C Bacillus E Bacillus E coagulans Zotu914 0.01 0 Bacilli Bacillaceae H Bacillus C Bacillus C aryabhattai Zotu885 0.01 0 Bacilli Sporolactobacillaceae Sporolactobacillus inulinus Zotu276 0.06 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus inulinus Zotu43 0.74 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus inulinus Zotu234 0.12 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus inulinus Zotu684 0.09 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus inulinus Zotu867 0.01 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu464 0.03 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu506 0.02 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu396 0.02 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu212 0.07 0.01 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu250 0.06 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu350 0.03 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu590 0.1 0.01 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu769 0.07 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu567 0.05 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu151 0.14 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus

257

Zotu655 0.02 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus laevolacticus Zotu15 2.86 0.01 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu10 4.3 0.14 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu216 0.08 0.01 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu786 0.02 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu521 0.03 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu785 0.01 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu571 0.02 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu596 0.01 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu524 0.03 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu278 0.05 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu58 0.55 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu72 0.45 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu125 0.21 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu193 0.1 0 Bacilli Sporolactobacillaceae Sporolactobacillus Sporolactobacillus pectinivorans Zotu18 0.91 1.89 Bacilli Planococcaceae Rummeliibacillus Rummeliibacillus pycnus Zotu87 0.09 0.25 Bacilli Planococcaceae Rummeliibacillus Rummeliibacillus pycnus Zotu860 0.12 0.01 Bacilli Planococcaceae Rummeliibacillus Rummeliibacillus pycnus Zotu880 0.09 0 Bacilli Planococcaceae Rummeliibacillus Rummeliibacillus pycnus Zotu55 0.03 0.7 Bacilli Erysipelotrichaceae Dielma Dielma fastidiosa Zotu26 0.03 1.83 Bacilli Erysipelotrichaceae Dielma Dielma sp001305055 Zotu260 0.01 0.07 Bacilli Haloplasmataceae Haloplasma Haloplasma contractile Zotu451 0.02 0 Bacilli Lactobacillaceae Leuconostoc Leuconostoc citreum Zotu262 0.04 0 Bacilli Lactobacillaceae Leuconostoc Leuconostoc mesenteroides Zotu184 0.1 0 Bacilli Streptococcaceae Lactococcus Lactococcus lactis Zotu629 0.01 0 Clostridia UBA1242 UBA11490 UBA11490_sp003524095 Zotu486 0.02 0 Clostridia Acetivibrionaceae Ruminiclostridium Ruminiclostridium sp001717745

258

Zotu477 0.02 0 Clostridia Acetivibrionaceae Ruminiclostridium Ruminiclostridium thermocellum Zotu695 0.02 0 Clostridia Acetivibrionaceae Ruminiclostridium Ruminiclostridium thermocellum Ruminiclostridium A Zotu349 0.03 0.01 Clostridia Acetivibrionaceae Ruminiclostridium A cellobioparum Zotu518 0.02 0 Clostridia Acetivibrionaceae Ruminiclostridium A Ruminiclostridium A papyrosolvens Ruminiclostridium F Zotu731 0.02 0 Clostridia Acetivibrionaceae Ruminiclostridium F thermosuccinogenes Zotu284 0.06 0 Clostridia CAG-138 UBA3792 UBA3792_sp900318745 Zotu907 0.01 0 Clostridia CAG-74 OEMS01 OEMS01_sp900199405 Zotu950 0.01 0 Clostridia CAG-74 OEMS01 OEMS01_sp900199405 Zotu942 0.01 0 Clostridia Christensenellaceae Christensenella Christensenella minuta Zotu352 0.04 0 Clostridia GCA-900066905 GCA-900066905 GCA-900066905_sp900066905 Zotu754 0.01 0 Clostridia Caloramatoraceae Caloramator B Caloramator B quimbayensis Zotu511 0.01 0 Clostridia Caloramatoraceae Fervidicella Fervidicella metallireducens Zotu142 0.01 0.22 Clostridia Clostridiaceae Clostridium Clostridium saccharobutylicum Zotu282 0.01 0.07 Clostridia Clostridiaceae Clostridium Clostridium sartagoforme Zotu450 0.02 0 Clostridia Clostridiaceae Clostridium Clostridium sp002760435 Zotu109 0 0.26 Clostridia Clostridiaceae Clostridium AD Clostridium AD estertheticum Zotu333 0.03 0 Clostridia Clostridiaceae Clostridium AE Clostridium AE oryzae Zotu736 0.01 0 Clostridia Clostridiaceae Clostridium B Clostridium B ljungdahlii Zotu144 0.17 0 Clostridia Clostridiaceae Clostridium B Clostridium B ljungdahlii Zotu152 0.14 0.02 Clostridia Clostridiaceae Clostridium B Clostridium B luticellarii Zotu183 0.13 0.02 Clostridia Clostridiaceae Clostridium B Clostridium B luticellarii Zotu110 0.27 0 Clostridia Clostridiaceae Clostridium B Clostridium B luticellarii Zotu42 0.82 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu57 0.57 0.25 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu316 0.04 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum

259

Zotu949 0.01 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu123 0.25 0.03 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu248 0.09 0.01 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu48 0.76 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu86 0.34 0.01 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu130 0.18 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu167 0.12 0 Clostridia Clostridiaceae Clostridium B Clostridium B tyrobutyricum Zotu145 0.12 0 Clostridia Clostridiaceae Clostridium C Clostridium C acetireducens Zotu340 0.03 0 Clostridia Clostridiaceae Clostridium F Clostridium F sp001276215 Zotu545 0.02 0 Clostridia Clostridiaceae Clostridium I Clostridium I pasteurianum_A Zotu119 0.19 0.01 Clostridia Clostridiaceae Clostridium S Clostridium S pasteurianum Zotu63 0.03 0.64 Clostridia Clostridiaceae Clostridium T Clostridium T intestinale Zotu446 0 0.02 Clostridia Clostridiaceae F0540 F0540_sp000466585 Zotu672 0.01 0 Clostridia Clostridiaceae F0540 F0540_sp000466585 Zotu245 0.01 0.07 Clostridia Clostridiaceae F0540 F0540_sp000466585 Zotu74 0.03 0.59 Clostridia Clostridiaceae F0540 F0540_sp000466585 Zotu501 0.01 0 Clostridia Anaerotignaceae Anaerotignum Anaerotignum propionicum Zotu555 0.01 0 Clostridia Cellulosilyticaceae Epulopiscium B Epulopiscium B sp000171335 Zotu279 0 0.06 Clostridia Cellulosilyticaceae Epulopiscium B Epulopiscium B sp000171335 Zotu403 0.03 0 Clostridia Acetivibrio A Acetivibrio A ethanolgignens Zotu239 0.08 0.01 Clostridia Lachnospiraceae Acetivibrio A Acetivibrio A ethanolgignens Zotu466 0.03 0 Clostridia Lachnospiraceae Acetivibrio A Acetivibrio A ethanolgignens Zotu420 0.02 0 Clostridia Lachnospiraceae Acetivibrio A Acetivibrio A ethanolgignens Zotu158 0.12 0 Clostridia Lachnospiraceae Acetivibrio A Acetivibrio A ethanolgignens Zotu149 0.02 0.11 Clostridia Lachnospiraceae Anaerocolumna Anaerocolumna jejuensis Zotu108 0.05 0.27 Clostridia Lachnospiraceae Anaerocolumna Anaerocolumna jejuensis Zotu298 0.01 0.05 Clostridia Lachnospiraceae Anaerocolumna Anaerocolumna sp000702945

260

Zotu83 0.31 0.06 Clostridia Lachnospiraceae CAG-45 CAG-45_sp900066395 Zotu589 0.02 0 Clostridia Lachnospiraceae CAG-603 CAG-603_sp900066105 Zotu503 0.04 0.24 Clostridia Lachnospiraceae Clostridium M Clostridium M sp000155435 Zotu252 0.05 0.01 Clostridia Lachnospiraceae Clostridium N Clostridium N fimetarium Zotu217 0.08 0.01 Clostridia Lachnospiraceae Clostridium N Clostridium N fimetarium Zotu670 0.01 0 Clostridia Lachnospiraceae Clostridium Q Clostridium Q sp003024715 Zotu203 0.03 0.08 Clostridia Lachnospiraceae Hespellia Hespellia stercorisuis Zotu202 0.09 0 Clostridia Lachnospiraceae Kineothrix Kineothrix alysoides Lachnoclostridium Zotu56 0.11 0.66 Clostridia Lachnospiraceae Lachnoclostridium phytofermentans A Zotu699 0 0.01 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu562 0.01 0 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu320 0 0.06 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu310 0 0.06 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu391 0.02 0 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu405 0.03 0 Clostridia Lachnospiraceae Lachnoclostridium Lachnoclostridium sp900078195 Zotu353 0.02 0 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu410 0.03 0 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu228 0.07 0.18 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu423 0.02 0 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu494 0.02 0 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu139 0.14 0.01 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu374 0.06 0 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu16 0.75 2.65 Clostridia Acutalibacteraceae MS4 MS4 sp000752215 Zotu592 0.02 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu556 0.01 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu361 0.03 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555

261

Zotu334 0.03 0.01 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu861 0.01 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu840 0.01 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu619 0.01 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu443 0.02 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu669 0.01 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu150 0.16 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu312 0.04 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu31 1.43 0.02 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu272 0.06 0 Clostridia Acutalibacteraceae UBA1033 UBA1033 sp001695555 Zotu277 0.05 0.01 Clostridia Acutalibacteraceae UBA4871 UBA4871_sp002119605 Zotu343 0.03 0 Clostridia Acutalibacteraceae UBA4871 UBA4871_sp002119605 Zotu227 0.06 0 Clostridia Acutalibacteraceae UBA4871 UBA4871_sp002119605 Zotu308 0.04 0 Clostridia Acutalibacteraceae UBA4871 UBA4871_sp002119605 Zotu243 0.06 0.01 Clostridia Ethanoligenenaceae Ethanoligenens Ethanoligenens harbinense Zotu408 0.03 0 Clostridia Ethanoligenenaceae Ethanoligenens Ethanoligenens harbinense Zotu599 0.02 0 Clostridia Ethanoligenenaceae Ethanoligenens Ethanoligenens harbinense Zotu539 0.02 0 Clostridia Ethanoligenenaceae Ethanoligenens Ethanoligenens harbinense Zotu311 0.05 0 Clostridia Ethanoligenenaceae Ethanoligenens Ethanoligenens harbinense Zotu791 0.01 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu936 0.01 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu214 0.06 0.03 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu286 0.04 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu76 0.3 0.05 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu687 0.01 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu258 0.06 0.01 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu101 0.3 0.02 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi

262

Zotu164 0.09 0.02 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu79 0.37 0.03 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu531 0.02 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu99 0.23 0.05 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu595 0.02 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu231 0.07 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu50 0.6 0.05 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu116 0.21 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu154 0.13 0 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu222 0.07 0.01 Clostridia Ethanoligenenaceae Ruminiclostridium D Ruminiclostridium D cellulosi Zotu487 0.01 0 Clostridia Oscillospiraceae Intestinimonas Intestinimonas massiliensis Zotu548 0.01 0 Clostridia Oscillospiraceae Intestinimonas Intestinimonas massiliensis Zotu513 0.02 0 Clostridia Oscillospiraceae Pseudoflavonifractor Pseudoflavonifractor sp900079765 Zotu172 0.1 0.01 Clostridia Oscillospiraceae UBA2658 UBA2658_sp002841545 Zotu925 0.02 0 Clostridia Oscillospiraceae UBA9475 UBA9475_sp002161675 Zotu230 0.07 0 Clostridia Ruminococcaceae Anaerotruncus Anaerotruncus sp000403395 Zotu232 0 0.09 Clostridia Ruminococcaceae D5 D5_sp900113995 Zotu836 0.01 0 Clostridia Ruminococcaceae Massilimaliae Massilimaliae timonensis Zotu140 0.14 0.03 Clostridia Ruminococcaceae Phocea Phocea massiliensis Zotu347 0.02 0 Clostridia Ruminococcaceae Phocea Phocea massiliensis Zotu709 0.01 0 Clostridia Ruminococcaceae Provencibacterium Provencibacterium massiliense Zotu406 0.02 0 Clostridia Ruminococcaceae Ruminococcus Ruminococcus flavefaciens F Zotu395 0.02 0 Clostridia Ruminococcaceae Ruminococcus_D Ruminococcus D albus C Zotu508 0.02 0 Clostridia Ruminococcaceae Ruminococcus_D Ruminococcus D albus C Zotu472 0.02 0 Clostridia Anaerovoracaceae S5-A14a S5-A14a_sp900112915 Zotu325 0.05 0 Clostridia Anaerovoracaceae S5-A14a S5-A14a_sp900112915 Zotu169 0.03 0.09 Clostridia Dethiosulfatibacteraceae Dethiosulfatibacter Dethiosulfatibacter aminovorans

263

Zotu903 0.01 0 Desulfitobacteriia Desulfitobacteriaceae Desulfitobacterium Desulfitobacterium dehalogenans Zotu470 0.02 0 Desulfitobacteriia Desulfitobacteriaceae Desulfitobacterium Desulfitobacterium hafniense Zotu383 0.03 0 Desulfitobacteriia Desulfitobacteriaceae Desulfitobacterium Desulfitobacterium hafniense Zotu766 0.02 0 Desulfitobacteriia Desulfitobacteriaceae Desulfosporosinus Desulfosporosinus orientis Zotu615 0.02 0 Negativicutes Dendrosporobacteraceae Dendrosporobacter Dendrosporobacter quercicolus Zotu500 0.02 0 Negativicutes Acetonemaceae Acetonema Acetonema longum Zotu540 0.02 0 Negativicutes Sporomusaceae Sporomusa Sporomusa sphaeroides Zotu828 0.01 0 Negativicutes Sporomusaceae Sporomusa Sporomusa sphaeroides Zotu608 0.02 0 Negativicutes Sporomusaceae Sporomusa Sporomusa sphaeroides Zotu397 0.04 0 Negativicutes Sporomusaceae Sporomusa Sporomusa sphaeroides Zotu845 0.01 0 Negativicutes Sporomusaceae Sporomusa_C Sporomusa C ovata Zotu268 0.06 0 Negativicutes Sporomusaceae Sporomusa_C Sporomusa C ovata Zotu855 0.02 0 Negativicutes Sporomusaceae Sporomusa_C Sporomusa C silvacetica Zotu438 0.03 0 Negativicutes Sporomusaceae Sporomusa_C Sporomusa C silvacetica Zotu326 0.04 0 Negativicutes Sporomusaceae Sporomusa_C Sporomusa C silvacetica Zotu367 0.02 0 DTU015 D2 DTU015 DTU015_sp001513185 Zotu422 0 0.03 Bacilli_A Brevibacillaceae Brevibacillus Brevibacillus choshinensis Zotu163 0.01 0.16 Bacilli_A Brevibacillaceae Brevibacillus Brevibacillus choshinensis Zotu394 0 0.03 Bacilli_A Brevibacillaceae Brevibacillus Brevibacillus choshinensis Zotu434 0.02 0 Bacilli_A Brevibacillaceae Brevibacillus Brevibacillus panacihumi Zotu404 0 0.03 Bacilli_A Paenibacillaceae Cohnella Cohnella kolymensis Zotu331 0 0.05 Bacilli_A Paenibacillaceae Cohnella Cohnella panacarvi Zotu285 0 0.05 Bacilli_A Paenibacillaceae Cohnella Cohnella panacarvi Zotu426 0.02 0 Bacilli_A Paenibacillaceae Cohnella Cohnella sp900169535 Zotu175 0 0.14 Bacilli_A Paenibacillaceae DXL2 DXL2_sp003217775 Zotu259 0 0.08 Bacilli_A Paenibacillaceae DXL2 DXL2_sp003217775 Zotu432 0 0.03 Bacilli_A Paenibacillaceae Gorillibacterium Gorillibacterium timonense

264

Zotu532 0 0.02 Bacilli_A Paenibacillaceae Gorillibacterium Gorillibacterium timonense Zotu937 0.01 0 Bacilli_A Paenibacillaceae Paenibacillus Paenibacillus durus B Zotu510 0 0.03 Bacilli_A Paenibacillaceae Paenibacillus Paenibacillus sp000787385 Zotu453 0.01 0.03 Bacilli_A Paenibacillaceae Paenibacillus G Paenibacillus G mucilaginosus Zotu360 0 0.04 Bacilli_A Paenibacillaceae Paenibacillus G Paenibacillus G mucilaginosus Zotu98 0 0.38 Bacilli_A Paenibacillaceae Paenibacillus G Paenibacillus G mucilaginosus Zotu187 0 0.16 Bacilli_A Paenibacillaceae Paenibacillus G Paenibacillus G mucilaginosus Zotu192 0 0.14 Bacilli_A Paenibacillaceae Paenibacillus S Paenibacillus S rigui Zotu471 0 0.02 Bacilli_A Paenibacillaceae UNC496MF UNC496MF_sp900116125 Zotu324 0 0.05 Bacilli_A Paenibacillaceae UNC496MF UNC496MF_sp900116125 Zotu305 0 0.06 Bacilli_A Paenibacillaceae UNC496MF UNC496MF_sp900116125 Zotu467 0 0.03 Bacilli_A Paenibacillaceae UNC496MF UNC496MF_sp900116125 Zotu418 0.02 0 Alicyclobacillia Tumebacillaceae Tumebacillus A Tumebacillus A avium Zotu314 0.03 0 Alicyclobacillia Tumebacillaceae Tumebacillus A Tumebacillus A avium Zotu594 0.01 0 Gemmatimonadetes Gemmatimonadaceae SCN-70-22 SCN-70-22_sp001724275 Zotu88 0.34 0 Myxococcia Vulgatibacteraceae Vulgatibacter Vulgatibacter incomptus Zotu572 0.03 0 Polyangia Polyangiaceae Labilithrix Labilithrix luteola Zotu593 0.01 0 Phycisphaerae UBA1161 UBA2421 UBA2421 sp002343075 Zotu480 0.02 0 Alphaproteobacteria Acetobacteraceae Acetobacter Acetobacter cerevisiae Zotu656 0.01 0 Alphaproteobacteria Acetobacteraceae Acetobacter Acetobacter indonesiensis Zotu4 14.35 4.69 Alphaproteobacteria Acetobacteraceae Acetobacter Acetobacter papayae Zotu597 0.02 0 Alphaproteobacteria Acetobacteraceae Acetobacter Acetobacter persici Zotu218 0.07 0 Alphaproteobacteria Acetobacteraceae Acetobacter Acetobacter persici Zotu8 1.33 3.98 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus Zotu17 2.08 0.2 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus Zotu585 0.1 0.01 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus Zotu44 0.7 0.03 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus

265

Zotu70 0.46 0.05 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus Zotu117 0.26 0.01 Alphaproteobacteria Acetobacteraceae Gluconacetobacter Gluconacetobacter diazotrophicus Zotu381 0.03 0 Alphaproteobacteria Azospirillaceae Azospirillum Azospirillum brasilense A Zotu358 0.03 0 Alphaproteobacteria Caulobacteraceae Caulobacter Caulobacter mirabilis Zotu579 0.02 0 Alphaproteobacteria Caulobacteraceae Phenylobacterium Phenylobacterium sp001557235 Zotu546 0.02 0 Alphaproteobacteria Beijerinckiaceae Microvirga Microvirga sp003151255 Zotu244 0.06 0.01 Alphaproteobacteria Devosiaceae Devosia Devosia sp001185205 Zotu112 0.18 0.03 Alphaproteobacteria Devosiaceae Devosia Devosia sp001185205 Zotu874 0.01 0 Alphaproteobacteria Hyphomicrobiaceae Hyphomicrobium A Hyphomicrobium A sp900117445 Zotu251 0.01 0.09 Alphaproteobacteria Kaistiaceae Kaistia Kaistia granuli Zotu205 0.08 0 Alphaproteobacteria Kaistiaceae Kaistia Kaistia granuli Zotu254 0.05 0 Alphaproteobacteria Rhizobiaceae Leaf454 Leaf454_sp001425485 Zotu12 0 0.04 Alphaproteobacteria Rhizobiaceae Ochrobactrum Ochrobactrum sp900470195 Zotu3 0.06 1.07 Alphaproteobacteria Rhizobiaceae Ochrobactrum_A Ochrobactrum A thiophenivorans Zotu180 0.07 0.04 Alphaproteobacteria Rhizobiaceae Paramesorhizobium Paramesorhizobium sp002980495 Zotu9 0.14 5.12 Alphaproteobacteria Rhizobiaceae Pararhizobium Pararhizobium sp001424985 Zotu66 0.04 0.57 Alphaproteobacteria Rhizobiaceae Sinorhizobium Sinorhizobium medicae Zotu591 0.01 0 Alphaproteobacteria Xanthobacteraceae Afipia Afipia sp001425835 Zotu30 0.02 1.63 Alphaproteobacteria Sphingomonadaceae Novosphingobium Novosphingobium pentaromativorans Zotu132 0 0.18 Alphaproteobacteria Sphingomonadaceae Novosphingobium Novosphingobium sp900113255 Zotu52 0.02 0.76 Alphaproteobacteria Sphingomonadaceae Novosphingobium Novosphingobium sp900113255 Zotu95 0.03 0.33 Alphaproteobacteria Sphingomonadaceae Sphingobium Sphingobium sp001421665 Zotu2 0.13 1.28 Gammaproteobacteria Burkholderiaceae Achromobacter Achromobacter denitrificans Zotu7 0.04 0.38 Gammaproteobacteria Burkholderiaceae Achromobacter Achromobacter denitrificans Zotu25 0 0.21 Gammaproteobacteria Burkholderiaceae Achromobacter Achromobacter marplatensis Zotu61 0.07 0.47 Gammaproteobacteria Burkholderiaceae Achromobacter Achromobacter mucicolens Zotu59 0 0.2 Gammaproteobacteria Burkholderiaceae Achromobacter Achromobacter ruhlandii

266

Zotu315 0.04 0 Gammaproteobacteria Burkholderiaceae Acidovorax_A Acidovorax A wautersii Zotu28 0.19 1.4 Gammaproteobacteria Burkholderiaceae Comamonas Comamonas aquatica Zotu527 0.02 0 Gammaproteobacteria Burkholderiaceae Paraburkholderia_B Paraburkholderia B tropica Zotu554 0.01 0 Gammaproteobacteria Burkholderiaceae Ralstonia Ralstonia solanacearum Zotu653 0.01 0 Gammaproteobacteria Enterobacteriaceae Citrobacter Citrobacter portucalensis Zotu304 0.05 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter bugandensis Zotu146 0.12 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cancerogenus A Zotu89 0.32 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cancerogenus A Zotu38 1.08 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cloacae I Zotu329 0.04 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cloacae Zotu114 0.2 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cloacae Zotu181 0.09 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter cloacae Zotu118 0.18 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter kobei Zotu155 0.12 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter kobei Zotu49 0.63 0 Gammaproteobacteria Enterobacteriaceae Enterobacter Enterobacter sp000568095 Zotu113 0.22 0 Gammaproteobacteria Enterobacteriaceae Enterobacter D Enterobacter D kobei Zotu335 0.03 0 Gammaproteobacteria Enterobacteriaceae Leclercia Leclercia adecarboxylata A Zotu385 0.03 0 Gammaproteobacteria Enterobacteriaceae Lelliottia Lelliottia jeotgali Zotu748 0.01 0 Gammaproteobacteria Enterobacteriaceae Lelliottia Lelliottia lapagei Zotu553 0.02 0 Gammaproteobacteria Enterobacteriaceae Pantoea Pantoea ludwigii B Zotu215 0.07 0 Gammaproteobacteria Enterobacteriaceae Pantoea Pantoea ludwigii B Zotu412 0.03 0 Gammaproteobacteria Enterobacteriaceae Pantoea Pantoea ludwigii B Zotu481 0.03 0 Gammaproteobacteria Enterobacteriaceae Pantoea Pantoea ludwigii B Zotu376 0.04 0 Gammaproteobacteria Enterobacteriaceae Serratia Serratia marcescens Zotu692 0.01 0 Gammaproteobacteria Enterobacteriaceae Serratia Serratia ureilytica Zotu659 0.01 0 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E monteilii C Zotu354 0.03 0 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E plecoglossicida

267

Zotu241 0.06 0 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E putida E Zotu442 0.03 0 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E putida_J Zotu5 10.13 5.51 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E sp000418555 Zotu288 0.06 0 Gammaproteobacteria Pseudomonadaceae Pseudomonas_E Pseudomonas E sp002113165 Zotu624 0.01 0 Gammaproteobacteria Steroidobacteraceae Steroidobacter Steroidobacter denitrificans Zotu811 0.01 0 Verrucomicrobiae UBA10450 AV55 AV55_sp003219415 Zotu476 0.02 0 Verrucomicrobiae UBA10450 AV55 AV55_sp003219415 Zotu178 0.09 0 Verrucomicrobiae UBA10450 AV55 AV55_sp003219415 Zotu263 0.05 0 Verrucomicrobiae Akkermansiaceae Haloferula Haloferula sp000739615

268

Table S4.5 C. elegans N2 monoxenic and N2_UGM survival on E. coli strains HP1, HG8 and the non-toxic control BD24

Overall test; Log-rank (Mantel-Cox) C. elegans N2 and N2_UGM survival on E. df = Chi square = p < 0.0001 coli HP1, HG8 and BD24 strains 5 40.83 One-way ANOVA C. elegans N2 and N2_UG survival on E. df = F = 477.5 p < 0.0001 coli HP1, HG8 and BD24 strains 5 Tukey's test Test strain Day q df p-value

(N2_UGM / HP1) vs. (N2_UGM / HG8) 2 2.24 12 0.6221 (N2_UGM / HP1) vs. (N2_UGM / BD24) 2 5.44 12 0.022 (N2_UGM / HP1) vs. (N2 / HP1) 2 5.45 12 0.0216 (N2_UGM / HP1) vs. (N2 / HG8) 2 0.87 12 0.9874 (N2_UGM / HP1) vs. (N2 / BD24) 2 54.48 12 < 0.0001 (N2_UGM / HG8) vs. (N2_UGM / BD24) 2 7.68 12 0.0016 (N2_UGM / HG8) vs. (N2 / HP1) 2 7.70 12 0.0016 (N2_UGM / HG8) vs. (N2 / HG8) 2 1.37 12 0.92 (N2_UGM / HG8) vs. (N2 / BD24) 2 56.72 12 < 0.0001 (N2_UGM / BD24) vs. (N2 / HP1) 2 0.01 12 > 0.9999 (N2_UGM / BD24) vs. (N2 / HG8) 2 6.31 12 0.0078 (N2_UGM / BD24) vs. (N2 / BD24) 2 49.04 12 < 0.0001 (N2 / HP1) vs. (N2 / HG8) 2 6.33 12 0.0076 (N2 / HP1) vs. (N2 / BD24) 2 49.03 12 < 0.0001 (N2 / HG8) vs. (N2 / BD24) 2 55.36 12 < 0.0001

269

Table S4.6 The difference of bacterial taxonomy abundance between the starting bacterial communities and the defined gut microbiota of C. elegans N2_DGM at the genus level

Two-way ANOVA

The difference of bacterial taxonomy F = abundance between the starting bacterial df = 8 p < 0.0001 63.67 communities and N2_DGM nematodes gut microbiota Sidak's test COMPARISON AT THE GENUS LEVEL t df p-value

Ochrobactrum 0.08682 36 > 0.9999 Citrobacter 9.842 36 < 0.0001 Achromobacter 16.07 36 < 0.0001 Sphingobacterium 7.098 36 < 0.0001 Pseudomonas 3.197 36 0.0257 Bosea 2.257 36 0.241 Leuconostoc 2.801 36 0.071 Stenotrophomonas 8.987 36 < 0.0001 Paramesorhizobium 0.2294 36 > 0.9999

270

Table S4.7 Survival of C. elegans N2 with a defined gut microbiota (N2_DGM) upon exposure to the Nkp-1 expressing E. coli clones

Overall test; Log-rank (Mantel-Cox)

C. elegans N2 and N2_DG survival on E. coli df = Chi square p < 0.0001 Nkp-1, HG8 and BD24 strains 5 = 58.73

One-way ANOVA C. elegans N2 and N2_DG survival on E. coli df = F = 330.6 p < 0.0001 Nkp-1, HG8 and BD24 strains 5

Tukey's test Test strain Day q df p-value (N2_DGM / HP1) vs. (N2_DGM / HG8) 2 13.34 12 < 0.0001 (N2_DGM / HP1) vs. (N2_DGM / BD24) 2 23.52 12 < 0.0001 (N2_DGM / HP1) vs. (N2 / HP1) 2 11.27 12 < 0.0001 (N2_DGM / HP1) vs. (N2 / HG8) 2 15.76 12 < 0.0001 (N2_DGM / HP1) vs. (N2 / BD24) 2 23.52 12 < 0.0001 (N2_DGM / HG8) vs. (N2_DGM / BD24) 2 36.87 12 < 0.0001 (N2_DGM / HG8) vs. (N2 / HP1) 2 2.07 12 0.6916 (N2_DGM / HG8) vs. (N2 / HG8) 2 2.42 12 0.5489 (N2_DGM / HG8) vs. (N2 / BD24) 2 36.87 12 < 0.0001 (N2_DGM / BD24) vs. (N2 / HP1) 2 34.80 12 < 0.0001 (N2_DGM / BD24) vs. (N2 / HG8) 2 39.29 12 < 0.0001 (N2_DGM / BD24) vs. (N2 / BD24) 2 0.00 12 > 0.9999 (N2 / HP1) vs. (N2 / HG8) 2 4.49 12 0.0678 (N2 / HP1) vs. (N2 / BD24) 2 34.80 12 < 0.0001 (N2 / HG8) vs. (N2 / BD24) 2 39.29 12 < 0.0001

271

Table S4.8 Comparison of fluorescence intensity of GFP-tagged E. coli clones colonising C. elegans monoxenic N2 and N2_DGM with defined gut microbiota

One-way ANOVA Comparison of E. coli HP1::GFP, HG8::GFP and F = p < BD24::GFP in C. elegans N2 monoxenic and df = 5 43.82 0.0001 N2_DGM (24 hours)

Comparison of E. coli HP1::GFP, HG8::GFP and F = p < BD24::GFP in C. elegans N2 monoxenic and df = 5 15.28 0.0001 N2_DGM (48 hours)

Comparison of E. coli HP1::GFP, HG8::GFP and F = p < BD24::GFP in C. elegans N2 monoxenic and df = 5 20.36 0.0001 N2_DGM (72 hours)

Sidak's test Time Test strain t df p-value (hours)

(N2_DGM / HP1::GFP) vs. (N2_DGM / BD24::GFP) 24 2.23 12 0.0454 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 24 1.09 12 0.2956 (N2_DGM / HG8::GFP) vs. (N2_DGM / BD24::GFP) 24 6.95 12 < 0.0001 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 24 4.99 12 0.0003 (N2_DGM / BD24::GFP) vs. (N2 / BD24::GFP) 24 0.26 12 0.7973

(N2_DGM / HP1::GFP) vs. (N2_DGM / BD24::GFP) 48 2.08 12 0.0592 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 48 1.03 12 0.3222 (N2_DGM / HG8::GFP) vs. (N2_DGM / BD24::GFP) 48 5.04 12 0.0003 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 48 1.60 12 0.1358 (N2_DGM / BD24::GFP) vs. (N2 vs BD24::GFP) 48 0.34 12 0.7403

(N2_DGM / HP1::GFP) vs. (N2_DGM / BD24::GFP) 72 2.62 12 0.0224 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 72 2.71 12 0.0188 (N2_DGM / HG8::GFP) vs. (N2_DGM / BD24::GFP) 72 3.25 12 0.0070 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 72 5.65 12 0.0001 (N2_DGM / BD24::GFP) vs. (N2 / BD24::GFP) 72 0.16 12 0.8744

272

Table S4.9 Proportion of morphological changes observed on N2 monoxenic and N2_DGM nematodes following exposure to GFP-tagged E. coli strains

Two-way ANOVA C. elegans N2 monoxenic and N2_DGM with internal df = 10 F = 37.23 p < 0.0001 hatching C. elegans N2 monoxenic and N2_DGM with pharynx df = 10 F = 5.006 p = 0.0006 distortion C. elegans N2 monoxenic and N2_DGM with internal df = 10 F = 64.85 p < 0.0001 organ damage C. elegans N2 monoxenic and N2_DGM with df = 10 F = 28.72 p < 0.0001 deformed anal region (dar) Tukey's test Time p- Test strain q df (hours) value C. elegans N2 monoxenic and N2_DGM with internal hatching (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 24 15.56 2 0.0289 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 24 12.00 4 0.0063 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 48 18.38 2 0.0208 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 48 1.90 2.94 0.7587 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 72 22.63 2 0.0141 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 72 3.00 4 0.4196 C. elegans N2 monoxenic and N2_DGM with pharynx distortion (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 24 14.14 2 0.0348 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 24 1.73 4 0.8094 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 48 1.27 2.94 0.9251 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 48 1.00 4 0.971 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 72 11.00 4 0.0087 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 72 1.41 2 0.8866 C. elegans N2 monoxenic and N2_DGM with internal organ damage (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 24 5.66 2 0.1919 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 24 15.56 2 0.0289 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 48 2.83 2 0.5302 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 48 1.41 2 0.8866 (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 72 17.68 2 0.0224 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 72 none none none C. elegans N2 monoxenic and N2_DGM with deformed anal region (dar) (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 24 5 2.56 0.1914 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 24 none none none (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 48 8.22 2.94 0.0489 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 48 none none none (N2_DGM / HP1::GFP) vs. (N2 / HP1::GFP) 72 7.78 3.2 0.0482 (N2_DGM / HG8::GFP) vs. (N2 / HG8::GFP) 72 none none none

273

Figure S4.1 Rarefaction curve (A) before and (B) after normalisation. Number of subsampling sequences: 41,027. The calculated rarefaction curves based on the rarefied and unrarefied data as well as Good’s coverage result of 99.24±2.31% indicated that the majority of the bacterial community in each sample was recovered by the surveying effort.

274

Figure S4.2 NMDS ordination of the SOIL and C. elegans N2_UGM bacterial communities using Bray-Curtis dissimilarity of transformed rarefied zOTU data (stress: 9.096333e-05).

275

(%)

Relative abundance

Figure S4.3 Taxonomic profiles of bacterial communities shown at the class level of all SOIL and N2_UGM samples. Bacterial communities with a relative abundance less than 1% are grouped with ‘Other’

276

(%)

Relative abundance

Figure S4.4 Taxonomic profiles of bacterial communities shown at the family level of all SOIL and N2_UGM samples. Bacterial communities with a relative abundance less than 1% are grouped with ‘Other’

277

A p < 0.0001

100

)

e

u

%

s

( s

i

n

t

o

e

i

h

g

t

e

r o

t

x

e

n

g

y

r

a

a

m h

the to tissue damage

a 50

p

d

e

h

h

t

t

i

g

w

n

i

e

d

d

n

o

t

u

a o

r

m

r

e

u

surrounding the pharynx pharynx regionsurrounding the(%)

s

Nematodes with Nematodes N 0 N2 N2_UGM

B 100

) %

( p = 0.0004

e

h

n

t

i

i

t

s

w

(%)

degenerated degenerated

e

t

e

n d

i

o 50

t

d

a

e

t

m

a

r

e intestine

e

N

n

e

g

e d

with Nematodes

0 N2 N2_UGM

Figure S4.5 Microscopic examination of C. elegans N2_UGM after the establishment of an undefined gut microbiota from the enriched SOIL. C. elegans N2 monoxenic were examined as the control. Nematode morphology was inspected using the Fluorescence Microscope Olympus BLX-61 under the DIC filter. Distinct morphological changes were observed in C. elegans N2_UGM, i.e. (A) damage to the tissue surrounding the pharynx region and (B) degenerated intestine. Those morphological phenotypes were not detected on C. elegans N2. Graphical summaries represent the average percentage of the visualised nematodes from triplicate samples ± standard error.

278

Eggs

Pharynx

Damage

Terminal pharyngeal bulb

Figure S4.6 Morphology of C. elegans N2 after the establishment of an undefined gut microbiota (UGM). The morphology of N2 monoxenic nematodes were also compared as a control and the animals show no changes of physical morphology (A-C). In contrast, the N2_UGM nematodes showed damage to the tissue surrounding the pharynx (D) and degenerated intestine (F), both are indicated by the white arrows. An in-depth visualisation of damages surrounding the nematode pharynx (G) C. elegans images were acquired using the DIC filter under 40x (A-F) or 100x (G) magnifications. Scale bars indicate 20 µm (A-F) or 10 µm (G).

279

Figure S4.7 E. coli HP1::GFP, HG8::GFP and non-toxic control BD24::GFP bacterial colonisation in C. elegans N2. The assayed C. elegans images were acquired under the 10x magnification using the Fluorescence Microscope Olympus BLX-61 equipped with the DIC and FITC filter. Scale bars indicates 100 µm. Bacterial colonisation was indicated by the green fluorescence emitted by the GFP-tagged bacterial cells. 280

Figure S4.8 E. coli HP1::GFP, HG8::GFP and non-toxic control BD24::GFP bacterial colonisation in in C. elegans N2_DGM. The assayed C. elegans images were acquired under the 10x magnification using the Fluorescence Microscope Olympus BLX-61 equipped with the DIC and FITC filter. Scale bars indicates 100 µm. Bacterial colonisation was indicated by the green fluorescence emitted by the GFP-tagged bacterial cells. 281

CHAPTER 5

General Discussion

282

5.1 Climate change; an inevitable factor that escalates parasitic nematode infection

The emergence of drug resistance among parasitic nematodes with an unpredictable infection trend is alarming (Chapter 1). In addition to prolonged treatment and incorrect drug dosage, climate change that contributes to increasing temperature, changing of weather, moisture and rainfall is also believed to be another prominent factor that escalates the prevalence of parasitic nematode diseases (Genchi et al., 2009; Fox et al., 2015; Verschave et al., 2016; McIntyre et al., 2017; Morgan et al., 2019).

Increasing temperature has resulted in dramatic changes in the development, prevalence and survival of free-living parasitic nematode such as Haemonchus contortus (Fox et al., 2015). In addition, changing of weather patterns e.g. increase of rainfall may enhance vectors distribution (such as for snails) and extends the period of parasitic nematode/helminths transmission or infectivity in ruminants or humans (Short et al., 2017; Caminade et al., 2019). For example, the rising schistosomiasis outbreaks in Ghana have been attributed to the changes of total rainfall that increased snail distribution and indirectly facilitated parasitic nematode transmission (Codjoe and Larbi, 2016). Rising temperatures also accelerate the development of blackfly (Simulium spp.) and its associated parasitic nematode Onchocerca volvulus which can cause onchocerciasis disease (river blindness) in humans. Even worse, heat stress resulting from climate change may profoundly affect the livestock’s fitness through reduction of innate immunity and disruption of cellular metabolism thus resulting in increased susceptibility to parasite infection (Lacetera, 2018).

Given the likely increase in infection rate of parasitic nematodes as a result of climate change (Fox et al., 2015), finding new antinematode therapeutic agents to combat those parasite infestations has become a global urgency (Garcia-Bustos et al., 2019). In this chapter, I will discuss how the findings in this thesis can contribute to the development of novel antinematode drug candidates to combat parasitic nematode infections (section 5.2) as well as providing details on the nematode response against the Nkp-1 toxic activity (section 5.3). I also will demonstrate how the natural gut microbiota of a nematode

283 impacts the host survival upon exposure to toxic compound such as Nkp-1 (section 5.4). An antinematode drug discovery pipeline from the marine environment is also proposed at the end of the chapter (Figure 5.1).

5.2 Blue gold from the ocean; a new antinematode compound from the surface associated marine bacterium Pseudoalteromonas tunicata D2

Terrestrial plants or soil microbial-derived bioactive compounds have been recognised as the main repository for antinematode drug discovery (Sekurova et al., 2019). However, due to the high re-discovery rate of terrestrial compounds, the current pharmaceutical interest shifts to the exploration and exploitation of bioactive compounds or ‘blue gold’ from the marine environment particularly from surface associated bacteria (Chapter 1) (Figure 1.1) (Sekurova et al., 2019).

A novel nematode-killing protein-1 (Nkp-1) produced by the Escherichia coli HG8 recombinant strain carrying the 13.8 kb P. tunicata DNA insert has been successfully identified and intensively characterised in this thesis (Chapter 2). The protein is encoded by a single gene hp1 that is responsible for the antinematode activity of the HG8 and HP1 clones against the model nematode Caenorhabditis elegans (Chapter 2). Whilst Nkp-1 does not show close genetic identity to any characterised protein, its mode of action (MOA) against C. elegans resembles the killing activity shown for pore-forming toxins (PFTs), for example the crystal Cry5B toxin produced by Bacillus thuringiensis and aerolysin toxin produced by Aeromonas sp. Those PFTs function by disrupting the animal’s intestinal cell membrane hence resulting in necrosis and severe gastrointestinal damage (Griffitts et al., 2005; Szczesny et al., 2011) (Chapter 3). Interestingly, a putative carbohydrate binding module (CBM6 or CBM35) was also found located in the N- terminal of the Nkp-1 that is suggested to bind to a yet unidentified glycoconjugate receptor of C. elegans cells (Chapter 2). Similar carbohydrate binding modules have also discovered in the Cry toxins of B. thuringiensis (Ficko-Blean and Boraston, 2012; Feng et al., 2015; Rodríguez-González et al., 2020). These fundamental findings in Chapter 2 and Chapter 3 may pave the way for Nkp-1 to become a future antinematode drug

284 candidate. Indeed, a similar pathway from discovery to development was taken for the drug avermectin first identified from a biosynthetic genes cluster in Streptomyces avermectinius genomic sequence. Thereafter the surrogate nematode C. elegans was used to facilitate developmental research including elucidation of the MOA and potential for avermectin to be used as a novel antinematode therapeutic agent (Burg et al., 1979; Ikeda et al., 1999; Deng et al., 2017).

In drug developmental procedures, microbial-derived bioactive compounds with medicinal value are regarded as renewable source for intensive research and large-scale drug production by the pharmaceutical industry (Bremer, 1998; Penesyan et al., 2010). However, the successful approval of an antinematode drug candidate is based on its known molecular function, its MOA against the target parasitic organism and its potential toxicity against other non-target cells or system (Garcia-Bustos et al., 2019). While the discoveries on nematicidal Nkp-1 MOAs and the responsible expressing gene may kickstart the development of Nkp-1 as a novel antinematode drug in the future, the efficacy of Nkp-1 against other common animal or plant parasitic nematodes (for example Ascaris lumbricoides, Meloidogyne incognita and H. contortus) (Jones et al., 2013; Mushonga et al., 2018; Bundy et al., 2020) in-vitro and in-vivo needs to be evaluated prior to the next large-scale trial phase against human, animals or plants (see https://www.fda.gov/patients/learn-about-drug-and-device-approvals/drug- development-process). (Figure 5.1).

Genetic characterisation and protein structural information of Nkp-1 (Chapter 2 and Chapter 3) may contribute to the development of antinematode drugs with improved pharmacological properties and higher bioactivity. The improvement of Nkp-1 nematicidal activity at the molecular level could be performed using protein engineering, mutation and gene manipulation techniques (Walters et al., 2008; Kamionka, 2011) (Figure 5.1). Similar approaches have been applied for the improvement of insecticidal activity of Cry toxins produced by Bacillus thuringiensis (Walters et al., 2008; Hou et al., 2019). For example, insertion of recognition sequence of cathepsin (a family) in the loop between α-helices 3 and 4 in Domain I of Cry3A successfully improved the toxin solubilisation by proteases found in the gut of western corn rootworm 285

(Diabrotica virgifera), thus resulting in increased Cry toxicity and parasite mortality (Walters et al., 2008). Another outstanding breakthrough was also found by Hou and colleagues (Hou et al., 2019) who discovered that fusing Escherichia coli maltose binding protein to the Cry8Hb significantly improved its insecticidal activity, presumably due to the increased toxin solubility in the parasite gut.

Moreover, synthetic biology involving DNA recombination and heterologous gene(s) expression are commonly applied in large scale microbial therapeutic drugs production (Yip et al., 2019). For this purpose, an industrial compliant host such as Escherichia coli is used as the recombinant expressing strain due to its rapid growth, short doubling time and simple scale up processes (Selas Castiñeiras et al., 2018; Yip et al., 2019). The hp1 cloning and heterologous Nkp-1 expression by E. coli producer in Chapter 2 may represent the initial step for future industrial level production of Nkp-1 (Figure 5.1). The next step for the upstream phase of Nkp-1 large scale production involves optimisation of fermentation conditions including both physical (i.e. fermentation process (batch, fed- batch, continuous), pH, temperature, aeration, agitation speed, equipment used, strategies to avoid contamination) and chemical parameters (microbial nutrient requirement i.e. carbon, nitrogen and phosphate sources) (Figure 5.1) (Singh et al., 2017; dos Santos et al., 2018). After a high concentration of Nkp-1 production is achieved, downstream processes that involve initial recovery (cell separation from the culture media and protein extraction), purification to separate overall contaminants and polishing phase to exclude specific contaminants need to be developed and optimised (Figure 5.1) (Shukla and Thömmes, 2010; dos Santos et al., 2018). Chromatography techniques based on protein affinity (Łącki and Riske, 2020), size exclusion (Yan et al., 2019) or expanded bed adsorption (EBA) (Mofidian et al., 2020) represent potential downstream strategies that are commonly used to purify the desired biopharmaceutical products. Heterologous production of several biopharmaceutical products for example human growth hormone (hGH) (Persson et al., 2005) recombinant interferon α‐1 (rhIFNα‐1) and α2b (IFN‐α2b) (Guan et al., 1996; Lin et al., 2013) and insulin (Hart et al., 1994) using E. coli corroborates the potential of Nkp-1 industrial scale production using a similar recombinant bacterial producer.

286

In addition, the fact that the antinematode activity of Nkp-1 can be produced by a single gene hp1 may pave the way for the development of nematode-resistance transgenic plants. Techniques such as the gene gun method (also known as “Micro-Projectile Bombardment” or “Biolistic” method) and Agrobacterium tumefaciens mediated transformation method can be used to transform the hp1 gene directly into the plant tissue or cells resulting in artificially inserted hp1 into the plant genome (Rani and Usha, 2013; Abbas, 2018). Successful foreign DNA transformation is confirmed via antibiotic selection and the plants tissue with hp1 can be regenerated to become a whole plant using the tissue culture techniques (Rani and Usha, 2013; Abbas, 2018). Such transgenic plants with insect-resistance phenotype have been successfully developed using similar methods. These include Bt cotton, Bt corn, Bt potato, Bt eggplant and Bt soy that were successfully transformed with gene expressing Cry toxin produced by B. thuringiensis (Perlak et al., 1990; Koziel et al., 1993; Federici, 2005; Koch et al., 2015). A transgenic tomato plant that is resistant to nematode has also been successfully generated using the similar approach (Li et al., 2007). While this idea may represent as an environmentally friendly biocontrol management against the drug resistant plant parasitic nematode as well as reducing the incorporation of hazardous chemical nematicides such as carbamates, organophosphates and fumigants in agriculture sector (Holden-Dye and Walker, 2014), a cytotoxicity test is required to confirm Nkp-1 as non-hazardous to the potential consumers (Geary et al., 2015).

5.3 Nematode behaviour and innate immunity regulation against Nkp-1

C. elegans have developed different defence strategies against toxic microorganisms. These include its tough cuticle as a physical barrier from microbial invasion, innate or induced associative learning avoidance behaviour against toxic repellents and enhancement of several highly conserved innate immunity pathways (Engelmann and Pujol, 2010). Those physical, behavioural and immunity responses are also demonstrated by parasitic nematodes as a survival strategy against the exposure of noxious compound or immunity response by the parasitized host (Ketschek et al., 2004; Cooper and Eleftherianos, 2016; Lok, 2016; Ruark-Seward et al., 2019). Investigation on nematode

287 response against the potential antinematode compound should be included as part of antinematode drug characterisation and development pipeline (Figure 5.1).

In this thesis, I highlighted the defence strategy employed by nematodes to survive against the Nkp-1 toxic activity (Chapter 3). It appears that C. elegans employs both an innate and an associative learning avoidance behaviour to escape from toxic Nkp-1 expressing strains; similar strategies that are employed by the nematode upon exposure to B. thuringiensis, Coxiella burnetii and Vibrio parahaemolyticus (Hasshoff et al., 2007; Sellegounder et al., 2011; Luo et al., 2013; Battisti et al., 2017). The dog hookworm Anclyostoma caninum also exhibited an innate avoidance behavioural response against sodium dodecyl sulphate (SDS) as has been shown by C. elegans (Ketschek et al., 2004). The avoidance behaviour is controlled by chemosensory neurons with ciliated projections that penetrate the cuticle and are exposed to the environment to detect different molecular cues (Bargmann, 2006; Meisel and Kim, 2014). C. elegans avoidance behaviour may indicate the natural response of parasitic nematode against the Nkp-1. Interestingly, the observed innate and associative learning avoidance behaviours of C. elegans against Nkp- 1 in this study (Chapter 3) might indicate the Nkp-1 potentials as a dual function-bio- nematicide either as a parasitic nematodes’ repellent or as a nematicidal agent which can kill the parasites upon Nkp-1 ingestion. The dual mechanisms of parasitic nematode biocontrol were also shown by Pseudomonas chlororaphis due to the extracellular production of pyrrolnitrin and hydrogen cyanide (Nandi et al., 2015).

Furthermore, C. elegans’ response against the Nkp-1 toxicity is also investigated in this thesis. I have found that the ILR signalling involving the daf-2/daf-16 and sek-1 p38- MAPK pathways were activated upon exposure to the Nkp-1 toxicity (Chapter 3). Given that immune systems in members of phylum nematoda are highly conserved (Schulenburg et al., 2004), this observation may indicate the parasitic nematode’s response to the antinematode drug treatment by Nkp-1 in the future. Studies have shown that the daf-2/daf-16 ILR signalling and p38-MAPK pathways are activated upon C. elegans exposure to several toxic bacteria i.e. Serratia marcescens, Microbacterium nematophilum and Pseudomonas aeruginosa (Mallo et al., 2002; O’Rourke et al., 2006; Troemel et al., 2006). The daf-2/daf-16 ILR signalling activates the downstream genes 288 encoding microbial effectors such as lysozyme and superoxide dismutase that are pivotal for nematode defence and longevity (Honda and Honda, 1999; Murphy et al., 2003; Bai et al., 2014). Interestingly, the daf-2/daf-16 pathway also activates the cyp-35B1 encoding cytochrome P450 enzyme that is important for xenobiotic or toxic compound detoxification including the antinematode drug by the host nematode (Iser et al., 2011). Genome analysis of parasitic nematodes i.e. Globodera pallida and Haemonchus contortus discovered more than 20 genes expressing Cytochrome P450 enzyme (Laing et al., 2013; Cotton et al., 2014) that may enhance the nematode resistance against antinematode drug through xenobiotic detoxification (Matoušková et al., 2016). The ability of nematodes to detoxify hazardous compound may affect the nematodes sensitivity or tolerance to drug treatment (Ménez et al., 2016).

5.4 Native nematode gut microbiota; a neglected factor in antinematode drug development

Gut microbiota association in organisms are ubiquitous in nature (Bates et al., 2006; López Nadal et al., 2020). A balanced host-gut microbiota interactions are believed to support host health and fitness, growth, tissue development, metabolism and behaviour (Sampson and Mazmanian, 2015; Tilg and Moschen, 2015; Thursby and Juge, 2017; Cani, 2018; Singhvi et al., 2020). Gut microbiota establishment also confers host protection against environmental stressors including toxic pollutants (Claus et al., 2016), excessive temperature (Raza et al., 2020) and colonisation by detrimental microorganisms (Rolhion and Chassaing, 2016). In fact, a number of exemplary studies support the importance of gut microbiota in improving C. elegans survival against toxic microorganisms including pathogens Pseudomonas aeruginosa (Montalvo-Katz et al., 2013), Enterococcus faecalis and Staphylococcus aureus (Kim and Mylonakis, 2012). However, until now, less is known about the impact of gut microbiota on C. elegans physiology and physical appearance upon exposure to toxic compound or detrimental bacterial (Ezcurra, 2018).

289

In this thesis, I demonstrated the impact of gut microbiota establishment on C. elegans survival and reduction of physical damages upon challenge with Nkp-1 expressing strains. Unlike the undefined gut microbiota (UGM), establishment of a defined gut microbiota (DGM) improved C. elegans N2 survival, reduced Nkp-1 expressing bacterial colonisation and diminished physical damages compared to the N2 monoxenic animals (Chapter 4). Using 16S rRNA gene amplicon sequencing I found a higher abundance of bacteria from the genera Xanthomonas and Sphingobacterium were observed in UGM while Achromobacter, Ochrobactrum and Stenotrophomonas were abundantly found in the DGM (Chapter 4). While those genera represent as part of C. elegans natural gut microbiota (Berg et al., 2016; Dirksen et al., 2016; Samuel et al., 2016), the difference of each bacterial abundance significantly affect the nematode survival and capability to control the toxic bacterial colonisation (HP1 and HG8) and the resulting physical damages (Samuel et al., 2016).

Given the difference of survival and severity of physical injuries observed between both N2_DGM and N2 monoxenic populations, there is little doubt that gut microbiota is pivotal also for other animal models that are frequently used in research such as C. elegans (Ezcurra, 2018), fruit fly Drosophilla melanogaster, zebra fish Danio rerio and mice Mus musculus (Round and Mazmanian, 2009). It has been shown that the germ-free surrogate organisms for example mouse were more susceptible to pathogenic microbial infection while showing distinct physiological and functional variances particularly in the gastrointestinal system compared to the normal counterparts (Al-Asmakh and Zadjali, 2015). For example, the germ-free mouse displays larger cecum due to mucus accumulation and undigested fibres which may indicate inefficient digestion system (Al- Asmakh and Zadjali, 2015). Furthermore, while being more susceptible to pathogen infection the germ-free zebrafish model also displays a lack of intestinal microvilli alkaline phosphatase activity, immature glycan expression on cell surface and reduction of goblet and enteroendocrine cells (Bates et al., 2006). Animal models for example C. elegans serve as powerful tools in scientific research including antinematode drug discovery (Leung et al., 2008; Burns et al., 2015; Wang et al., 2018; Queirós et al., 2019) (Chapter 1). However, the differences in survival between germ-free and gut microbiota- associated animal models should be considered and further tests using animal models that are established with their native gut microbiota (for example C. elegans in this study) 290 performed to obtain an understanding of bioactivity in a near natural setting (Figure 5.1) (Turner, 2018).

291

CONSUMERS

Figure 5.1 Schematic diagram representing the antinematode drug discovery pipeline from the marine environment. Antinematode compound can be isolated either from marine floras, faunas, planktonic and surface associating microorganism or from the culture-independent metagenomic libraries construction. Molecular cloning, compound characterisation and improvement are performed prior to drug approval and industrial scale drug production involving the upstream and downstream processes. Figure was designed using Biorender at https://biorender.com/

292

5.5 Conclusion, research limitation and future perspective

This research project addressed scientific concepts that span from the discovery, characterisation and determination of mode of action (MOA) of a novel antinematode compound from the marine bacterium P. tunicata D2 to its effect against nematodes established with gut microbiota. Transposon mutagenesis, genetic complementation and individual gene cloning led to the discovery of hp1 as the responsible gene in expressing the nematicidal protein Nkp-1 that is highly toxic to C. elegans larvae and hermaphrodites but ineffective against the nematode eggs. Importantly, Nkp-1 harbours a putative carbohydrate binding module (CBM) located at the N-terminal that is predicted to bind to C. elegans glycoconjugate receptor prior to initiation of pore-forming activity against the nematode resulting in necrotic cell death and severe physical damages. This thesis also shows the nematode’s avoidance behaviour and differentially regulated immune response involving the daf-2/daf-16 ILR signalling and sek-1 p38-MAPK pathway as a defence mechanism against the Nkp-1 toxicity. While many studies used C. elegans N2 monoxenic as the animal model in deciphering antinematode activity by the target compound or organisms, few have interrogated the effect of natural gut microbiota association in the model nematode upon exposure to the toxic compound or microorganisms of interest. My work has addressed this knowledge gap to show that gut microbiota establishment may improve the nematode survival, reduce toxic bacterial colonisation and diminish physical damages.

However, this research also has several limitations. While the Nkp-1 has been intensively characterised, the specific nematode glycoconjugate receptor targeted for Nkp-1 binding is still unknown (Chapter 2). Furthermore, while two models of Nkp- 1 containing carbohydrate binding modules (CBM6 and CBM35) have been proposed, the accurate molecular structure of Nkp-1 is still undetermined (Chapter 2). From the perspective of C. elegans response towards Nkp-1 toxicity, the overall immune defence strategies employed by the nematode cannot be fully unravelled given that the qPCR analysis only focused on certain immunity pathways and several downstream genes (Chapter 3). In addition, while the 16S rRNA gene sequencing and microbial community analysis successfully discovered the difference of microbial structure in

293 the SOIL and nematode samples in form of relative abundance, the data itself cannot make conclusions about absolute abundances of these microbiomes (Chapter 4). Therefore, future work on the identification of glycoconjugate receptor targeted for binding by Nkp-1 is crucial given that mutation on the binding receptor may result in nematode resistance against the Nkp-1 killing activity. This research could be performed through forward genetic mutation on C. elegans to identify gene(s) that are responsible for the expression of glycoconjugate receptors targeted by Nkp-1. Furthermore, an X-ray crystallography or NMR spectroscopy should be conducted to identify the precise three-dimensional structure of most of the atoms in Nkp-1, thus revealing the best visualisation of the Nkp-1 structure and improve the understanding on its physical-chemical properties. A complete understanding on C. elegans response towards the Nkp-1 toxicity should be achieved via comprehensive phenotyping and transcriptomic profiling on the nematodes upon exposure to the Nkp-1. The transcriptomic investigation can provide a global snapshot of molecular alteration in C. elegans including its immune defence strategies against the Nkp-1 toxicity. In addition, a quantitative PCR experiment also could be performed to provide absolute abundances of different bacterial species. Finally, the toxicity of Nkp-1 against human (e.g. Ancylostoma duodenale, Brugia malayi), animal (Anisakis simplex, Haemonchus contortus) or plant parasitic nematodes (e.g. Meloidogyne incognita, Radopholus Similis) in-vitro and in-vivo should be performed to understand the broad scale potential of this new antinematode agent.

294

5.6 REFERENCES

Abbas, M. S. T. (2018). Genetically engineered (modified) crops (Bacillus thuringiensis crops) and the world controversy on their safety. Egyptian Journal of Biological Pest Control, 28(1), 52. Al-Asmakh, M., & Zadjali, F. (2015). Use of germ-free animal models in microbiota- related research. Journal of Microbiology and Biotechnology, 25(10), 1583- 1588. Bai, Y., Zhi, D., Li, C., Liu, D., Zhang, J., Tian, J., Wang, X., Ren, H., & Li, H. (2014). Infection and immune response in the nematode Caenorhabditis elegans elicited by the phytopathogen Xanthomonas. Journal of Microbiology and Biotechnology, 24(9), 1269-1279. Bargmann, C. I. (2006). Chemosensation in C. elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, Wormbook. doi: 10.1895/wormbook.1.123.1. Bates, J. M., Mittge, E., Kuhlman, J., Baden, K. N., Cheesman, S. E., & Guillemin, K. (2006). Distinct signals from the microbiota promote different aspects of zebrafish gut differentiation. Developmental Biology, 297(2), 374-386. Battisti, J. M., Watson, L. A., Naung, M. T., Drobish, A. M., Voronina, E., & Minnick, M. F. (2017). Analysis of the Caenorhabditis elegans innate immune response to Coxiella burnetii. Innate immunity, 23(2), 111-127. Berg, M., Zhou, X. Y., & Shapira, M. (2016). Host-specific functional significance of Caenorhabditis gut commensals. Frontiers in Microbiology, 7, 1622. Bremer, G. (1998). Marine microorganisms as a source of novel bioactive compounds; current status, prospects and strategies for the future. Marine Microorganisms for Industry, 112-120. Bundy, D. A., de Silva, N., Appleby, L. J., & Brooker, S. J. (2020). Intestinal Nematodes: Ascariasis. In Hunter's Tropical Medicine and Emerging Infectious Diseases (pp. 840-844): Elsevier. Burg, R. W., Miller, B. M., Baker, E. E., Birnbaum, J., Currie, S. A., Hartman, R., Kong, Y. L., Monaghan, R. L., Olson, G., Putter, I., Tunac, J. B., Wallick, H., Stapley, E. O., Oiwa, R., & Omura, S. (1979). Avermectins, new family of potent anthelmintic agents: producing organism and fermentation. Antimicrobial Agents and Chemotherapy, 15(3), 361-367. Burns, A. R., Luciani, G. M., Musso, G., Bagg, R., Yeo, M., Zhang, Y., Rajendran, L., Glavin, J., Hunter, R., Redman, E., Stasiuk, S., Schertzberg, M., Angus McQuibban, G., Caffrey, C. R., Cutler, S. R., Tyers, M., Giaever, G., Nislow, C., Fraser, A. G., MacRae, C. A., Gilleard, J., & Roy, P. J. (2015). Caenorhabditis elegans is a useful model for anthelmintic discovery. Nature Communications, 6, 7485-7485. Caminade, C., McIntyre, K. M., & Jones, A. E. (2019). Impact of recent and future climate change on vector-borne diseases. Annals of the New York Academy of Sciences, 1436(1), 157-173. Cani, P. D. (2018). Human gut microbiome: hopes, threats and promises. Gut, 67(9), 1716-1725. Claus, S. P., Guillou, H., & Ellero-Simatos, S. (2016). The gut microbiota: a major player in the toxicity of environmental pollutants? Biofilms and Microbiomes, 2(1), 1-11.

295

Codjoe, S. N. A., & Larbi, R. T. (2016). Climate change/variability and schistosomiasis transmission in Ga district, Ghana. Climate and Development, 8(1), 58-71. Cooper, D., & Eleftherianos, I. (2016). Parasitic nematode immunomodulatory strategies: recent advances and perspectives. Pathogens, 5(3), 58. Cotton, J. A., Lilley, C. J., Jones, L. M., Kikuchi, T., Reid, A. J., Thorpe, P., Tsai, I. J., Beasley, H., Blok, V., & Cock, P. J. (2014). The genome and life-stage specific transcriptomes of Globodera pallidaelucidate key aspects of plant parasitism by a cyst nematode. Genome Biology, 15(3), R43. Deng, Q., Zhou, L., Luo, M., Deng, Z., & Zhao, C. (2017). Heterologous expression of Avermectins biosynthetic gene cluster by construction of a Bacterial Artificial Chromosome library of the producers. Synthetic and Systems Biotechnology, 2(1), 59-64. Dirksen, P., Marsh, S. A., Braker, I., Heitland, N., Wagner, S., Nakad, R., Mader, S., Petersen, C., Kowallik, V., Rosenstiel, P., Félix, M.-A., & Schulenburg, H. (2016). The native microbiome of the nematode Caenorhabditis elegans: gateway to a new host-microbiome model. BMC Biology, 14(1), 38. dos Santos, N. V., de Carvalho Santos-Ebinuma, V., Pessoa Junior, A., & Pereira, J. F. B. (2018). Liquid–liquid extraction of biopharmaceuticals from fermented broth: trends and future prospects. Journal of Chemical Technology & Biotechnology, 93(7), 1845-1863. Engelmann, I., & Pujol, N. (2010). Innate immunity in C. elegans. In Invertebrate Immunity (pp. 105-121): Springer. Ezcurra, M. (2018). Dissecting cause and effect in host-microbiome interactions using the combined worm-bug model system. Biogerontology, 19(6), 567-578. Federici, B. A. (2005). Insecticidal bacteria: An overwhelming success for invertebrate pathology. Journal of Invertebrate Pathology, 89(1), 30-38. Feng, D., Chen, Z., Wang, Z., Zhang, C., He, K., & Guo, S. (2015). Domain III of Bacillus thuringiensis Cry1Ie toxin plays an important role in binding to peritrophic membrane of Asian Corn Borer. PloS One, 10(8), e0136430. Ficko-Blean, E., & Boraston, A. B. (2012). Insights into the recognition of the human glycome by microbial carbohydrate-binding modules. Current Opinion in Structural Biology, 22(5), 570-577. Fox, N. J., Marion, G., Davidson, R. S., White, P. C., & Hutchings, M. R. (2015). Climate-driven tipping-points could lead to sudden, high-intensity parasite outbreaks. Royal Society Open Science, 2(5), 140296. Garcia-Bustos, J. F., Sleebs, B. E., & Gasser, R. B. (2019). An appraisal of natural products active against parasitic nematodes of animals. Parasites & Vectors, 12(1), 306. Geary, T. G., Sakanari, J. A., & Caffrey, C. R. (2015). Anthelmintic drug discovery: into the future. The Journal of Parasitology, 101(2), 125-133. Genchi, C., Rinaldi, L., Mortarino, M., Genchi, M., & Cringoli, G. (2009). Climate and Dirofilaria infection in Europe. Veterinary Parasitology, 163(4), 286-292. Griffitts, J. S., Haslam, S. M., Yang, T., Garczynski, S. F., Mulloy, B., Morris, H., Cremer, P. S., Dell, A., Adang, M. J., & Aroian, R. V. (2005). Glycolipids as receptors for Bacillus thuringiensis crystal toxin. Science, 307(5711), 922-925. Guan, Y., Lilley, T. H., Treffry, T. E., Zhou, C.-L., & Wilkinson, P. B. (1996). Use of aqueous two-phase systems in the purification of human interferon-α1 from recombinant Escherichia coli. Enzyme and Microbial Technology, 19(6), 446- 455.

296

Hart, R. A., Lester, P. M., Reifsnyder, D. H., Ogez, J. R., & Builder, S. E. (1994). Large scale, in situ isolation of periplasmic IGF–I from E. coli. Nature Biotechnology, 12(11), 1113-1117. Hasshoff, M., Böhnisch, C., Tonn, D., Hasert, B., & Schulenburg, H. (2007). The role of Caenorhabditis elegans insulin-like signaling in the behavioral avoidance of pathogenic Bacillus thuringiensis. The FASEB Journal, 21(8), 1801-1812. Holden-Dye, L., & Walker, R. (2014). Anthelmintic drugs and nematocides: studies in Caenorhabditis elegans. In C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology, WormBook. Available at http://www.wormbook.org/chapters/www_anthelminticdrugs.2/anthelminticd rugs.2.html Honda, Y., & Honda, S. (1999). The daf-2 gene network for longevity regulates oxidative stress resistance and Mn-superoxide dismutase gene expression in Caenorhabditis elegans. The FASEB Journal, 13(11), 1385-1393. Hou, J., Cong, R., Izumi-Willcoxon, M., Ali, H., Zheng, Y., Bermudez, E., McDonald, M., Nelson, M., & Yamamoto, T. (2019). Engineering of Bacillus thuringiensis Cry proteins to enhance the activity against western Corn Rootworm. Toxins, 11(3), 162. Ikeda, H., Nonomiya, T., Usami, M., Ohta, T., & Ōmura, S. (1999). Organization of the biosynthetic gene cluster for the polyketide anthelmintic macrolide avermectin in Streptomyces avermitilis. Proceedings of the National Academy of Sciences of the USA, 96(17), 9509. Iser, W. B., Wilson, M. A., Wood, W. H., III, Becker, K., & Wolkow, C. A. (2011). Co-regulation of the DAF-16 target gene, cyp-35B1/dod-13, by HSF-1 in C. elegans dauer larvae and daf-2 Insulin pathway mutants. PloS One, 6(3), e17369. Jones, J. T., Haegeman, A., Danchin, E. G., Gaur, H. S., Helder, J., Jones, M. G., Kikuchi, T., Manzanilla‐López, R., Palomares‐Rius, J. E., & Wesemael, W. M. (2013). Top 10 plant‐parasitic nematodes in molecular plant pathology. Molecular Plant Pathology, 14(9), 946-961. Kamionka, M. (2011). Engineering of therapeutic proteins production in Escherichia coli. Current Pharmaceutical Biotechnology, 12(2), 268-274. Ketschek, A. R., Joseph, R., Boston, R., Ashton, F. T., & Schad, G. A. (2004). Amphidial neurons ADL and ASH initiate sodium dodecyl sulphate avoidance responses in the infective larva of the dog hookworm Anclyostoma caninum. International Journal for Parasitology, 34(12), 1333-1336. Kim, Y., & Mylonakis, E. (2012). Caenorhabditis elegans immune conditioning with the probiotic bacterium Lactobacillus acidophilus strain NCFM enhances Gram-positive immune responses. Infection and Immunity, 80(7), 2500-2508. Koch, M. S., Ward, J. M., Levine, S. L., Baum, J. A., Vicini, J. L., & Hammond, B. G. (2015). The food and environmental safety of Bt crops. Frontiers in Plant Science, 6, 283. Koziel, M. G., Beland, G. L., Bowman, C., Carozzi, N. B., Crenshaw, R., Crossland, L., Dawson, J., Desai, N., Hill, M., Kadwell, S., Launis, K., Lewis, K., Maddox, D., McPherson, K., Meghji, M. R., Merlin, E., Rhodes, R., Warren, G. W., Wright, M., & Evola, S. V. (1993). Field performance of elite transgenic maize plants expressing an insecticidal protein derived from Bacillus thuringiensis. Nature Biotechnology, 11(2), 194-200. Lacetera, N. (2018). Impact of climate change on animal health and welfare. Animal Frontiers, 9(1), 26-31. 297

Łącki, K. M., & Riske, F. J. (2020). Affinity chromatography: An enabling technology for large-scale bioprocessing. Biotechnology Journal, 15(1), 1800397. Laing, R., Kikuchi, T., Martinelli, A., Tsai, I. J., Beech, R. N., Redman, E., Holroyd, N., Bartley, D. J., Beasley, H., & Britton, C. (2013). The genome and transcriptome of Haemonchus contortus, a key model parasite for drug and vaccine discovery. Genome Biology, 14(8), 1-16. Leung, M. C. K., Williams, P. L., Benedetto, A., Au, C., Helmcke, K. J., Aschner, M., & Meyer, J. N. (2008). Caenorhabditis elegans: An emerging model in biomedical and environmental toxicology. Toxicological Sciences, 106(1), 5- 28. Li, X. Q., Wei, J. Z., Tan, A., & Aroian, R. V. (2007). Resistance to root-knot nematode in tomato roots expressing a nematicidal Bacillus thuringiensis crystal protein. Plant Biotechnology Journal, 5(4), 455-464. Lin, Y. K., Ooi, C. W., Tan, J. S., Show, P. L., Ariff, A., & Ling, T. C. (2013). Recovery of human interferon alpha-2b from recombinant Escherichia coli using alcohol/salt-based aqueous two-phase systems. Separation and Purification Technology, 120, 362-366. Lok, J. B. (2016). Signaling in Parasitic Nematodes: Physicochemical communication between host and parasite and endogenous molecular transduction pathways governing worm development and survival. Current Clinical Microbiology Reports, 3(4), 186-197. López Nadal, A., Ikeda-Ohtsubo, W., Sipkema, D., Peggs, D., McGurk, C., Forlenza, M., Wiegertjes, G. F., & Brugman, S. (2020). Feed, microbiota, and gut immunity: using the zebrafish model to understand fish health. Frontiers in Immunology, 11, 114. Luo, H., Xiong, J., Zhou, Q., Xia, L., & Yu, Z. (2013). The effects of Bacillus thuringiensis Cry6A on the survival, growth, reproduction, locomotion, and behavioral response of Caenorhabditis elegans. Applied Microbiology and Biotechnology, 97(23), 10135-10142. Mallo, G. V., Kurz, C. L., Couillault, C., Pujol, N., Granjeaud, S., Kohara, Y., & Ewbank, J. J. (2002). Inducible antibacterial defense system in C. elegans. Current Biology, 12(14), 1209-1214. Matoušková, P., Vokřál, I., Lamka, J., & Skálová, L. (2016). The role of xenobiotic- metabolizing enzymes in anthelmintic deactivation and resistance in helminths. Trends in Parasitology, 32(6), 481-491. McIntyre, K. M., Setzkorn, C., Hepworth, P. J., Morand, S., Morse, A. P., & Baylis, M. (2017). Systematic assessment of the climate sensitivity of important human and domestic animals pathogens in Europe. Scientific Reports, 7(1), 1- 10. Meisel, J. D., & Kim, D. H. (2014). Behavioral avoidance of pathogenic bacteria by Caenorhabditis elegans. Trends in Immunology, 35(10), 465-470. Ménez, C., Alberich, M., Kansoh, D., Blanchard, A., & Lespine, A. (2016). Acquired tolerance to ivermectin and moxidectin after drug selection pressure in the nematode Caenorhabditis elegans. Antimicrobial Agents and Chemotherapy, 60(8), 4809-4819. Mofidian, R., Barati, A., Jahanshahi, M., & Shahavi, M. H. (2020). Fabrication of novel agarose-nickel bilayer composite for purification of protein nanoparticles in expanded bed adsorption column. Chemical Engineering Research and Design, 159, 291-299.

298

Montalvo-Katz, S., Huang, H., Appel, M. D., Berg, M., & Shapira, M. (2013). Association with soil bacteria enhances p38-dependent infection resistance in Caenorhabditis elegans. Infection and Immunity, 81(2), 514-520. Morgan, E. R., Aziz, N.-A. A., Blanchard, A., Charlier, J., Charvet, C., Claerebout, E., Geldhof, P., Greer, A. W., Hertzberg, H., & Hodgkinson, J. (2019). 100 questions in livestock helminthology research. Trends in Parasitology, 35(1), 52-71. Murphy, C. T., McCarroll, S. A., Bargmann, C. I., Fraser, A., Kamath, R. S., Ahringer, J., Li, H., & Kenyon, C. (2003). Genes that act downstream of DAF-16 to influence the lifespan of Caenorhabditis elegans. Nature, 424(6946), 277-283. Mushonga, B., Habumugisha, D., Kandiwa, E., Madzingira, O., Samkange, A., Segwagwe, B. E., & Jaja, I. F. (2018). Prevalence of Haemonchus contortus infections in sheep and goats in Nyagatare District, Rwanda. Journal of Veterinary Medicine, 2018, 3602081. O’Rourke, D., Baban, D., Demidova, M., Mott, R., & Hodgkin, J. (2006). Genomic clusters, putative pathogen recognition molecules, and antimicrobial genes are induced by infection of C. elegans with M. nematophilum. Genome Research, 16(8), 1005-1016. Penesyan, A., Kjelleberg, S., & Egan, S. (2010). Development of novel drugs from marine surface associated microorganisms. Marine Drugs, 8(3), 438-459. Perlak, F. J., Deaton, R. W., Armstrong, T. A., Fuchs, R. L., Sims, S. R., Greenplate, J. T., & Fischhoff, D. A. (1990). Insect resistant cotton plants. Nature Biotechnology, 8(10), 939-943. Persson, J., Andersen, D. C., & Lester, P. M. (2005). Evaluation of different primary recovery methods for E. coli‐derived recombinant human growth hormone and compatibility with further down‐stream purification. Biotechnology and Bioengineering, 90(4), 442-451. Queirós, L., Pereira, J., Gonçalves, F., Pacheco, M., Aschner, M., & Pereira, P. (2019). Caenorhabditis elegans as a tool for environmental risk assessment: emerging and promising applications for a “nobelized worm”. Critical Reviews in Toxicology, 49(5), 411-429. Rani, J. S., & Usha, R. (2013). Transgenic plants: Types, benefits, public concerns and future. Journal of Pharmacy Research, 6(8), 879-883. Raza, M. F., Wang, Y., Cai, Z., Bai, S., Yao, Z., Awan, U. A., Zhang, Z., Zheng, W., & Zhang, H. (2020). Gut microbiota promotes host resistance to low- temperature stress by stimulating its arginine and proline metabolism pathway in adult Bactrocera dorsalis. PLoS Pathogens, 16(4), e1008441. Rodríguez-González, Á., Porteous-Álvarez, A. J., Val, M. D., Casquero, P. A., & Escriche, B. (2020). Toxicity of five Cry proteins against the insect pest Acanthoscelides obtectus (Coleoptera: Chrisomelidae: Bruchinae). Journal of Invertebrate Pathology, 169, 107295. Rolhion, N., & Chassaing, B. (2016). When pathogenic bacteria meet the intestinal microbiota. Philosophical Transactions of the Royal Society B Biological Sciences, 371(1707), 20150504. Round, J. L., & Mazmanian, S. K. (2009). The gut microbiota shapes intestinal immune responses during health and disease. Nature Reviews Immunology, 9(5), 313-323. Ruark-Seward, C. L., Davis, E. L., & Sit, T. L. (2019). Electropermeabilization-based fluorescence in situ hybridization of whole-mount plant parasitic nematode specimens. MethodsX, 6, 2720-2728.

299

Sampson, T. R., & Mazmanian, S. K. (2015). Control of brain development, function, and behavior by the microbiome. Cell Host Microbe, 17(5), 565-576.

Samuel, B. S., Rowedder, H., Braendle, C., Félix, M.-A., & Ruvkun, G. (2016). Caenorhabditis elegans responses to bacteria from its natural habitats. Proceedings of the National Academy of Sciences of the USA, 113(27), E3941- E3949. Schulenburg, H., Léopold Kurz, C., & Ewbank, J. J. (2004). Evolution of the innate immune system: the worm perspective. Immunological Reviews, 198(1), 36- 58. Sekurova, O. N., Schneider, O., & Zotchev, S. B. (2019). Novel bioactive natural products from bacteria via bioprospecting, genome mining and metabolic engineering. Microbial Biotechnology, 12(5), 828-844. Selas Castiñeiras, T., Williams, S. G., Hitchcock, A. G., & Smith, D. C. (2018). E. coli strain engineering for the production of advanced biopharmaceutical products. FEMS Microbiology Letters, 365(15). Sellegounder, D., Pandian, S., & Balamurugan, K. (2011). Changes in Caenorhabditis elegans exposed to Vibrio parahaemolyticus. Journal of Microbiology and Biotechnology, 21, 1026-1035. Short, E. E., Caminade, C., & Thomas, B. N. (2017). Climate change contribution to the emergence or re-emergence of parasitic diseases. Infectious Diseases, 10, 1-7. Shukla, A. A., & Thömmes, J. (2010). Recent advances in large-scale production of monoclonal antibodies and related proteins. Trends in Biotechnology, 28(5), 253-261. Singh, V., Haque, S., Niwas, R., Srivastava, A., Pasupuleti, M., & Tripathi, C. K. M. (2017). Strategies for fermentation medium optimization: an in-depth review. Frontiers in Microbiology, 7, 2087-2087. Singhvi, N., Gupta, V., Gaur, M., Sharma, V., Puri, A., Singh, Y., Dubey, G. P., & Lal, R. (2020). Interplay of human gut microbiome in health and wellness. Indian Journal of Microbiology, 60(1), 26-36. Szczesny, P., Iacovache, I., Muszewska, A., Ginalski, K., van der Goot, F. G., & Grynberg, M. (2011). Extending the aerolysin family: from bacteria to vertebrates. PloS One, 6(6), e20349. Thursby, E., & Juge, N. (2017). Introduction to the human gut microbiota. The Biochemical Journal, 474(11), 1823-1836. Tilg, H., & Moschen, A. R. (2015). Food, immunity, and the microbiome. Gastroenterology, 148(6), 1107-1119. Troemel, E. R., Chu, S. W., Reinke, V., Lee, S. S., Ausubel, F. M., & Kim, D. H. (2006). p38 MAPK regulates expression of immune response genes and contributes to longevity in C. elegans. PLoS Genetics, 2(11), e183. Turner, P. V. (2018). The role of the gut microbiota on animal model reproducibility. Animal Models and Experimental Medicine, 1(2), 109-115. Verschave, S. H., Charlier, J., Rose, H., Claerebout, E., & Morgan, E. R. (2016). Cattle and nematodes under global change: transmission models as an ally. Trends in Parasitology, 32(9), 724-738. Walters, F. S., Stacy, C. M., Lee, M. K., Palekar, N., & Chen, J. S. (2008). An engineered chymotrypsin/cathepsin G site in domain I renders Bacillus thuringiensis Cry3A active against Western Corn Rootworm larvae. Applied and Environmental Microbiology, 74(2), 367-374.

300

Wang, H., Park, H., Liu, J., & Sternberg, P. W. (2018). An efficient genome editing strategy to generate putative null mutants in Caenorhabditis elegans using CRISPR/Cas9. G3: Genes|Genomes|Genetics, 8(11), 3607. Yan, Y., Xing, T., Wang, S., Daly, T. J., & Li, N. (2019). Coupling mixed-mode size exclusion chromatography with native mass spectrometry for sensitive detection and quantitation of homodimer impurities in bispecific IgG. Analytical Chemistry, 91(17), 11417-11424. Yip, C.-H., Yarkoni, O., Ajioka, J., Wan, K.-L., & Nathan, S. (2019). Recent advancements in high-level synthesis of the promising clinical drug, prodigiosin. Applied Microbiology and Biotechnology, 103(4), 1667-1680.

301

APPENDIX I

Table 1A Example of parasitic nematode infection in human and clinical symptoms

Parasitic nematode species Common Parasitic Estimated number of Pathophysiology or clinical symptoms Reference nematode's name worldwide infections Ancylostoma Hookworm 470 million Anaemia due to blood loss, hypoalbuminemia, (Ghodeif and Jain, duodenale/ Necator malnutrition 2019) americanus Ascaris lumbricoides Roundworm 819 million Abdominal pain, vomiting, liver abscess, (Bundy et al., 2020) cholangitis, pancreatitis, appendicitis, Intestinal obstruction, growth retardation Trichuris trichiura Whipworm 450 million to 1 billion Abdominal discomfort, painful defecation, mucus (Viswanath and discharge, anaemia, impaired growth and Williams, 2019) cognitive development Enterobius vermicularis Pinworm 200 million Perianal pruritus, restlessness, appendicitis, a (Kucik et al., 2004; secondary infection which resulted in urinary tract Fan et al., 2019) infection or pelvic inflammation Strongyloides stercoralis Threadworm 300 million Mainly affect immunocompromised patients (Bisoffi et al., resulting in a severe problem with the 2013; Miglioli- gastrointestinal and respiratory system. Galvão et al., 2020) philippinensis or none 500 million Weight loss, chronic diarrhoea, oedema, (Attia et al., 2012) philippinensis abdominal pain, cachexia and weakness. Dracunculus medinensis Guinea worm 3.5 million Severe pain especially at joints, skin abscess, (Pichakacheri, oedema, blister, ulcer, fever, nausea and vomiting 2019; Hopkins, 2020) Trichinella spiralis Trichina worm 11 million Fever, facial oedema, diarrhoea, inflammation in (Dupouy-Camet, muscle or myositis, fatal 2000; Zhao et al., 2017) Trichostrongylus spp. i.e.T. none 5.5 million Abdominal pain, diarrhoea, flatulence, nausea, (Ghadirian and colubriformis, T. orientalis, T. fatigue, anaemia, eosinophilia Arfaa, 1975; Sato et vitrinus, T. axei, T. capricola, al., 2011) T. probolurus and T. skrjabini 302

Table 1A (Continue) Example of parasitic nematode infection in human and clinical symptoms

Parasitic nematode species Common Parasitic Estimated number Pathophysiology or clinical symptoms Reference nematode's name of worldwide infections Brugia malayi, B. timori, Filarial worm 893 million Lymphatic filariasis leading to tissue swelling WHO (2020) (Reaves Wuchereria bancrofti (lymphoedema) or skin/tissue thickening et al., 2018) (elephantiasis), scrotal swelling Dioctophyme renale Giant Kidney Worm Unknown nephritis, loin pain, hematuria, urgent (Yang et al., 2019) urination, abdominal pain, anaemia, fever, renal colic, papule with itching Halicephalobus gingivalis none Unknown meningoencephalitis (inflammation in the (Papadi et al., 2013) (also reffered to as H. brain), myelitis (inflammation in the spinal deletrix or Micronema cord), fever, mental changes, lethargy, fatal deletrix) Angiostrongylus Rat lungworm Unknown Headache, neck stiffness, paraesthesia, (Wang et al., 2008; cantonensis vomiting and nausea. Symptoms like Niebuhr et al., 2020) coughing, rhinorrhea, sore throat, fever and malaise upon nematodes entrance into the lung, eosinophilic meningitis, and meningoencephalitis. Gnathostoma spinigerum, G. none Unknown Subcutaneous tissues infection and can invade (Rusnak and Lucey, doloresi, G. nipponicum, G. to deeper tissue leading to severe swelling, 1993; Nomura et al., hispidum, G. malaysiae, and localised pain, pruritus, erythema and damage 2000; Kim et al., 2013) G.binucleatum on organs Onchocerca volvulus none 187 million Subcutaneous tissue infection, hypopigmented (Richards Jr et al., skin lesions, severe itching, keratitis, 2020) inflammation and blindness (river blindness disease) Dirofilaria repens none none Nodules formation in the subcutaneous tissue (Genchi and Kramer, or ocular conjunctiva. Can cause tumour-like 2017) infection in the lung

303

Table 2A Example of plants parasitic nematodes, the associated host and clinical symptoms

Parasite Parasitic nematode species common Plant host Pathophysiology or clinical symptoms Reference name Meloidogyne arenaria, Root-knot Rice (Oryza sativa) Swelling or hook-shaped galls formation at the (Jones et al., 2013; M. incognita, M. javanica, M. nematode root tips. Disruption of water and nutrients Mantelin et al., 2017) graminicola and M. hapla absorption due to modification of the vascular system, resulting in retarded plant development and reproduction

Heterodera avenae, H. Cyst- Cereal; wheat (Triticum (Smiley, 2010; Jones et glycines and H. filipjevi nematode aestivum), barley al., 2013; Masonbrink (Hordeum vulgare) and et al., 2019; S. Yang et Root penetration and formation of "syncytium" oat (Avena sativa), al., 2019) (a specialised, fused root cells for feeding and tobacco (tobacco nutrient absorption) by the sedentary parasite. (Nicotiana The syncytium cells are metabolically active benthamiana), Soybean with enriched cytoplasm, shrinking vacuole and (Glycine max) enlarged nuclei Globodera pallida and Cyst- Potato (Solanum (Eves-van den Akker et G. rostochiensis nematode tuberosum) al., 2016) Pratylenchus penetrans, Root-lesion Alfalfa (Medicago sativa Plant root are degraded and invaded by the (Jones et al., 2013; P. thornei, P. neglectus, P. zeae, nematode L.), wheat (Triticum nematode, resulting in poor root growth, lesions, Vieira et al., 2019; P. vulnus, P. flakkensis, P. aestivum), soybean necrosis, plant browning and exposure to Mbatyoti et al., 2020; scribneri and P. coffeae (Glycine max) secondary infection by bacteria and fungi Rahman et al., 2020)

Ditylenchus destructor Potato-rot Garlic (Allium sativum) Nematode penetrate the garlic bulb or tubers (Zheng et al., 2016; nematode through lenticels by digesting starch and pectin Takada et al., 2020) for feeding, resulting in damages and rotting tubers and bulbs

304

Table 2A (Continue) Example of plants parasitic nematodes, the associated host and clinical symptoms

Parasitic nematode Parasite common Plant host Pathophysiology or clinical symptoms Reference species name Radopholus Similis Burrowing Ornamental plants, banana Root cortex cells are invaded by nematode for (Volcy, 2011; Xiaoling nematode (Musa sp.), blalck paper feeding. Cells are damaged and cavities fuse to et al., 2011; Huang et (Piper nigrum) form a red-brown lesion. Roots are damaged, al., 2019) resulting in diseases in the infected plants, e.g. 'toppling disease' in banana.

Bursaphelenchus Pine wilt nematode Pine tree (Pinus sp.) Nematode feeds on parenchymal cells, migrate (Kikuchi et al., 2011) Xylophilus within the plant tissues, resulting in wilting and plant death Rotylenchulus Reniform Tomato (Lycopersicon Nematode infection resulted in poor plant root (Robinson et al., 1997; Reniformis Nematode esculentum), cotton development, stunted tomatoes and reduced plant Redding et al., 2017; (Gossypium sp.), soybean productivity Zhang et al., 2018) (Glycine max)

Xiphinema index Dagger nematode Grapevines (Vitis vinifera) Nematode feeds on the growing tips of grapevine (Villate et al., 2012) plants and transmits Grapevine Fanleaf Virus (GFLV), resulting in distortion on leaves associated with a yellow mosaic pattern, abnormal branching, double nodding and softer fruits compared to normal Nacobbus Aberrans False-root knot Potato (Solanum tuberosum) Nematodes can either migrate into the plant roots (Eves-van den Akker et nematode resulting in lesions and necrosis or establish al., 2014) syncytium for feeding Aphelenchoides Rice white tip Rice (Oryza sativa), Nematodes can cause whitening of leaves, leading (Wang et al., 1993; Besseyi nematode strawberry (Fragaria to necrosis, distortion and disentangled leaves Wang et al., 2014; grandiflora), sweet corn enclosing the panicles Çelik et al., 2020) (Zea mays), sweet potato (Dioscorea esculenta)

305

Table 3A Example of fish parasitic nematodes, associated fish host and clinical symptoms

Zoonotic Parasitic nematode species Fish host Pathophysiology or clinical symptoms Reference diseases Philometra rara, P. saltatrix deep-water Haifa fish Nematode infections cause inflammation of Nematode (Moravec et al., 2017) (Hyporthodus haifensis) and visceral organs, haemorrhage, damaged blood invades an open blue fish (Pomatomus saltatrix) vessel and joint and poor fish development lesion in the human hand Hysterothylacium fabri, H. Brevoort (Conger myriaster), Nematodes migrate from the rectum to head, Commonly (Balbuena et al., aduncum, H. sinense, H. Tanaka snailfish (Liparis penetrate the host's muscle resulting in reduced associated with 2000; Guo et al., amoyense, H. zhoushanense, tanakae) host's health and mortality Anisakis sp. 2014; Chen et al., H. liparis infection 2018)

Anguillicoloides crassus Wild American eel (Anguilla Infected fish shows signs of inflammation, None (Pratt et al., 2019) rostrata), Japanese eel (Anguilla haemorrhage, lesion and fibrosis, resulting in japonica) swimming or migration inability

Cystidicola farionis Rainbow trout (Oncorhynchus Nematode cause damage on swim bladder, None (Menconi et al., 2019; mykiss), Pink salmon chronic irritation on fish, ulcers, inflammation Millane et al., 2019) (Oncorhynchus gorbuscha) with high nematode burden resulting in fish mortality

Anisakis simplex, A. Hake (Merluccius merluccius), Nematodes penetrate and migrate between the Anisakiasis and (Levsen et al., 2018; pegreffii, A. physeteris anchovies (Engraulis fish muscles resulting in damage in sarcomere allergic reaction Aibinu et al., 2019; encrasicholis), cod (Gadus structure and formation of granules. Nematode's Shamsi, 2019) morhua), mackerel (Scomber antigens are present surrounding the sarcomere japanicus), herring (Clupea structure and dispersed into the muscle of the harengus), sardine (Sardina infected fish. pilchardus) and tiger flathead (Neoplatycephalus richardsoni)

306

Table 4A Example of livestock parasitic nematodes, associated animal host and clinical symptoms

Parasite common Parasitic nematode species Animal host Pathophysiology or clinical symptoms Reference name Haemonchus contortus Barber's pole worm Sheep, goat Adult nematode feeds on blood resulting in (Mushonga et al., anaemia, oedema, pale mucous membrane, dark 2018) coloured faeces, weight loss, decreased appetite and animal death

Teladorsagia circumcincta Brown stomach worm Cattle, sheep, goat, Infected animals show reduced appetite, growth (Elseadawy et al., (Ostertagia circumcincta) human (zoonotic retardation, diarrhoea, gastroenteritis 2019; Ashrafi et al., disease) 2020) Ostertagia ostertagi Medium stomach Cattle/bovine Parasite infection resulted in appetite and weight (Mendez, 2018) worm loss, reduced reproductivity and milk production Trichostrongylus colubriformis, Black scour worm Cattle, sheep, Infected animals show reduced nutrient absorption (Ashrafi et al., 2020; T. vitrinus goats, human and poor growth and skeletal development due to Hidalgo et al., 2020) (zoonotic disease) enteritis or damages in the intestine Nematodirus battus Thin-necked intestinal Sheep Usually cause clinical symptoms (diarrhoea, (Taylor and Thomas, worm reduced weight and appetite, low protein in the 1986; McMahon et blood) in association with other parasites. al., 2017) Cooperia oncophora Hair worm Cattle, sheep Infected animals show weight loss and enteritis (Sun et al., 2020) Oesophagostomum radiatum Nodule worm Cattle, sheep, goat, Infected animals show intestinal lesion (nodule) (Delano et al., 2002; buffaloes due to inflammation resulting from nematode Deeba et al., 2019) penetration. Animals having alternating constipation and diarrhoea Dictyocaulus viviparus Lungworm Bovine/cattle Collapsed lung function, shortness of breath, (McCarthy and van cardiac lobes, exposure to secondary bacterial Dijk, 2020) infection, weight loss and reduced milk production

307

REFERENCES (FOR TABLE 1A – 4A)

Aibinu, I. E., Smooker, P. M., & Lopata, A. L. (2019). Anisakis nematodes in fish and shellfish-from infection to allergies. International Journal for Parasitology. Parasites and Wildlife, 9, 384-393. Ashrafi, K., Sharifdini, M., Heidari, Z., Rahmati, B., & Kia, E. B. (2020). Zoonotic transmission of Teladorsagia circumcincta and Trichostrongylus species in Guilan province, northern Iran: molecular and morphological characterizations. BMC Infectious Diseases, 20(1), 28. Attia, R. A., Tolba, M. E., Yones, D. A., Bakir, H. Y., Eldeek, H. E., & Kamel, S. (2012). Capillaria philippinensis in upper Egypt: has it become endemic? The American Journal of Tropical Medicine and Hygiene, 86(1), 126-133. Balbuena, J. A., Karlsbakk, E., Kvenseth, A. M., Saksvik, M., & Nylund, A. (2000). Growth and emigration of third-stage larvae of Hysterothylacium aduncum (Nematoda: Anisakidae) in larval herring Clupea harengus. International Journal for Parasitology, 86(6), 1271-1275. Bisoffi, Z., Buonfrate, D., Montresor, A., Requena-Méndez, A., Muñoz, J., Krolewiecki, A. J., Gotuzzo, E., Mena, M. A., Chiodini, P. L., Anselmi, M., Moreira, J., & Albonico, M. (2013). Strongyloides stercoralis: a plea for action. PLoS Neglected Tropical Diseases, 7(5), e2214-e2214. Bundy, D. A. P., de Silva, N., Appleby, L. J., & Brooker, S. J. (2020). 112 - Intestinal nematodes: Ascariasis. In E. T. Ryan, D. R. Hill, T. Solomon, N. E. Aronson, & T. P. Endy (Eds.), Hunter's Tropical Medicine and Emerging Infectious Diseases (Tenth Edition) (pp. 840-844). London. Çelik, E. S., Tülek, A., & Devran, Z. (2020). Development of a novel scale based on qPCR for rapid and accurate prediction of the number of Aphelenchoides besseyi in paddy rice. Crop Protection, 127, 104975. Chen, H.-X., Zhang, L.-P., Gibson, D. I., Lü, L., Xu, Z., Li, H.-T., Ju, H.-D., & Li, L. (2018). Detection of ascaridoid nematode parasites in the important marine food- fish Conger myriaster (Brevoort) (Anguilliformes: Congridae) from the Zhoushan fishery, China. Parasites & Vectors, 11(1), 274-274. Deeba, F., Qureshi, A. S., Kashif, A. R., & Saleem, I. (2019). Epidemiology of different gastrointestinal helminths in buffaloes in relation to age, sex and body condition of the host. Journal of Entomology and Zoology Studies, 7(1): 1533-1540. Delano, M. L., Mischler, S. A., & Underwood, W. J. (2002). Chapter 14 - Biology and Diseases of Ruminants: Sheep, Goats, and Cattle. In J. G. Fox, L. C. Anderson, F. M. Loew, & F. W. Quimby (Eds.), Laboratory Animal Medicine (Second Edition) (pp. 519-614). Burlington: Academic Press. Dupouy-Camet, J. (2000). Trichinellosis: a worldwide zoonosis. Veterinary Parasitology, 93(3-4), 191-200. Elseadawy, R., Abbas, I., Al-Araby, M., Hildreth, M. B., & Abu-Elwafa, S. (2019). First Evidence of Teladorsagia circumcincta Infection in Sheep from Egypt. Journal of Parasitology, 105(4), 484-490.

308

Eves-van den Akker, S., Laetsch, D. R., Thorpe, P., Lilley, C. J., Danchin, E. G. J., Da Rocha, M., Rancurel, C., Holroyd, N. E., Cotton, J. A., Szitenberg, A., Grenier, E., Montarry, J., Mimee, B., Duceppe, M.-O., Boyes, I., Marvin, J. M. C., Jones, L. M., Yusup, H. B., Lafond-Lapalme, J., Esquibet, M., Sabeh, M., Rott, M., Overmars, H., Finkers-Tomczak, A., Smant, G., Koutsovoulos, G., Blok, V., Mantelin, S., Cock, P. J. A., Phillips, W., Henrissat, B., Urwin, P. E., Blaxter, M., & Jones, J. T. (2016). The genome of the yellow potato cyst nematode, Globodera rostochiensis, reveals insights into the basis of parasitism and virulence. Genome Biology, 17(1), 124-124. Eves-van den Akker, S., Lilley, C. J., Danchin, E. G. J., Rancurel, C., Cock, P. J. A., Urwin, P. E., & Jones, J. T. (2014). The transcriptome of Nacobbus aberrans reveals insights into the evolution of sedentary endoparasitism in plant-parasitic nematodes. Genome Biology and Evolution, 6(9), 2181-2194. Fan, C.-K., Chuang, T.-W., Huang, Y.-C., Yin, A.-W., Chou, C.-M., Hsu, Y.-T., Kios, R., Hsu, S.-L., Wang, Y.-T., Wu, M.-S., Lin, J.-W., Briand, K., & Tu, C.-Y. (2019). Enterobius vermicularis infection: prevalence and risk factors among preschool children in kindergarten in the capital area, Republic of the Marshall Islands. BMC Infectious Diseases, 19(1), 536. Genchi, C., & Kramer, L. (2017). Subcutaneous dirofilariosis (Dirofilaria repens): an infection spreading throughout the old world. Parasites & Vectors, 10(2), 517- 517. Ghadirian, E., & Arfaa, F. (1975). Present status of trichostrongyliasis in Iran. The American Journal of Tropical Medicine and Hygiene, 24(6), 935-941. Ghodeif A.O. & Jain H. Hookworm. [Updated 2020 Mar 24]. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2019. Available at: https://www.ncbi.nlm.nih.gov/books/NBK546648/ Guo, Y.-N., Xu, Z., Zhang, L.-P., Hu, Y.-H., & Li, L. (2014). Occurrence of Hysterothylacium and Anisakis nematodes (Ascaridida: Ascaridoidea) in the Tanaka’s snailfish Liparis tanakae (Gilbert & Burke) (Scorpaeniformes: Liparidae). Parasitology Research, 113(4), 1289-1300. Hidalgo, A., Gacitúa, P., Melo, A., Oberg, C., Herrera, C., & Fonseca-Salamanca, F. (2020). First molecular characterization of Trichostrongylus colubriformis infection in rural patients from Chile. Acta Parasitologica, 1-6. Hopkins, D. R. (2020). 118 - Dracunculiasis. In E. T. Ryan, D. R. Hill, T. Solomon, N. E. Aronson, & T. P. Endy (Eds.), Hunter's Tropical Medicine and Emerging Infectious Diseases (Tenth Edition) (pp. 878-881). London. Huang, X., Xu, C.-L., Yang, S.-H., Li, J.-Y., Wang, H.-L., Zhang, Z.-X., Chen, C., & Xie, H. (2019). Life-stage specific transcriptomes of a migratory endoparasitic plant nematode, Radopholus similis elucidate a different parasitic and life strategy of plant parasitic nematodes. Scientific Reports, 9(1), 6277. Jones, J. T., Haegeman, A., Danchin, E. G., Gaur, H. S., Helder, J., Jones, M. G., Kikuchi, T., Manzanilla‐López, R., Palomares‐Rius, J. E., & Wesemael, W. M. (2013). Top 10 plant‐parasitic nematodes in molecular plant pathology. Molecular Plant Pathology, 14(9), 946-961.

309

Kikuchi, T., Cotton, J. A., Dalzell, J. J., Hasegawa, K., Kanzaki, N., McVeigh, P., Takanashi, T., Tsai, I. J., Assefa, S. A., Cock, P. J. A., Otto, T. D., Hunt, M., Reid, A. J., Sanchez-Flores, A., Tsuchihara, K., Yokoi, T., Larsson, M. C., Miwa, J., Maule, A. G., Sahashi, N., Jones, J. T., & Berriman, M. (2011). Genomic insights into the origin of parasitism in the emerging plant pathogen Bursaphelenchus xylophilus. PLoS Pathogens, 7(9), e1002219. Kim, J. H., Lim, H., Hwang, Y.-S., Kim, T. Y., Han, E. M., Shin, E.-H., & Chai, J.-Y. (2013). Gnathostoma spinigerum infection in the upper lip of a Korean woman: an autochthonous case in Korea. The Korean Journal of Parasitology, 51(3), 343- 347. Kucik, C. J., Martin, G. L., & Sortor, B. V. (2004). Common intestinal parasites. American family physician, 69(5), 1161-1168. Levsen, A., Svanevik, C. S., Cipriani, P., Mattiucci, S., Gay, M., Hastie, L. C., Bušelić, I., Mladineo, I., Karl, H., Ostermeyer, U., Buchmann, K., Højgaard, D. P., González, Á. F., Pascual, S., & Pierce, G. J. (2018). A survey of zoonotic nematodes of commercial key fish species from major European fishing grounds—Introducing the FP7 PARASITE exposure assessment study. Fisheries Research, 202, 4-21. Mantelin, S., Bellafiore, S., & Kyndt, T. (2017). Meloidogyne graminicola: a major threat to rice agriculture. Molecular Plant Pathology, 18(1), 3-15. Masonbrink, R., Maier, T. R., Muppirala, U., Seetharam, A. S., Lord, E., Juvale, P. S., Schmutz, J., Johnson, N. T., Korkin, D., Mitchum, M. G., Mimee, B., den Akker, S. E., Hudson, M., Severin, A. J., & Baum, T. J. (2019). The genome of the soybean cyst nematode (Heterodera glycines) reveals complex patterns of duplications involved in the evolution of parasitism genes. BMC Genomics, 20(1), 119. Mbatyoti, A., Daneel, M., Swart, A., Marais, M., De Waele, D., & Fourie, H. (2020). Plant-parasitic nematode assemblages associated with glyphosate tolerant and conventional soybean cultivars in South Africa. African Zoology, 55(1),1-16. McCarthy, C., & van Dijk, J. (2020). Spatiotemporal trends in cattle lungworm disease Dictyocaulus viviparus in Great Britain from 1975 to 2014. Veterinary Record, vetrec-2019-105509. McMahon, C., Edgar, H. W. J., Barley, J. P., Hanna, R. E. B., Brennan, G. P., & Fairweather, I. (2017). Control of Nematodirus spp. infection by sheep flock owners in Northern Ireland. Irish Veterinary Journal, 70(1), 31. Menconi, V., Pastorino, P., Cavazza, G., Santi, M., Mugetti, D., Zuccaro, G., & Prearo, M. (2019). The role of live fish trade in the translocation of parasites: the case of Cystidicola farionis in farmed rainbow trout (Oncorhynchus mykiss). Aquaculture International, 27(6), 1667-1671. Mendez, J. (2018). Bovine neutrophils release extracellular traps upon stimulation with Ostertagia ostertagi. Scientific Reports, 8, 17598. Miglioli-Galvão, L., Pestana, J. O. M., Lopes-Santoro, G., Torres Gonçalves, R., Requião Moura, L. R., Pacheco Silva, Á., Camera Pierrotti, L., David Neto, E., Santana Girão, E., Costa de Oliveira, C. M., Saad Abboud, C., Dias França, J. Í., Devite Bittante, C., Corrêa, L., & Aranha Camargo, L. F. (2020). Severe Strongyloides stercoralis infection in kidney transplant recipients: A multicenter case-control study. PLoS Neglected Tropical Diseases, 14(1), e0007998. Millane, M., Walsh, L., Roche, W. K., & Gargan, P. G. (2019). Unprecedented widespread occurrence of pink salmon Oncorhynchus gorbuscha in Ireland in 2017. Journal of Fish Biology, 95(2), 651-654. 310

Moravec, F., Chaabane, A., Neifar, L., Gey, D., & Justine, J.-L. (2017). Species of Philometra (Nematoda, Philometridae) from fishes off the Mediterranean coast of Africa, with a description of Philometra rara n. sp. from Hyporthodus haifensis and a molecular analysis of Philometra saltatrix from Pomatomus saltatrix. [Espèces de Philometra (Nematoda, Philometridae) parasites de poissons en Méditerranée au large de l’Afrique, avec une description de Philometra rara n. sp. de Hyporthodus haifensis et une analyse moléculaire de Philometra saltatrix de Pomatomus saltatrix]. Parasite, 24, 8-8. Mushonga, B., Habumugisha, D., Kandiwa, E., Madzingira, O., Samkange, A., Segwagwe, B. E., & Jaja, I. F. (2018). Prevalence of Haemonchus contortus infections in sheep and goats in Nyagatare District, Rwanda. Journal of Veterinary Medicine, 2018, 3602081. Niebuhr, C. N., Jarvi, S. I., Kaluna, L., Torres Fischer, B. L., Deane, A. R., Leinbach, I. L., & Siers, S. R. (2020). Occurrence of rat lungworm Angiostrongylus cantonensis in invasive Coqui frogs Eleutherodactylus coqui and other hosts in Hawaii, USA. Journal of Wildlife Diseases, 56(1), 203-207. Nomura, Y., Nagakura, K., Kagei, N., Tsutsumi, Y., Araki, K., & Sugawara, M. (2000). Gnathostomiasis possibly caused by Gnathostoma malaysiae. Tokai Journal of Experimental and Clinical Medicine, 25(1), 1-6. Papadi, B., Boudreaux, C., Tucker, J. A., Mathison, B., Bishop, H., & Eberhard, M. E. (2013). Halicephalobus gingivalis: a rare cause of fatal meningoencephalomyelitis in humans. The American Journal of Tropical Medicine and Hygiene, 88(6), 1062-1064. Pichakacheri, S. (2019). Guinea-worm Dracunculus medinensis infection presenting as a diabetic foot abscess: A case report from Kerala. The National Medical Journal of India, 32(1), 22-23. Pratt, T. C., O'Connor, L. M., Stacey, J. A., Stanley, D. R., Mathers, A., Johnson, L. E., Reid, S. M., Verreault, G., & , J. (2019). Pattern of Anguillicoloides crassus infestation in the St. Lawrence River watershed. Journal of Great Lakes Research, 45(5), 991-997. Rahman, M. S., Linsell, K. J., Taylor, J. D., Hayden, M. J., Collins, N. C., & Oldach, K. H. (2020). Fine mapping of root lesion nematode (Pratylenchus thornei) resistance loci on chromosomes 6D and 2B of wheat. Theoretical and Applied Genetics, 133(2), 635-652. Reaves, B. J., Wallis, C., McCoy, C. J., Lorenz, W. W., Rada, B., & Wolstenholme, A. J. (2018). Recognition and killing of Brugia malayi microfilariae by human immune cells is dependent on the parasite sample and is not altered by ivermectin treatment. International Journal for Parasitology: Drugs and Drug Resistance, 8(3), 587-595. Redding, N. W., Agudelo, P., & Wells, C. E. (2017). Multiple nodulation genes are up- regulated during establishment of reniform nematode feeding sites in soybean. Phytopathology, 108(2), 275-291. Richards Jr, F. O., Eigege, A., Umaru, J., Kahansim, B., Adelamo, S., Kadimbo, J., Danboyi, J., Mafuyai, H., Saka, Y., & Noland, G. S. (2020). The interruption of transmission of human onchocerciasis by an annual mass drug administration program in Plateau and Nasarawa States, Nigeria. The American Journal of Tropical Medicine and Hygiene, 102(3), 582-592. Robinson, A., Inserra, R., Caswell-Chen, E., Vovlas, N., & Troccoli, A. (1997). Rotylenchulus species: Identification, distribution, host ranges, and crop plant resistance. Nematropica, 27(2), 127-180. 311

Rusnak, J. M., & Lucey, D. R. (1993). Clinical Gnathostomiasis: case report and review of the English-language literature. Clinical Infectious Diseases, 16(1), 33-50. Sato, M., Yoonuan, T., Sanguankiat, S., Nuamtanong, S., Pongvongsa, T., Phimmayoi, I., Phanhanan, V., Boupha, B., Moji, K., & Waikagul, J. (2011). Short report: Human Trichostrongylus colubriformis infection in a rural village in Laos. The American Journal of Tropical Medicine and Hygiene, 84(1), 52-54. Shamsi, S. (2019). Seafood-borne parasitic diseases: a “one-health” approach is needed. Fishes, 4(1), 9, 1-11. Smiley, R. W. (2010). Root-lesion nematodes: Biology and management in Pacific Northwest wheat cropping systems. Available from https://catalog.extension.oregonstate.edu/pnw617 Sun, M.-M., Han, L., Zhou, C.-Y., Liu, G.-H., Zhu, X.-Q., & Ma, J. (2020). Mitochondrial genome evidence suggests Cooperia sp. from China may represent a distinct species from Cooperia oncophora from Australia. Parasitology International, 75, 102001. Takada, N., Sutoh, S., Toyota, M., Yamazaki, Y., Kitano-Yamashita, N., Ushida, C., & Yamashita, K. (2020). Methiin as a nematode attractant in Allium sativum. Canadian Journal of Chemistry, 98(3): 164-168. Taylor, D. M., & Thomas, R. J. (1986). The development of immunity to Nematodirus battus in lambs. International Journal for Parasitology, 16(1), 43-46. Vieira, P., Mowery, J., Shao, J., Eisenback, J. D., & Nemchinov, L. G. (2019). Cellular and transcriptional responses of resistant and susceptible cultivars of alfalfa to the root lesion nematode, Pratylenchus penetrans. Frontiers in Plant Science, 10, 971. Villate, L., Morin, E., Demangeat, G., Van Helden, M., & Esmenjaud, D. (2012). Control of Xiphinema index populations by fallow plants under greenhouse and field conditions. Phytopathology, 102(6), 627-634. Viswanath, A., & Williams, M. (2019). Trichuris Trichiura (Whipworm, Roundworm). In StatPearls [Internet]: StatPearls Publishing. Volcy, C. (2011). Past and present of the nematode Radopholus similis (Cobb) Thorne with emphasis on Musa: a review. Agronomía Colombiana, 29(3), 433-440. Wang, F., Li, D., Wang, Z., Dong, A., Liu, L., Wang, B., Chen, Q., & Liu, X. (2014). Transcriptomic analysis of the rice white tip nematode, Aphelenchoides besseyi (Nematoda: Aphelenchoididae). PloS One, 9(3), e91591-e91591. Wang, K., Tsay, T., & Lin, Y. (1993). The occurrence of Aphelenchoides besseyi on strawberry and its ecology in Taiwan. Plant Protection Bulletin (Taipei), 35(1), 14-29. Wang, Q.-P., Lai, D.-H., Zhu, X.-Q., Chen, X.-G., & Lun, Z.-R. (2008). Human angiostrongyliasis. The Lancet Infectious Diseases, 8(10), 621-630. WHO (World Health Organization) (2020) Lymphatic filariasis. https://www.who.int/news-room/fact-sheets/detail/lymphaticfilariasis. Accessed on 13 April 2020. Xiaoling, Z., Xingxing, J., & Senquan, Z. (2011). Identification and description of Radopholus similis extracted from three imported ornamental plants. Plant Protection, 37(2), 95-98.

312

Yang, F., Zhang, W., Gong, B., Yao, L., Liu, A., & Ling, H. (2019). A human case of Dioctophyma renale (giant kidney worm) accompanied by renal cancer and a retrospective study of dioctophymiasis. [Un cas humain de Dioctophyma renale (le ver géant du rein) accompagné d’un cancer du rein, et étude rétrospective de la dioctophymiose]. Parasite, 26, 22-22. Yang, S., Dai, Y., Chen, Y., Yang, J., Yang, D., Liu, Q., & Jian, H. (2019). A novel G16B09-like effector from Heterodera avenae suppresses plant defenses and promotes parasitism. Frontiers in Plant Science, 10, 66. Zhang, F., Wang, Y., Zhan, X., Dai, D., Guo, G., Guo, S., Sun, M., & Zhang, J. (2018). First report of Rotylenchulus reniformis on tomato in Henan, China. Plant Disease, 103(5), 1044-1044. Zhao, L., Shao, S., Chen, Y., Sun, X., Sun, R., Huang, J., Zhan, B., & Zhu, X. (2017). Trichinella spiralis calreticulin binds human complement C1q as an immune evasion strategy. Frontiers in Immunology, 8, 636. Zheng, J., Peng, D., Chen, L., Liu, H., Chen, F., Xu, M., Ju, S., Ruan, L., & Sun, M. (2016). The Ditylenchus destructor genome provides new insights into the evolution of plant parasitic nematodes. Proceedings of the Royal Society B Biological Sciences, 283(1835), 20160942.

313

APPENDIX II

1. Lysogeny Broth medium (LB10) (per litre of distilled water)* NaCl: 10 g Tryptone: 10 g Yeast Extract: 5 g

2. Nematode Growth Medium (NGM, per litre of distilled water)* NaCl: 3 g Tryptone: 2.5 g MgSO4: 1mM CaCl2: 1mM Cholesterol Solution: 5 μg/mL KH2PO4 Solution pH 6.0: 25 mM Add sterile water to 1 L

3. Bleaching solution NaClO 4%: 3 mL 5M NaOH: 1 mL Sterile H2O: 1 mL

4. Minimal salt 5x (M9, per litre of distilled water) Na2HPO4: 33.9 g KH2PO4: 15.0 g NaCl: 2.5 g NH4Cl: 5.0 g

5. Marine Broth (MB) (Difco Laboratories, US)* Bacto peptone: 5 g Yeast extract: 1 g Fe(III) citrate: 0.1 g NaCl: 19.45 g MgCl2 (anhydrous): 5.9 g NaSO4: 3.24 g CaCl2: 1.8 g KCl: 0.55 g Na2CO3: 0.16 g KBr: 0.08 g SrCl2: 0.034 g H3BO3: 0.022 g Na-silicate: 0.004 g NaF: 0.0024 g NH4NO3: 0.0016 g Na2HPO4: 0.008 g Adjust pH to 7.6

314

6. Potato Dextrose Agar (PDA) Potato extract: 4g Dextrose: 20g Agar: 15 g

7. S Basal buffer (per litre of distilled water) NaCl: 5.85 g K2 HPO4: 1 g KH2PO4: 6 g Cholesterol (5 mg/mL in ethanol): 1 mL

8. Phosphate Buffer Saline (PBS) NaCl: 8 g KCl: 0.2g Na2HPO4: 1.44g KH2PO4: 0.24g Distilled water: 1 litre pH adjusted to 7.4 with HCl (1.0 M)

9. Skim milk agar* Skim milk solution (2% w/v) in distilled water (autoclaved separately at 110°C for 10 minutes) LB10 agar (components are described as above and autoclaved separately at 121°C for 15 minutes) Skim milk solution was added to the agar mixture at 50°C prior to plating

10. Gelatine agar* Gelatine 1.2% (w/v) Peptone 0.4% (w/v) Yeast extract 0.1% (w/v) Gelatine was dissolved in 50°C prior to mixing with the agar solution and autoclaving at 121°C for 15 minutes

11. Tween 80 agar* Tween 80 2% (v/v) Victoria Blue B 0.01% (v/v) Tween 80 solution and LB10 agar was autoclaved separately at 121°C for 15 minutes while the Victoria Blue B solution was filter-sterilized through the 0.2 µm pore size- filter (Amicon). The LB10 agar solution, Tween 80 and Victoria Blue B were mixed at 50°C prior to plating.

12. Chitin agar* Chitin: 20g HCL 32%: 300 mL LB10 agar (components are described as above) Chitin were mixed with HCl 32% with continuous stirring for two hours at 4°C. The suspension was repeatedly washed with 1 L water five times and filtered through coarse filter paper. The pH of the colloidal chitin suspension was adjusted to pH 7.0 by adding NaOH (5.0 M) followed by three times washing with distilled water for desalting. The suspension was then centrifuged at 8000 rpm for 10 minutes and the 315

precipitated colloidal chitin was supplemented into the LB10 agar (2% w/v) prior to autoclaving at 121°C for 15 minutes.

13. Coomassie blue staining Coomassie Blue G-250: 1g Methanol: 50 % (v/v) Glacial acetic acid: 10% (v/v) Distilled water: 40% (v/v)

14. Running Buffer (5x) Tris base: 15.1g Glycine: 72g SDS: 5g Distilled water: 850 mL pH: 8.3

15. Sample Loading Buffer (4x) (10 mL stock) Tris-HCl (1.0 M, pH 6.8): 2 mL SDS: 0.8g Glycerol 10% (v/v): 4 mL β-mercaptoethanol (14.7M): 0.4 mL EDTA (0.5 M): 1 mL Bromophenol Blue: 8 mg

16. Destaining solution Methanol: 100 mL Glacial acetic acid: 100 mL Distilled water: 800 mL

17. Lowry’s method (reagents preparation) Solution A: CuSO4.5H2O (0.5g) and Na3C6H5O7.2H2O (1g) were dissolved in 100 mL of distilled water Solution B: Na2CO3 (20g) and NaOH (4g) were dissolved in 100 mL distilled water Solution C: Solutions A and B were mixed (ratio A:B; 1:50) Solution D: Folin-Ciocalteus Phenol reagent was mixed with distilled water (1 : 1)

18. Wash Buffer A Triton 100 (0.5% v/v) Tris 5 mM EDTA 50 mM pH 8.0

19. Wash Buffer B Glycerol 5% (v/v) Tris 5 mM EDTA 50 mM pH 8.0

316

20. Ethanol/EDTA precipitation method Add 10 µL of TE buffer and 5 μL of EDTA (125mM). Add 70 µl of 95% ethanol. Mix well. The, incubate at room temperature for 15 min. Pellet the DNA by centrifugation at 3,000 x g at 4˚C for 30 minutes. Supernatant was discarded and DNA pellet was washed in 70 % ethanol to remove salts. The DNA was pelleted using centrifugation at 1,650 x g at 4˚C for 15 min. The supernatant was discarded and DNA was vacuum dried using vacuum dessicator for 5 minutes.

21. SOC media Tryptone: 20g Yeast extract: 5g NaCl: 0.5g Glucose (20 mM) (filter sterilised) Tryptone, yeast extract and NaCl were dissolved in distilled water (adjusted to final volume 1 litre, pH 7.0) and autoclaved at 121°C for 15 minutes. Glucose solution was added when the initial solution was cooled to 55°C.

22. 2x Yeast Tryptone (2YT) Broth Media Tryptone: 16g Yeast extract: 10g NaCl: 5g Distilled water: 1 litre pH adjusted to 7.0 using NaOH (5.0M) and autoclaved at 121°C for 15 minutes.

23. Soft marine agar Marine Broth (MB) (Difco Laboratories, US) Distilled water Agar 0.6% (w/v)

24. LM Buffer 60 µg/mL levamisole in M9 buffer

*Added with 1.5% (w/v) Agar Bacteriological (Oxoid) to make solid media

317

APPENDIX III

1.2

y = 2.1281x + 0.0738 1 R² = 0.9847

0.8

0.6 OD (750 nm) 0.4

0.2

0 0 0.1 0.2 0.3 0.4 0.5 0.6

BSA (mg/mL)

Figure 1 Standard curve for protein quantification generated from a serially diluted Bovine Serum Albumin (BSA)

318

APPENDIX IV

Figure 1 pBAD24 plasmid map sequence. The figure was derived from Addgene at https://www.addgene.org/vector-database/1845/

319

hp1

Figure 2 pBAD24hp1 plasmid map sequence. The figure was generated using Snapgene software. The hp1 amplified DNA fragment (highlighted in red) was ligated downstream the araBAD promoter

320