<<

CRISPR-based innovative genetic tools for control of gambiae mosquitoes

The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters

Citation Smidler, Andrea. 2019. CRISPR-based innovative genetic tools for control of mosquitoes. Doctoral dissertation, Harvard University, Graduate School of Arts & Sciences.

Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:42029729

Terms of Use This article was downloaded from Harvard University’s DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http:// nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA

CRISPR-based innovative genetic tools for control of Anopheles gambiae mosquitoes

A dissertation presented

by

Andrea L. Smidler

to

The Committee on Higher Degrees in Biological Sciences in

in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

In the subject of

Biological Sciences in Public Health

Harvard University

Cambridge, Massachusetts

April 2019

i

© 2019 – Andrea Smidler

All rights reserved

ii

Dissertation Advisor: Dr. Flaminia Catteruccia, Dr. George Church Andrea Smidler

CRISPR-based innovative genetic tools for control of Anopheles gambiae mosquitoes ABSTRACT

Malaria and other -borne diseases pose an immense burden on mankind. Since the turn of the century, control campaigns have relied on the use of -impregnated bed nets and indoor residual sprays to stop Anopheles mosquitoes from transmitting the parasite. Although these are our best strategies to control the spread of disease, wild mosquito populations are developing resistance to at an alarming rate, making disease control increasingly challenging. In the search for new powerful strategies aimed at controlling malaria-transmitting Anopheles populations, we can now exploit a suite of powerful genome engineering tools to control wild populations and monitor releases. In this dissertation we utilize CRISPR/Cas9 technology and other genetic engineering tools in Anopheles gambiae to generate genetically sterile males for population suppression, to assess the feasibility of developing evolutionarily stable gene drives for population replacement, and to expand the genetic toolkit for field releases of genetically modified mosquitoes.

We develop CRISPR/Cas in A. gambiae for male genetic sterilization and for basic biological study.

Using this system we generate a line of mutant mosquitoes with deletions in Zero Population Growth

(ZPG), a gene critical for germ cell development. Resulting male mutants show no sperm in the testes and sterilize the females with which they mate, demonstrating that similar systems could be adapted for use in Sterile-Insect Technique (SIT)-like release campaigns.

We also test whether CRISPR/Cas9 can facilitate the sustainable and stable spread of gene drives in natural mosquito populations. Specifically, we design gene drives that have the potential of being evolutionarily stable by insertion into haplolethal ribosomal genes. To facilitate this goal, we create docking lines via a novel knockin technology for insertion of complex DNA templates into genetically

iii intractable loci. We identify multiple challenges associated with such systems, including the occurrence of Minute-like mutant phenotypes that present severe fitness costs when targeting haploinsufficient genes, a general decay in gRNA function over time that has consequences for all gene drives designed to date, and the critical need for precisely controlling Cas9 expression to avoid large fitness costs.

Finally, we develop and validate a novel transgenic tool for monitoring GM field releases. We generate transgenic lines expressing a fusion of a fluorescent marker with a male seminal protein specifically in male reproductive tissues. Incorporation of this fluorescence fusion into the mating plug, which is transferred to the females during mating, allows the visual identification of successful mating events for a few hours after copulation. This transgenic tool enables effective monitoring of GM male mating competitiveness in field trials, overcoming current limitations.

The work outlined here significantly expands the genetic toolkit for the manipulation of Anopheles mosquitoes, facilitating the implementation of genetic control strategies aimed at malaria-transmitting vector populations.

iv

TABLE OF CONTENTS ABSTRACT …………………………………..…………………………………………………………..………………………………………. iii TABLE OF CONTENTS ……………………………..…………………………..………………………………………………………….…. v LIST OF FIGURES AND TABLES …………..……………………………………………………………………………………….…… viii GENOTYPE NOMENCLATURE……………………………………………………………………………………………………………… x ACKNOWLEDGEMENTS ……………..………………..…………………………………………………………………………….…….. xi

CHAPTER 1: Introduction …………..……………………………………………………………………………………………………… 1 1.1 The burden of malaria and its current control …………………………………………………………………. 2 1.2 A brief and its eradication from and the United States – a precedent for mosquito elimination to control malaria……………………………………………….. 4 1.3 Mosquito genetic control – An introduction …………………………………………..………………………. 6 1.4 Population suppression systems ……………….…………………………………………..……………………….. 8 1.4 A. for mosquito population suppression ………………..…………………..... 8 1.4 B. RIDL-based transgenic population suppression systems ………………………..………………………… 9 1.5 Population replacement systems ………………………………………………………..…………………….….. 10 1.5 A. Endonuclease gene drives – Mode of action and history of development .………….…..……. 11 1.5 B. TALEN and ZFN-mediated gene drives ………………………………………..………..…………….…………. 12 1.5 C. CRISPR-mediated gene drives …………………………………………………..………..……………………….... 13 1.5 D. Safety considerations concerning gene drive……………………………………….………………………….. 14 1.5 E. Anti-parasite cargoes for population replacement …………………………………………………..……. 15 1.6 The problem of drive-resistant alleles (DRAs) ..…………………………………………………………….. 16 1.7 Learning from other CRISPR gene drives ..……………………………………………………………………… 18 1.8 Fitness considerations for gene drives …………………………………………………………………………… 20 1.8 A. CRISPR off-target mutagenesis ..…………………………………………………………………………………… 21 1.8 B. Gene drive-imposed evolutionary selective pressures ..…………..……………………………………. 22 1.9 Designing evolutionarily stable gene drives for population replacement ..…………………….. 25 1.10 Mosquito genetic control strategies need more tools ..……………………………………………….. 26 1.11 Summary of Aims ..……………………………………………………………………………………………………… 28 1.12 References ..………………………………………………………………………………………………………………… 30

CHAPTER 2: CRISPR-mediated germline knockouts for genetic sterilization of male Anopheles gambiae ………………………………………………………..………………………………. 46 ABSTRACT …………………………………………………………………………………………………………………………… 47 INTRODUCTION ………………………………………………………………………………………………………………….. 48 RESULTS ……………………………………………………………………………………………………………………………… 49 2.1 Generation of transgenic CRISPR-expression lines …………………………………………….. 50 2.2 ZPG mutants fail to develop normal testes ……………………………………………………….. 51

v

2.3 Male ZPG mutants are highly sterile ……………………………………………………………….... 52 2.4 Characterization of fertile ZPG mutant males ……………………………………………………. 53 2.5 Female ZPG mutants display severe ovarian atrophy ………………………………………… 54 DISCUSSION ………………………………………………………………………………………………………………………… 56 FUTURE DIRECTIONS …………………………………………………………………………………………………………… 58 METHODS …………………………………………………………………………………………………………………………… 61 REFERENCES ……………………………………………………………………………………………………………………….. 65 CHAPTER 3: Advancement towards developing evolutionarily stable gene drives for Anopheles gambiae population replacement …………………………………………………………. 69 ABSTRACT …………………………………………………………………………………………………………………………… 70 INTRODUCTION ………………………………………………………………………………………………………………..... 71 RESULTS ………………………………………………………………………………………………………………………………. 72 3.1 Gene drive design ……………………………………………………………………………………………… 72 3.2 Gene drive docking site locus identification ………………………………………………………. 72 3.3 Designing two docking sites in RpL11-Rpt1 – benefits and drawbacks of each design …………………………………………………………….. 74 3.3 A. Designing dRPLT – a gene drive docking site with Rpt1 knocked-out …………..…… 76 3.3 B. Designing dRPLTu – a gene drive docking site with Rpt1 function maintained…. 77 3.4 Knockin of dRPLT docking line into RpL11-Rpt1 by HDR …………………………………….. 77 3.5 Developing transgenic Interlocus Gene Conversion (IGC) for knockin into genetically intractable loci ……………………………………………………….77 3.5 A. Generating dRPLT by CrIGCkid ………………………………………………………………………….. 79 3.5 B. Generating dRPLTu by CrIGCkid …………….…………………………………………………………. 80 3.5 C. CrIGCkid does not occur in females, and infertility phenotypes lessen over time. 84 3.6 Characterization of dRPLT and dRPLTu docking lines …………………………………………. 84 3.6 A. dRPLT is homozygous lethal, experiences proteasome dysfunction, and is likely a Rpt1 knockout ………………………………………………………………………………….. 84 3.6 B. dRPLTu homozygotes are viable but display Minute-like fertility phenotypes ..… 85 3.7 Gene drives in dRPLT and dRPLTu may be lethal ……………………………………………….. 87 3.7 A. Inserting B2GD into dRPLT or dRPLTu results in death of all transgenic individuals ……………………………………………………………………………………. 87 3.7 B. Lethality of gene drives in docking sites may be due to aberrant Cas9 overexpression ………………………………………………………………………….…… 89 3.8 Developing drive-enabling germline-Cas9 promoters independent of docking sites ……………………………………………………………………………. 92 3.8 A. Vasa2 transgenic promoter is too leaky for drive ……………………………………………. 92 3.8 B. A Vasa protein fusion to fine-tune Cas9 expression ...... 94 3.8 C. Two versions of the ZPG promoter for gene drive Cas9 expression ………………….. 95

vi

DISCUSSION ………………………………………………………………………………………………………………………… 98 METHODS ……………………………………………………………………………………………………………………….... 106 REFERENCES ………………………………………………………………………………………………………………………. 122

CHAPTER 4: A transgenic tool to assess Anopheles mating competitiveness in the field …………………. 127 ABSTRACT …………………………………………………………………………………………………………………………. 128 INTRODUCTION …………………………………………………………………………………………………………………. 129 RESULTS ……………………………………………………………………………………………………………………………. 131 4.1 Generation and characterization of the plugin-tdTomato transgenic line.………… 131 4.2 Plugin-tdTomato localizes to vesicles within the anterior compartment of the MAGs ………………………………………………………………………………………………….… 134 4.3 Plugin-tdTomato is sexually transferred to the female atrium in the mating plug ...... 136 4.4 Testing PluTo males for in vivo assessment of successful mating …………………….. 136 4.5 Expression of the plugin-tdTomato fusion protein does not affect the mating competitiveness of PluTo males and does not impair female reproductive fitness ……………………………………………………………………………………….. 137 DISCUSSION ……………………………………………………………………………………………………………………... 139 CONCLUSIONS……………………………………………………………………………………………………………………. 141 METHODS ………………………………………………………………………………………………………………….……... 141 REFERENCES ……………………………………………………………………………………………………………………… 148

CHAPTER 5: Discussion ………………………………………………………………………………………………………………..… 152 5.1 Overview …………………………………………………………………………………………………………………….. 153 5.2 CRISPR-mediated genetically sterile males for Sterile Insect Technique ……………..………… 153 5.3 Developing a transgenic tool to assess Anopheles mating competitiveness in the field …………………………………………………………………………………………………………………… 156 5.4 Developing evolutionarily stable gene drives for Anopheles gambiae population replacement ………………….……………………………………………. 157 5.5 Concluding remarks……………………………………………………………………………………………………… 161 5.6 REFERENCES………………………………………………………………………………………………………………… 162

APPENDIX A: Plasmid and transgene maps…………………………………………………………….………………………. 166 APPENDIX B: Full Text: Targeted Mutagenesis in the Malaria Mosquito Using TALE Nucleases……….. 168 APPENDIX C: Full Text: Engineering the control of mosquito-borne infectious diseases…………………… 178 APPENDIX D: Full Text: Concerning RNA-guided gene drives for the alteration of wild populations... 188

vii

LIST OF FIGURES AND TABLES CHAPTER 1. Figure 1.1 | Dynamics of population suppression and population replacement vector control strategies …..…………………………………………………………………………………………………………. 7 Figure 1.2 | A basic diagram of ZFN, TALEN, and CRISPR/Cas9 DNA cleavage……………………….…. 14 Figure 1.3 | Biased inheritance of Gene drives through mosquito populations and the genetic outcomes of cleavage and homing………………………………………………………… 17

CHAPTER 2. Figure 2.1 | Transgenic map and fluorescence properties of VasCas9 and gZPG transgenic lines…………………………………………………………………………………...... 50 Figure 2.2 | The spermless male testicular phenotype…………………………………………………………….. 51 Figure 2.3 | A composite image of mutations observed in {+/gZPG;+/VasCas9} male testes……. 52 Figure 2.4 | Genetically spermless males cause significant infertility in females…………………….… 53 Figure 2.5 | ZPG mutant females have significantly ablated ovarian development and are defective in 20E synthesis ………………………………………………………………………………..…… 55 Figure 2.6 | Schematic for the Inducible Eunuch system…………………………………………….……………. 60

CHAPTER 3. Figure 3.1 | An evolutionarily stable CRISPR gene drive design for population replacement ….. 73 Figure 3.2 | Constructing gene drives into RpL11-Rpt1……………………………………………………………. 75 Table 3.1 | Summary and comparison of characteristics of the dRPLT and dRPLTu gene drive docking sites………………………………………………………..…………………………………….… 76 Figure 3.3 | Putative mechanism for IGC to generate gene drive docking lines …………………….… 79 Table 3.2 | Summary of IGCdRPLTupB insertion sites……………………………………………………………….. 80 Figure 3.4 | PCR confirmation of dRPLTu knockin into RpL11-Rpt1 following IGC…………………….. 81 Figure 3.6 | Characterization of CrIGCkid to create dRPLTu……………………………………………………… 83 Figure 3.7 | Fitness characterization of dRPLT…………………………………………………………………………. 85 Figure 3.8 | dRPLT and dRPLTu individuals display many hallmark Minute phenotypes…………… 86 Figure 3.9 | Gene drive and distal gene drive transgene maps developed in this work……………. 88 Table 3.3 | Number embryos injected, transgenics recovered and survived to adulthood………. 88 Figure 3.10 | Western blot analysis for Cas9 expression in drive lines……………………………………… 89 Figure 3.11 | No biased inheritance of mNosGDdRPLT………………………………………………………………… 90 Figure 3.12 | B2dGDpB does not facilitating drive of dRPLTu in males……………………………………… 91 Figure 3.13 | Rescuing mNosGDdRPLT Cas9 function with Vasa-Cas9 results in death or infertility of all females and infertility in males………………………………………..……….… 94

viii

Figure 3.14 | A Vasa fusion distal drive, VFdGDpB is fertile but fails to facilitate drive of dRPLTu. 95 Figure 3.15 | A ZPG2 distal drive, ZPG2dGDpB fails to facilitate drive of dRPLTu……………………….. 96 Figure 3.16 | A ZPG1 distal drive, ZPG1dGDpB causes complete female infertility, and incomplete male infertility which decreases in subsequent egg-lays …….……..….. 97 Figure 3.17 | Summary of Cas9 Goldilocks Conundrum……………………………………………………….… 105

CHAPTER 4 Figure 4.1 | PluTo transgene design and male phenotype……………………………………………………... 132 Figure 4.2 | Endogenous Plugin and PluTo transgene expression…………………………………………… 133 Figure 4.3 | Endogenous Plugin localization in the MAGs ……………………………………………………… 134 Figure 4.4 | Plugin-tdTomato localization in the MAGs and mating plug……………………………….. 135 Figure 4.5 | PluTo transgenic males allow in vivo assessment of male mating success within the female…………………………………………………………………………………………………………… 137 Figure 4.6 | Reproductive fitness of females mated to PluTo males…………………………………..….. 138

ix

GENOTYPE NOMENCLATURE

To clarify the transgenes and genotypes outlined herein, here are presented the nomenclature rules used in this work. They are inspired by Drosophila nomenclature standards, but have been modified for the specific needs of the work undertaken here while attempting to maximize clarity and ease of readability.

Italics are reserved for gene and locus names, with capitalized letters denoting wild type loci, and lowercase letters denoting a disrupted or knockout gene product, occasionally specified with Δ. Sterile males are termed ‘ZPG mutants’, because as mosaics, they cannot be considered true knockouts.

Transgene names vary, but attempt to convey important information about the transgene’s function, purpose, and properties. The first time the transgene name is discussed, the abbreviation for its name is listed. All relevant transgene maps can be found in APPENDIX A, and for sequence specifics, the email

([email protected]) can be reached for inquiries ad infinitum.

Superscripts attempt to convey locus information for transgenes. For example if a drive is inserted into the dRPLT docking line, it will be conveyed as a dRPLT superscript adjacent to the transgene name. Therefore the mNosGD line inserted in dRPLT is denoted mNosGDdRPLT. Similarly if a transgene is piggyBac based, and the insertion(s) has been identified and purified as an iso-female family, it will be denoted simply as pB#. If an experiment was performed with a mixed population with a single clonal transgene inserted within many varied and unspecified sites in the genome in a mixed population, it will be denoted pBmix.

Mixed genotypes and phenotypes follow similar conventions to those used in Drosophila. Genotypes are fully contained within brackets {}, and the genotype of each locus is denoted over a slash ‘/’ . A wild type allele is conveyed with the standard ‘+’ symbol, and the genotypes of separate loci are separated by a semicolon ‘;’. Therefore the genotype of an individual heterozygous for each of the dRPLT and VasCas9 transgenes will be denoted {+/dRPLT ;+/VasCas9}. Similarly an individual transheterozygous for the mNosGD inserted into dRPLT, and unintegrated dRPLT will be denoted {mNosGDdRPLT/dRPLT}.

x

ACKNOWLEDGEMENTS

Thank you:

To all my scientific mentors through the years. To my childhood teachers who first instilled in me a reverence for genetics, to my genetics summer camp counselors, and to my undergraduate advisors, Dr.

Jeff Dangl and Dr. Nuria S. Coll who believed in me enough to let me loose in a lab. The seemingly small doors you opened have become monumental to me. To Eric Marois, who first inspired my love of mosquito genetic engineering - I can’t imagine a life in which we did not meet.

To Flaminia. Who saw the potential behind that uncompromising awkwardness. To your unending patience and support despite our disagreements and my bullheadedness. For pushing me to be better, but giving me the space to follow my heart, and for letting me learn the hard way. For the opportunities you’ve provided, and for the lessons you’ve taught. I thank you from the bottom of my heart, I would never be the scientist I am today without you.

To Kevin Esvelt. For our inspiring conversations, and belief in the impossible. For showing me the immense motivation required to change the world, and for bringing me along as you do so. I am forever grateful for your guidance, inspiration, and generosity.

To George Church. For showing me how to think like an innovator, and how to embrace the improbable.

For your phenomenal mentorship, and the examples of excellence you provide. For your truly generous and unwavering support, and for teaching me that I belong at the table.

To other members of my Dissertation Advisory Committee: Dr. Dyann Wirth, Dr. Matthias Marti, and Dr.

Dan Neafsey. Thank you for keeping me focused on the important questions, and for providing your incredible feedback and guidance over the years. My project would not be where it is today without your

xi wisdom. To my Dissertation Defense Committee: Dr. Manoj Duraisingh, Dr. Jay Mitchell, Dr. Conor

McMeniman, and Dr. Jeff Dvorin, thank you for your time and expertise as I defend my PhD.

Thank you to the funding that made this work possible, the Ruth L. Kirschtein NIH Fellowship.

Thank you to my fellow lab mates. You all have been amazing, and I’ve learned so much from you. Thank you for the support and comradery through the years. You each hold a special place in my heart, and it is impossible to properly thank you all. Especially to Rob, Emily, Manu, and Eryney for all your guidance, support and help through the years.

To my friends. To Mike and RJ, for showing me what academic dedication looked like. For teaching me how to write from the ground up, and for helping me discover my calling. To Gabe for being the best listener alive; for being inspiring, encouraging and pragmatic somehow all simultaneously, and to Erica,

Vanessa and Eli, for being my Bean dip therapist, and a shoulder to cry on, and for keeping my teeth sharp.

To my family. Mom and Dad, I would never be the woman I am today without you. For teaching me how to be strong, smart, and independent, and for tolerating my crazy childhood years. For the unwavering love and laughs, and for always being there through it all.

To Dan, my first nerd partner in crime. From competing at doing Punnett Squares at the breakfast table, to constructing suits of chain-maille. You have kept me grounded, humbled, and inspired.

To Alex, my love. For being my rock [troll]. When the world around me is full of variables, you are my constant. For teaching me how to love life, and how to live it well. For showing me the joy that serenity can bring, and for teaching me the limits of my eye-rolling capabilities. For loving me as a garbage person, as well as a knockout, and for being the best father our dogter could ask for. You have no idea how much

I look forward to our lives together.

xii

CHAPTER 1. Introduction

Respective Contributions: This chapter was written by Andrea Smidler, with editorial input from Flaminia

Catteruccia. Portions of this chapter include work from published Research and Review Articles co-written by Andrea Smidler in collaboration with colleagues. This chapter includes some images, phrases, and ideas from these articles with permission for use under Creative Commons Attribution 4.0. These articles are included in full in the Appendix for reference.

APPENDIX C

1Gabrieli, P., Smidler, A. & Catteruccia, F. Engineering the control of mosquito-borne infectious diseases. Genome biology 15, 535, doi:10.1186/s13059-014-0535-7 (2013). APPENDIX D

2 Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife, e03401, doi:10.7554/eLife.03401 (2014). 1.1 The burden of malaria and its current control

Vector-borne diseases have plagued humanity since the dawn of time, possibly killing half of all humans who have ever lived3,4. The burden these diseases impose on mankind has left scars across our genomes as a testament to their prevalence, persistence, and pandemicity5, and despite humanity’s best efforts, we have yet to eradicate even one. Some of the most destructive pathogens include the Dengue, Yellow

Fever, Zika, and Chikungunya arboviruses, but the malaria-causing protozoan parasites are the most deadly6. Spread by the bite of an anopheline mosquito, these parasites cause the death of approximately half a million people annually with another 200 million incapacitated for days7 – grinding lives and productivity to a halt, and affecting some countries’ GDP by as much as 6%8.

Since the beginning of this century, funding for malaria research and control has increased almost 10- fold9. Despite recent strides towards developing effective prophylactics and vaccines, the World Health

Organization (WHO) predicts that “ is the only intervention that can reduce transmission from very high levels to close to zero”10. Increased application of vector-control measures has contributed to a staggering reduction in mortality rates11, with the use of long lasting insecticide-treated nets (LLINs) and (IRS) contributing to over 75% cases averted since the turn of the century12,13.

However, these once-reliable control methods are becoming increasingly ineffective due to insecticide resistance mechanisms emerging in mosquito populations.

LLINs are fine-mesh nets hung above beds which protect people from mosquito bites when they are asleep14, and when impregnated with insecticides they provide further protection to the community15.

This intervention targets vector populations which blood-feed indoors and at night, generally targeting a female once per gonotrophic (egg development) cycle, but neglect mosquitoes using different feeding strategies. Following a ramp-up of LLINs distribution, it is estimated that nearly half of the malaria at-risk population are now protected by bednets16, but their widespread use has induced resistance to

2 pyrethroids in mosquitoes17-21, to the point that has become difficult to find populations that are fully susceptible22. Some reports even suggest that mosquitoes may be evolving novel feeding behaviors to circumvent these interventions23-25, suggesting we may be reaching the limit of LLIN effectiveness.

Similar considerations are applicable to the use of IRS, which relies on insecticide application to interior house walls where female mosquitoes usually rest after taking a blood meal. Because this intervention targets the interiors of human dwellings, disease-transmitting populations which feed and rest outdoors are not affected by this strategy. Due to the reduced contact risks for humans compared to insecticides on LLINs, a number of insecticide classes are approved for IRS use, including organophosphates and carbamates. Even the highly toxic and persistent DDT can be utilized for malaria control purposes in some cases of extreme resistance to all other insecticides26. Currently only an estimated 7% of the African at- risk population lives in homes treated with IRS16, however these numbers are increasing following a shift towards using non-pyrethroids17,21, in response to increasing resistance amongst wild populations.

Indeed, the majority of anophelines are now resistant to pyrethroids27, and resistance rates to carbamates and organophosphates are on the rise28. Despite best practices recommending annual rotation of IRS insecticides to limit the emergence of resistance21, alarmingly mosquito populations have emerged with resistance to all four classes of insecticides available for malaria control29,30 – with some resistance mechanisms causing resistance to multiple classes simultaneously29,31. Since 2010, 68 countries have reported resistance to at least one class of insecticide with 57 of those countries reporting resistance to

2 or more14. Although this suggests that the lethal effect of even newly-developed insecticides may be thwarted at an unexpectedly fast time-scale, both LLINs and IRS remain the core intervention strategies used to control malaria worldwide32 making the development of novel vector control technologies – especially those capable of targeting outdoor biting and resting populations – increasingly urgent.

3

1.2. A brief history of malaria and its eradication from Italy and United States - A precedent for mosquito elimination to control malaria

While deliberately eradicating an entire species seems an ecological nightmare, it is not without precedent. Malaria has so significantly changed the course of human history that it can be argued to have shaped the fate of empires, explaining the urgency with which many countries have sought to eliminate this disease through elimination of its mosquito vector. We will briefly examine malaria’s impact on human history, and the control efforts in two countries where the disease was successfully eliminated, in an attempt to glean insights on how similar campaigns may look like in the future, what the ecological impacts might be, and which motivations may drive them.

Malaria was endemic to Italy throughout most of recorded history, with first-century Roman scholars understanding enough about its to encourage building of homes on high land away from swamps33. In fact, “having forsaken its best farmland to malaria, ancient Rome could not feed itself, but

[following expansion of the empire] the riches of conquest paid for grain, olives, fish sauce, and oil imported from North Africa. It paid for elaborate aqueducts, allowing wealthy Romans to move away from the most mosquito-ridden water’s edge”34. Malaria posed such an immense burden that Romans would pray to the demon goddess of , Febris, in three dedicated temples around the city to attempt to beg for respite34,35. Therefore it is no surprise that when the technology arose, the Italian people sought to eradicate the malaria mosquito at their first opportunity36. Cases were treated with , larvae were killed by application of petroleum derivatives and “Paris Green”(copper acetoarsenite), and breeding sites removed through land reclamation36. These efforts resulted in significant disease reduction, with some regions falling to almost zero by 193937. However, only once DDT became available after the conclusion of WWII was eradication complete with virtually all cases eliminated by 194836.

4

At approximately the same time the United States began a similar eradication campaign38. Malaria had raged endemically in the United States since before it was a nation, with some epidemics occurring as far north as Sheffield Massachussets39. However most scientists agree that North America was free of malaria before colonization34, with one of – if not – the first cases being imported by a Jamestown settler infected with Plasmodium vivax40. Later, the more deadly was introduced during the slave trade41 and was spread endemically by native Anopheles quadrimaculatus, Anopheles crucians, and

Anopheles pseudopunctipennis42. It was such a significant part of American life that it may have even contributed to the formation of the nation, with the environmental historian J.R. McNeill once writing;

“Those tiny amazons [mosquitoes] conducted covert biological warfare against the British army”43. When the British Army attempted to invade South Carolina in 1780 they were ultimately thwarted by malaria, succumbing to due to their lack of exposure to the parasite through childhood44. The disease burden in the region was so great that it led Johan David Shoepf to write in 1788; “Carolina was in the spring a paradise, in the summer a hell, and in the autumn a hospital”45.

The battle with malaria raged through the Civil War, and “mosquito soldiers”46 may have even prolonged it. Despite the Confederate army being greatly outnumbered, the malaria resistance its soldiers had acquired from repeated exposure gave them an advantage against the Union when fighting in Southern territories, potentially extending the war by years46. Therefore almost 80 years later when WWII ignited, the impact of malaria on any war effort was not lost on the American government47, so in 1942 the predecessor of the CDC, The Office of Malaria Control in War Areas, was founded47. Its primary objective was to direct malaria eradication efforts, which officially commenced in 194747 but had technically begun years prior, seeking to reduce mosquito numbers by mass spraying of DTT as well as oiling to drown aquatic-stage larvae47. DTT was highly effective against mosquitoes and helped suppress their populations to near extinction, but its peripheral ecological effects later became known following the publication of the book Silent Spring48. Besides causing a significant ecological impact, DDT did not achieve

5 complete elimination of the target mosquito populations, which were able to rebound at the end of the intervention thereby restoring original local ecologies.

Mosquitoes are one of many small insect prey species, and are not known to be the sole pollinators of any plants, perhaps with the exception of one orchid species in North America49,50. It is generally agreed by the scientific community that if they were eliminated, their ecological niche could be easily filled by other insects51. However only recently have studies begun to better understand the mosquito’s role in its ecological habitat52. While no resounding ecological consequences were observed in Italy or the United

States after their near-elimination, it would be foolish to presume that all ecologies are identical.

Therefore suppression of mosquito vectors – even if only temporary – should be carried out only with the outmost care and with constant ecological monitoring53.

1.3 Mosquito genetic control – An introduction

Genetic control of vector-borne diseases through manipulation of the vector has long been considered a panacea for disease control, as genetic technologies have consistently proven effective in humanity’s fight against insects for the better part of the past century. Briefly, first proposed in the 1930s by E.F. Knipling54,

Sterile Insect Technique, a technique which suppresses wild populations following release of Sterile males

(SIT, discussed further in Section 1.4A) enabled the eradication of a major agricultural pest, the screwworm Cochliomyia hominivorax, from all of North and Central America55,56. Following the development of genetic sexing strains57 (discussed further in Section 1.10B), eradication of the

Mediterranean fruit from the United States and was achieved soon after58. More recently these advances were followed by the first releases of transgenic insects to suppress Dengue-spreading Aedes populations in Grand Cayman59 and Brazil60 (discussed in Section 1.4B). Now, with the advent of

CRISPR/Cas9 officially heralding the “dawn of the gene-editing age”61, breakthroughs in developing

6 genetic control strategies against important disease vectors have occurred at a break-neck pace, providing new lymph to the prospect of eradicating the first vector-borne disease.

Technologies aimed at controlling vector-borne diseases using genetic modification of the insect vectors can be broadly grouped into two general categories; those that seek to eradicate the vector population

(Figure 1.1A), and those that seek to replace them with populations incapable of spreading disease (Figure

1.1B). Termed suppression and replacement technologies respectively, they are not mutually exclusive and can be achieved using a variety of methodologies. These range from mass release of sterile males which sterilize the females with whom they mate58 (Figure 1.1A), to drive systems which – through biased inheritance – spread factors resulting in population crashes62,63, or which can replace pathogen-permissive populations with pathogen-refractory ones64 (Figure 1.1B). Each type of design has its own benefits and drawbacks ecologically, biologically, and logistically (Reviewed in1,65) with one of the major distinctions being whether their modifications are designed to disappear following release (self-limiting) or are designed to persist on a long-term basis (self-sustaining).

Figure 1.1 | Dynamics of [A] population suppression and [B] population replacement vector control strategies. [A] Population suppression techniques generally require iterative releases of sterilizing mosquitoes (green arrows) over a number of generations (X-axis), resulting in decreased population size. [B] Population replacement strategies require fewer releases (green arrow) which results in the original wild type population going to extinction (grey), and being replaced with the engineered strain (green) ideally with the engineered drive strain going to allelic fixation (Drive frequency). Reproduced from 1 with permission from Creative Commons 4.0. Also available in original text in Appendix C.

7

1.4 Population suppression systems

1.4 A. Sterile Insect Technique for mosquito population suppression.

Some of the first genetic insect control campaigns for insect population suppression used a technology termed Sterile Insect Technique (SIT). “Classic” SIT is the use of chemicals or radiation to sterilize males, which when released en masse, compete with wild males for females mates, and effectively sterilize the females following copulation66. This system is most efficient in monandrous species in which females mate once per lifetime, and in which failure to transfer viable sperm to the female renders her permanently infertile. Effectively acting as a species-specific chemical-free insecticide by their ability to suppress specific populations, such vector control programs require massive numbers of males be released iteratively until population elimination is achieved66. Importantly for mosquitoes in particular, male- exclusive releases are compulsory as females bite and spread disease.

SIT was first successfully implemented to eliminate a number of important agricultural pests, most notably the Screwworm55 and Medfly58, however its successful implementation in mosquitoes has proven elusive.

While chemo- or radiation-based sterilization of male Aedes albopictus67, An. arabiensis68,69, An. stephensi and An. gambiae70 can be achieved, it generally causes a reduction in male mating fitness71,72. Because mosquito males must compete in swarms for rare mates, such fitness defects necessitate that more males, than initially predicted by simple models, need be released to achieve suppression73 making such systems impractical. However, a notable hurdle for genetic SIT systems exists; in order to produce the substantial numbers required, the desired sterility phenotypes must either be suppressed during mass rearing or induced for release. Towards this end, alternative technologies for mass male sterilization have been designed and implemented for GM mosquito control.

8

1.4 B. RIDL-based transgenic population suppression systems.

In the face of these challenges, the Classical SIT system was updated for more effective use in mosquitoes.

In the absence of effective sterilization systems using classic methods, the company OXITEC developed a transgenic system in , which although distinct from SIT74, can achieve similar aims75,76.

Termed Release of Insects carrying Dominant Lethals (RIDLTM), a few permutations of this system were developed, all largely characterized by release of males carrying a tetracycline-controlled lethal transgene that kills their offspring. RIDL transgenic systems are suppressed by the addition of tetracycline during rearing, enabling mass quantities of mosquitoes to be produced, which are then activated following release resulting in offspring death. The absence of tetracycline in the wild larval environment permits aberrant accumulation of protein products from the transgene ultimately causing leathality75,76. Therefore in this system the males are fertile although their offspring are not viable, resulting in a population suppression effect similar to that of classic SIT1.

A similar system, fsRIDLTM, causes lethality specifically in females rather than all individuals, by disrupting their ability to fly. By expressing the aforementioned lethal transgene specifically in the female’s flight muscles, the daughters of released males still contribute to larval suppression by competition, but never fly away to bite, mate, or spread disease77. First developed in Ae. aegypti, but eventually transferred to

Ae. albopictus78 and An. stephensi79, this transgenic system has the added benefit that males are unaffected by the transgene – thus allowing them to mate and continue to sire offspring. This permits the population suppression effect to persist for a few generations following the initial release, making the initial investment more effective, until the fitness costs of the large transgenic cargo causes all transgenics to be outcompeted and die. In the end, both the RIDL and fsRIDL technologies rely on mass release of males whose offspring are ultimately inviable, and in which the transgene does not persist in the population long term, but does achieve transient population suppression.

9

With successful field trials suppressing Ae. aegypti populations by as much as 80-95%80, Oxitec recently announced its intent to develop similar technologies in the Central American malaria vector An. albimanus in conjunction with the Bill and Melinda Gates Foundation81. Despite having already developed the technology in the anopheline species An. stephensi79, they notably have never announced intention to develop their technology in any African malaria vectors – leaving a critical gap in the development of genetic sterilization technologies in these important species.

1.5 Population replacement systems

While eradication of mosquitoes may be alluring to some, it remains mostly unfeasible across many large geographic expanses of Africa, or where vector populations are particularly dense1,82,83. Since the creation of the first transgenic anopheline84, scientists have discovered a number of native genes or synthetic genetic cargoes that can render mosquitoes refractory to parasite transmission (reviewed in85,86, discussed further in Section 1.5E). But to affect transmission in the field, these traits would need to be spread into every wild mosquito. If we could develop a method for biasing inheritance of these genetic cargoes, they may be able to spread to most members of the population over many generations in spite of fitness costs. Expanding such technologies to other organisms beyond mosquitoes could have broader impacts, enabling us to address several other major world problems, including the rise of and herbicide resistance, and the agricultural and environmental damage wrought by invasive species2.

In nature, certain selfish genes ‘drive’ themselves through populations by increasing their inheritance probability87. Some well-characterized examples include endonuclease genes that copy themselves into homologous chromosomes88, segregation distorters that destroy competing chromosomes during meiotic germline development89, transposons that insert multiple copies of themselves throughout the genome90, and Medea elements that kill competing siblings who do not inherit them91,92. In this class we can also broadly place vertically-transmitted bacterial endosymbionts such as Wolbachia93 that bias their

10 inheritance by manipulating reproduction of their insect hosts. Such selfish genetic elements have been identified in species ranging from flies94,95, yeast96, and nematodes97; to corn98, beetles91, and lemmings99

– with many hundreds likely lurking in species yet to be studied87.

1.5 A. Endonuclease gene drives – Mode of action and history of development.

One such type of selfish gene spurred early scientific curiosity by the mode in which it spread, and has ultimately inspired a new field of genetic engineering. Naturally occurring homing endonuclease (HEG) gene drives are selfish genetic elements, which encode a large endonuclease within a specific locus, capable of cleaving the wild type homologous locus. When the cell repairs this break via the Homology

Directed Repair (HDR) DNA-repair pathway88, it is forced to use the HEG-encoding gene as the template for repair, inducing the cell to copy its sequence into the break. This copying process is termed ‘homing’, and it ultimately results in the heterozygous cell becoming homozygous. When this occurs in the germline of sexually reproducing species, the gene is inherited accordingly, resulting in super-Mendelian inheritance of the HEG in the progeny. Over many generations, this caused the HEG allele to increase in frequency, effectively ‘driving’ it through the population (Figure 1.1B, Figure 1.3A). Over time, the HEG drive – and any associated genetic cargo – spreads into all individuals in a population, enabling deliberate engineering of the entire wild populace83,87,100-102. This basic copy-and-paste mode of action provides the basis for many modern synthetic gene drives, including the gene drives outlined in this dissertation.

In 2003, Austin Burt first suggested that naturally-occurring meganucleases, such as OnuI, I-PpoI or I-SceI, could be modified to recognize new host sequences in the insect genome to enable manipulation of natural insect populations103. Several early efforts focused on building such HEG-enabled gene drives in mosquitoes with the aim to block malaria transmission104-109, with the first successful demonstration of the drive phenomenon in mosquitoes occurring in 2011107. Concurrently, efforts were underway to redesign HEGs to target the ribosomal rDNA repeats on the Anopheles X chromosome106, whose

11 expression from the male Y-chromosome110 would create a Selfish-Y system111 capable of ‘shredding’ X- containing sperm during spermatogenesis, ensuring male-only progeny112. This type of ‘genetic load’ drive100,111-114 would spread rapidly but the resulting all-male population would inevitably crash100,115, the speed of which depending on the dynamics of gene flow116. Among a variety of technical challenges, development was ultimately thwarted by difficulties in engineering the HEGs to target new host sequences117-119, making implementation impractical if not impossible.

1.5 B. TALEN and ZFN-mediated gene drives.

At this time, newer designer nucleases arose enabling more straightforward targeting of genomic sequences by designer DNA-binding modules. These zinc finger (ZF), or transcription activator-like effector

(TALE) DNA-binding proteins, are composed of multiple large DNA binding domains fused in sequence, with each domain characterized by slight amino acid differences enabling binding to distinct DNA bases.

Assembling multiple DNA-binding domains into repetitive modules allows for binding up to 20 consecutive bases, and fusing a pair to the homodimerizing endonucleolytic domains of the type II endonuclease,

Fok1120, enables targeted DNA cleavage of unique genomic sites (Figure 1.2A,B). This technology was successfully used to generate the first knockout and knock-in mosquito mutants121-125.

During the course of my Master’s Thesis, from 2011-2012, research I conducted at the University of

Strasbourg, France, successfully developed the use of paired transgenic TALE endonucleases (TALENs) to make the first knockout anopheline121 (Full text in Appendix B). However, attempts to build a basic gene drive targeting a neutral genomic site126 using the same transgenic TALEN technology resulted in shuffling of the repetitive TALE DNA-binding domain sequences, rendering the transgene nonfunctional and precluding the possibility of drive (Smidler, Marois et al. unpublished). Similar TALEN- and ZFN-enabled gene drives published in Drosophila suffered from the same instability due to their repetitive nature127.

12

Taken together, these studies revealed that highly repetitive gene drives would not likely be tolerated in biological systems, necessitating development of less repetitive endonucleases for gene drives.

1.5 C. CRISPR-mediated gene drives.

Following the discovery that it could be used as a powerful engineering tool128,129, the Clustered Regularly

Interspaced Short Palindromic Repeat/Cas9 system (CRISPR/Cas9) has completely revolutionized synthetic biology. Along with many other groups, we recognized its promise for use in gene drives2,53 following the first application in cell systems130-133. Based on the Type II CRISPR acquired immune systems which enables bacteria to ‘remember’ viral and recognize and cleave the DNA of new invaders134,135, CRISPR has been adopted to edit the genomes of a number of species (Reviewed in136). The

Cas9 endonuclease can be targeted to nearly any DNA sequence in the genome through the action of a synthetic small guide RNA (gRNA), with the only DNA sequence requirement being the bases ‘NGG’

(termed a PAM) at the 3’end. The gRNA is less than 100 bp in length, and is composed of the first 17-20 bp ‘spacer’ sequence which directs Cas9 to the ‘target’ sequence through direct Watson-Crick-Franklin basepairing, followed by a sequence which forms a stem-loop secondary structure recognized by Cas9128.

The Cas9 enzyme and small guide RNA work together as a two-part system to bind and cleave DNA, and are capable of targeting every 8 bp on average128 (Figure 1.2C). Moreover, because unlike TALENs and

ZFNs targeting different sequences does not require any protein redesign – instead only requiring redesign of a small gRNA – this system is much more inherently flexible, enabling straightforward redesign and simultaneous targeting of multiple sequences with less sequence repetitiveness. Because CRISPR- mediated genome editing relies on the same copy-and-paste mechanism as prior endonuclease gene drives, we reasoned it could be capable of enabling powerful gene drives if properly engineered2,137.

13

Figure 1.2 | A basic diagram of ZFN, TALEN, and CRISPR/Cas9 DNA cleavage. [A] A representation of a ZFN binding and cleaving DNA. ZFN repeats (blue) bind 3 bp each, when fused together are capable of binding unique genomic sequences (two ZFNs capable of binding 12 bp each shown). Fusion to a FokI domain (orange) mediates DNA cleavage when heterodimerized, necessitating use of two adjacent ZFNs for genome targeting. [B] A representation of TALENs binding and cleaving DNA. TALE-DNA binding repeats (purple) each bind a single DNA base, fusion of multiple together in sequence enables binding of unique genomic sequences (two TALENs capable of binding 12 bp each shown). Fusion to a FokI domain (orange) mediates DNA cleavage when heterodimerized, necessitating use of two adjacent TALENs for genome targeting. [C] The S. pyogenes Cas9 endonuclease (Cas9, salmon) in complex with a synthetic guide RNA (gRNA, red) can be programmed to target specific DNA sequences for cleavage (Genome, grey). The spacer sequence on the gRNA (blue) can guide the CRISPR complex to a target site (yellow) in the genome through direct Watson-Crick-Franklin basepairing. Wild type Cas9 endonuclease can crease dsDNA breaks (scissors) which can be repaired by endogenous cellular DNA repair pathways. The only sequence requirement is the presence of a protospacer adjacent motif (PAM, green) which in the case of canonical Cas9 is the bases ‘NGG’. This figure was created with Biorender.com.

1.5 D. Safety considerations concerning gene drive

Due to this ease of design the general flexibility of CRISPR to engineer innumerable species, CRISPR gene drives could pose a threat as a potential dual-use technology in some species. With this in mind we were motivated to discuss safety regulations for monitoring and development even prior to beginning laboratory experiments on the topic53. However, in keeping with the molecular biological focus of this dissertation, this work will not be discussed further.

14

1.5 E. Anti-parasitic cargoes for population replacement.

The population-replacement gene drives outlined in Chapter 3 have the ultimate aim to spread anti- malarial genetic cargoes into wild populations. A number of potential mechanisms capable of rendering mosquitoes genetically refractory to malaria infection have been developed, including activation or augmentation of the mosquito host immunity138-146, impairment of parasite development by disrupting factors necessary for parasite invasion of the midgut or salivary glands147,148, and introduction of artificial constructs with parasite-killing properties149-151.

Mosquitoes mount a multifaceted immune response to invading parasites and viruses (reviewed in 138,139), which can be artificially manipulated to reduce infection or transmission of the parasite. The NF-κB-like immune pathway in mosquitoes has proven to be a good target for upregulating the native immune response. As an example, overexpression of the NF-κB transcription factor Rel2 causes a significant decrease in parasite prevalence in engineered Ae. aegypti and An. stephensi140,141, and upregulation of

Rel2’s downstream effector genes also causes a reduction in parasite load142,143. While promising, no transgenic system relying solely on native immune activation has been successful at completely ablating parasite development, necessitating the identification of additional anti-malarial cargoes for population replacement.

Parasite killing can also be achieved by transgenic expression of parasite-targeting single-chain antibodies borrowed from mammals152, which when expressed in the midgut153, salivary glands154, or both tissues155, can significantly reduce parasite loads. While originally developed in An. stephensi, these combined transgenes may achieve similar inhibition of parasite development in An. gambiae and therefore represent good cargo candidates for population replacement gene drives.

15

Disruption of host-pathogen interactions required for parasite invasion and replication also represents a promising strategy. For example, transgenic expression of the small synthetic peptide SM1 that binds to both midgut and salivary gland epithelia has been shown to block putative receptors required for invasion of the rodent malaria parasite P. berghei, resulting in reduced parasite loads156. However this strategy was ineffective against the human malaria parasite P. falciparum infections157. In another example, knockout of the enzyme kynurenine 3-monooxygenase (kmo) resulted in a decrease in Xanthuric Acid (XA) production in the midgut148 – a waste product of tryptophan metabolism, and a known Plasmodium exflagellation stimulator147. Such mutants displayed significant decreases in oocyst intensity, suggesting that if the phenotype can be recapitulated in infected An. gambiae, knockout during the course of drive spread by the addition of drive-encoded kmo-targeting gRNAs could be a viable candidate anti-malarial cargo. Despite these promising advances, no anti-malarial gene described to date achieves complete suppression of the deadliest malaria parasite, P. falciparum, necessitating further development for population replacement strategies.

1.6 The problem of drive-resistant alleles (DRAs)

Gene drives only copy and spread when the CRISPR-generated DNA breaks are repaired by the HDR pathway. HDR rates are known to vary across cell types133, developmental stages158,159, species160, and phase of the cell cycle161. Use of any one of the alternative repair pathways, such as the error-prone non- homologous end-joining (NHEJ) or microhomology-mediated end joining (MMEJ), will not only result in failure to copy the drive, but may also prevent future drive spread via the generation of mutations within the target sequences recognized by the gRNA. Once one such mutation arises, or the drive encounters a naturally occurring one, the target sequence can no longer be recognized by the gRNA and cannot be cleaved by Cas9. The inability to cleave and home renders the drive no more than a large burdensome transgene. If the fitness burden of these uncleavable alleles is negligible, mutations will increase in

16 frequency and will out compete the drive, forcing it to extinction. Such alleles thus earn the name ‘drive- resistant alleles’ (shortened to ‘DRAs’ for brevity) (Figure 1.3B). For clarity, and in keeping with nomenclature standards of the field162, DRAs which maintain the wild type phenotype will be termed

‘r1DRAs’, while those that disrupt gene function will be termed ‘r2DRAs’ within this text.

Figure 1.3 | Biased inheritance of Gene drives through mosquito populations and the genetic outcomes of cleavage and homing. [A] During perfect drive, mosquitoes heterozygous for the gene drive (blue) always pass on the transgene to the progeny even when mating to wild type (grey). Over many generations, the biased inheritance of the drive results in the drive increasing in frequency, until all individuals in the population have it. [B] Individuals encoding the gene drive will initiate DNA cleavage of the corresponding wild type chromosome. This DNA break can repair by a number of endogenous cellular repair pathways, which can result in multiple distinct outcomes. The DNA break can repair back to the original wild type sequence, in which case cleavage can re-occur iteratively until one of the other two outcomes have occurred. The second outcome results in mutation at the site of the break, often the result of an End-Joining repair pathway. These mutations can inhibit subsequent cleavage and homing and are termed Drive Resistant alleles (DRAs). The final repair outcome is that of successful drive homing into the break, which occurs as a result of repair by the Homology Directed Repair (HDR) pathway. This homing has resulted in the duplication of the drive sequence from the original chromosome into the homolog, effectively making a heterozygous cell homozygous for the drive. If occurring in the germline, this causes biased inheritance of the drive. Reproduced from2 under Creative Commons 4.0. Also available in original text in Appendix D.

17

Modeling predicts that the key factor to drive success is its ability to suppress DRA formation, with homing efficiency, fitness burden, and introduction frequency only minimally contributing to drive success163. The generation of innovative drive architectures that are evolutionarily stable164 and capable of overcoming this issue is therefore a priority for the successful implementation of gene drives, as discussed in one of our studies (Appendix D2) and the topic of a Patent Application137.

1.7 Learning from other CRISPR gene drives

In 2015 the first demonstrations of CRISPR-mediated gene drives in mosquitoes were published for population suppression and population replacement in An. gambiae and An. stephensi respectively62,165.

These gene drives were characterized by use of a single gRNA targeting a sequence which could tolerate some degree of mutation. While these papers demonstrated successful drive in many individual couplings, they both failed to demonstrate invasion of the drive into cage populations presumably due to inhibition by DRAs. The replacement drive in An. stephensi addressed this in depth by demonstrating significant DRA formation (>77%) in the progeny of females64, likely due to leaky maternal deposition of Cas9 – expressed by the germline Vasa2 promoter166 – into developing oocytes at a stage in which homing cannot occur.

The population suppression drive in An. gambiae addressed this phenomenon in later publications by characterizing the generation of DRAs in the gRNA binding site, and demonstrated that they not only prevented drives from reaching fixation but also competed them to extinction167. These examples underscore the importance of designing drives with architectures that i) prevent the emergence of DRAs within the population, and ii) can ‘delete’ DRAs were they to arise.

Similar drives developed concurrently in Drosophila also demonstrated the emergence of drive-resistant mutant alleles, which could be slightly attenuated by using the less leaky nanos germline promoter for

Cas9 expression168. Modified population suppression gene drive designs soon followed in An. gambiae which also attenuated DRA formation of earlier drive designs62 by use of the more tightly regulated

18 germline promoters nanos and zpg, however achieving only slight decreases in the rate of DRA-based drive inhibition169. These were some of the first demonstrations that fine-tuning Cas9 expression to specific cell types and stages could have a notable impact on not just homing efficiency, but also could contribute to minimizing DRA formation.

In our opinion, all drive designs relying on a single gRNA for cleavage should not be considered truly

‘evolutionarily stable’170,171, as even those targeting the most conserved loci run the serious risk of developing DRAs, thwarting drive spread. The best way to overcome the issue of previously existing or recently evolved DRAs is to target multiple sites172. Because mutations in target sites are evolutionarily favored only when they are beneficial in providing resistance to the gene drive, using many target sites would reduce the chances of mutations emerging at all of the sites, as long as cutting rates are high103. In fact, numerous groups have modelled the rate of DRA formation and the number of gRNAs required to achieve fixation in spite of this with ~4 gRNAs predicted to be sufficient for most systems172. However, drives relying on multiple gRNAs have specific design requirements to prevent cleavage of the drive- containing chromosome, which necessitates the target sequence be removed or changed during homing to prevent self-cleavage within the drive chromosome. A multi-gRNA design was recently demonstrated in Drosophila in which four gRNAs targeted a gene for female fertility to induce population suppression173.

While this drive demonstrated super-Mendelian inheritance, it was highly ineffective at homing despite nearly 100% cleavage rate in the majority of individuals. It was also plagued by the creation of DRAs, many of which corresponded to aberrant sequence shuffling over repetitive regions of the drive, or by incomplete or aborted homing. This drive may have experienced instability due to its design. Two of the four gRNAs targeted internal sequences far from the homology arms, while the remaining two targeted regions adjacent to the homology boundary. With rates of HDR-mediated repair being highly dependent on the homology’s proximity to the cut site174, we suspect that only the two outermost gRNAs generated cleavage events which enabled HDR repair. Further, the authors reported that one of the outermost

19 gRNAs was a disproportionately prolific cleaver while the other gRNA near the homology boundary was not, suggesting this drive may have struggled to complete HDR repair at this end. Taken together these suggest that gRNA localization proximal to the homology boundary may be critical to attenuate instability.

Another population suppression gene drive in An. gambiae attempted evolutionarily stability by targeting highly conserved DNA sequences intolerant of SNPs due to their position over a critical splice site63. This drive was designed to interrupt splicing of the female-specific isoform of doublesex, a gene required for insect sex-determination. Either disruptive mutations or successful homing prevented female-specific splicing, resulting in an intersex phenotype and complete female sterility, while leaving males unaffected.

This drive homed efficiently via a single gRNA causing extinction of cage populations in only 7-11 generations, without succumbing to competition from DRAs. However, the drive was released at what we consider an unrealistic allele frequency of 25%, thus not permitting the population sufficient time to evolve DRAs as would be expected at lower release frequencies. Furthermore, the site targeted by the single gRNA is polymorphic in wild populations, and while the authors demonstrated that such a polymorphism can be cleaved by their gRNA in vitro, the in vivo cutting efficiency over this polymorphism may not be sufficient to prevent resistance. Importantly the existence of a naturally occurring polymorphism demonstrates that this site can in fact tolerate some variation, suggesting that r1DRAs will eventually emerge. Therefore although this type of design architecture is promising, it is important to note it is not wholly ‘resistance proof’63, and moreover it could be only utilized for population suppression campaigns, and not replacement, as it relies on knocking out an essential gene.

1.8 Fitness considerations for gene drives

Gene drives are large, nuclease-encoding, genetically burdensome transgenes whose fitness costs on the individual and population are a significant and critical factor to consider when designing potential drives for release. Two examples of these potential costs are provided below.

20

1.8 A. CRISPR off-target mutagenesis.

Despite its promise and flexibility, the CRISPR system has a well-documented tendency to cleave off-target sequences175, sometimes targeting sites with as many as five basepair mismatches between the target and gRNA131,174,176-178. The unintended DNA damage may be even greater than that reported so far, as commonly used methods to identify off-target mutations were recently shown to miss perturbations such as large chromosome rearrangements, deletions, and insertions 179. CRISPR paired-nickase strategies have been developed that have more stringent sequence requirements and can reduce off-target effects by 50-

1,500-fold180. However because CRISPR-based gene editing tools are still incipient, no one has yet tested this strategy for gene drive.

Fitness effects due to off-target mutagenesis are likely to become more pronounced over time as drives spread longer, farther, and faster. A gene drive expressing CRISPR for many generations would be expected to accumulate many off-target mutations over time, causing increasing fitness defects in the drive-containing genetic background until it homes into a new genetic background, potentially inhibiting drive spread over the long term, a side-effect ignored during modeling. Moreover, Cas9-expressing transgenic lines in many species have been hinted to experience fitness defects. For instance some Cas9- expressing Drosophila lines have notable fitness costs (personal communication Dr. Kevin Esvelt, MIT), and attempts to generate An. stephensi lines expressing Cas9 ubiquitously resulted in early larval lethality

(unpublished, personal communication Dr. Anthony James, UC Irvine). Fitness costs in An. stephensi were contained only when Cas9 expression was confined to the germline, following our suggestion, resulting in the generation of the first successful population replacement gene drive in mosquitoes165.

Interestingly, work from our lab suggests that Cas9-expression may not only affect mosquito host fitness, but also transmission of malaria parasite. Analysis of a germline Cas9-expressing transgenic line developed in the course of this work (further detailed in Section 2.1) revealed that females expressing the nuclease

21 in the absence of any gRNA supported significantly higher levels of P. falciparum infections. This effect was not mitigated by RNAi suppression of Cas9 or by outcrossing, suggesting Cas9 had caused a permanent alteration in the genome (unpublished, Dr. W. Robert Shaw). While the exact mechanism behind this phenotype is still unknown, these results are a stark warning that Cas9 activity in the absence of gRNAs may introduce unwanted mutations in the mosquito genome, as previously demonstrated by

Sundaresan et al. in vitro181 , with possible negative consequences for disease transmission.

1.8 B. Gene drive-imposed evolutionary selection pressures.

Not all gene drive designs exert the same selective pressures on their mosquito host. Theoretically a drive designed to efficiently avoid the emergence of DRAs should be self-optimizing through purifying selection182,183. Under purifying selection, broken or suboptimal drives within a population will be counterselected, while those most efficient will spread prolifically. Gene drives are transgenes which are subject to at least the same rate of baseline random cellular mutagenesis as other genomic sequences.

Therefore Cas9- and gRNA-encoding sequences are expected to eventually mutate, rendering gene drives simply non-driving transgenes. However, even if such mutations occur in some individuals, as long as others still harbor a functioning drive, then drives should continue to spread if DRAs are absent from the population184. An entire field of study is dedicated to exploring this phenomenon in naturally occurring

HEG selfish genetic elements. For these reasons, such HEGs have been described as “molecular parasites, which have to be regarded as distinct evolutionary units, and whose fate is intertwined, but separate, from the fate of the [] host organism”184. Similar to what would be expected of well-designed population- replacement CRISPR gene drives, naturally occurring HEGs undergo cycles of self-optimization until fixation is achieved, followed by mutation, then counter-selection, near extinction, and then reinvasion184.

Such self-optimizing properties have enabled some HEGS to drive within their hosts for over 500 million years185, has allowed them to hop into distantly related species186-188, and has made them hop back-and-

22 forth between closely related organisms as evidenced by phylogeny discordance between many HEGs and their hosts189-193. This interspecies transfer and persistence of HEGs underscores the natural ability of such selfish genetic elements to self-optimize, and the potential for a well-designed CRISPR gene drive to achieve the same.

Naturally occurring HEGs can however only achieve fixation if they impose minimal fitness costs on their host184. It is therefore expected that any CRISPR gene drives with significant effects on host fitness may struggle to spread. For example, CRISPR gene drives which cause infertility or population suppression62,170 exert extremely high selection pressure on their mosquito hosts because survival of the species depends on emergence of resistance mechanisms. Similarly, population replacement gene drives seeking to spread infection-inhibiting factors may impose selective pressures depending on the phenotypes they cause. For example, while a cargo ablating the mosquito salivary glands might not be able to transmit pathogens194, significant selection pressure to maintain critical organ function could thwart its spread. Therefore, less disruptive cargoes, such as those which simply enhance the mosquito’s ability to mount an immune response against the parasite (Section 1.5E), may be preferred to achieve robust population replacement.

Intriguingly, synthetic CRISPR-mediated population replacement gene drives could also further self- optimize beyond their original design. It is possible that such drives may, for example, select for polymorphisms in Cas9 regulatory regions to increase or fine-tune expression to ideal cell types and developmental stages, or may select for mutations in Cas9 itself to reduce off-target mutagenesis. Such drives may select for altered gRNA sequences to enable drive into polymorphic sequences, or even duplicate gRNAs to enable faster drive. After many generations a drive could therefore have a very different genetic make-up relative to its original design, posing the important question of possible associated risks.

23

It is important to note that even the most robust CRISPR-mediated population-replacement gene drives are likely to succumb to a variety of other inhibition mechanisms evolved by eukaryotes to combat selfish genetic elements. Selfish genetic elements have evolved innumerable times throughout history, and one type, transposons, account for nearly one-third to one-half of the Drosophila and human genomes respectively195. Many organisms have evolved robust systems to defend against transposable elements196, because remobilization results in genomic instability due to mutation at excision and insertion sites195.

These selfish genetic elements all share the feature of generating long-noncoding RNAs with significant secondary structure, capable of being recognized by the cell as ‘non-self’197-199. Generally recognized by formation of hairpins in nascent transcripts, the genetic elements which encode them ultimately become the target of gene silencing200,201 by the PIWI pathway, first post-transcriptionally, and then followed by more permanent heterochromatinization of the encoding locus. But there is further evidence that the selfish-genetic element burden is so high in insects, that some species have evolved additional RNAi-like silencing202 pathways to deal with the burden imposed by transposons. Perhaps unsurprisingly then, attempts to create dsRNA-expressing transgenic systems in Aedes mosquitoes resulted in the loss of RNAi- silencing from the transgene after 17 successive generations while leaving the tightly-associated fluorescence reporter gene unaffected203. And subsequent work by the same group to generate a completely new set of dsRNA-expressing transgenic lines experienced the same fate204. These observations suggest that mosquitoes retain powerful cellular mechanisms to suppress selfish genetic elements through the recognition and selective silencing of transgenic stem-loop RNAs. How does this affect CRISPR-enabled gene drives? Perhaps the power of such gene drives – the ability to target multiple sites by gRNAs – could also be their downfall. Although slightly different structurally, the gRNAs encoded by CRISPR gene drives are distinctly similar to long noncoding RNAs (lncRNAs); small transposon- associated RNA species205, with no protein-encoding counterparts in the genome, and with distinct stem- loop secondary structures. Similarly, gRNAs could be targeted for inhibition by similar cellular mechanisms

24 used to silence the noncoding RNAs of transposons due to their similar structure and properties. Although the current focus of the field is to suppress DRAs to enable drive spread, it is possible that once this issue is overcome gene drive scientists may also face the hurdle of overcoming gRNA silencing.

1.9 Designing evolutionarily stable gene drives for population replacement

As mentioned before, for a gene drive to be evolutionarily stable, DRAs must be prevented and if they do arise a mechanism must be designed to delete them. This could be achieved using a drive architecture which is inserted within – but does not disrupt – the C-terminus of a haploinsufficient gene (also termed haplolethal), i.e. a gene in which loss-of-function in a single diploid allele results in cell lethality. During drive homing, most mutations would result in cell death, precluding the spread of DRAs and ensuring that correct drive homing is the only outcome which restores gene functionality and permits cell survival. In contrast, designs targeting exclusively haplosufficient genes would generate many r2DRAs, greatly inhibiting drive spread in {r2DRA/drive} heterozygotes by their persistence.

Haplolethal genes are rare within eukaryotic genomes, and nearly all function in ribosome protein synthesis206, with significant conservation from yeast to humans207. Functionally, their haploinsufficiency is due to the average cell’s requirement for an astonishing number of ribosome units, with individual mammalian cells requiring as many as 10 million ribosomes each208. Despite being caused by different mutations in distinct subunits, all mutants share similar growth defects whose phenotypes are practically identical209. Although haplolethal genes are difficult to study due to their inability to tolerate genetic perturbation, a number of screens in D. melanogaster have identified ribosomal genes by disruption of their regulatory sequences209. Such mutants are termed Minutes for their diminished size and growth defects; with nearly all characterized by prolonged development, female-infertility and low male fertility, small body size, abnormally short bristles on the body, and failure to eclose210. Minute phenotypes fall on a dose-dependent spectrum with more intense phenotypes corresponding to greater loss of ribosome

25 protein209, but significant phenotypes can be observed with as little as a ~10% decrease in mRNA levels211.

To date the only known Minute demonstrated in mosquitoes has been an ablation of ovarian development following transient ribosome depletion212, and most ribosomal genes in An. gambiae have over 80% protein identity to D. melanogaster despite being separated by >200 million years of evolution213, supporting functional conservation of this critical cellular process. Therefore, while on the one hand gene drives designed within ribosomal genes would be extremely challenging to engineer, leaving no room for error, on the other hand success would provide perhaps the strongest selection possible against the emergence of DRAs.

1.10 Mosquito genetic control strategies need more tools

Field releases of GM mosquitoes have already been carried out in the Cayman Islands, and since 200960,80, and releases of population suppressing gene drive mosquitoes are on the horizon. With

GM mosquito releases in Africa becoming increasingly likely, the lack of effective tools to monitor and assess the effectiveness of field these trials has become apparent. Regardless of whether SIT-like trials

(Section 1.4A) or gene drive releases (Section 1.5) are being considered, all strategies require the exclusive release of males because females bite and spread disease. The fitness and competitiveness of these males in a wild mating setting is critical to trial success and is a very important variable to monitor. This is particularly relevant in mosquito species that, such as the most important malaria vectors An. gambiae,

An. coluzzii and An. funestus, mate in large swarms where competition between males is fierce214.

However there are currently no straightforward methods to determine the mating fitness of released males without sacrificing the female. Currently available methods necessitate dissections of females shortly after copulation to assess the presence of sperm in the spermatheca and the mating plug in the atrium. The only non-invasive assessment of male mating success available at present is to blood feed mated females, allow them to oviposit, and screen their progeny for specific genetic traits to determine

26 paternity, thereby making analysis of mating events laborious. Novel methods to study the fitness and mating competitiveness of these released males are needed if GM field releases are to be successfully monitored. Additionally, genetic strategies for vector control rely on the release of male-only populations, because female mosquitoes bite and spread disease. This necessitates the generation of genetic sexing technologies to separate males from females prior to a release.

Mechanical sorting methods are possible in Aedes and species due to sexual size dimorphism215, with robots recently being developed to enable this on a large scale216-218. However this technology is not readily transferable to anophelines due to close similarities between the sexes, necessitating development of genetic technologies to aid in the sorting process. Transgenic lines have been developed which differentially express fluorescent markers in the sexes110,219-221, enabling fluorescence-based sorting to remove females. However these systems suffer from multiple problems including the fact that half of the population is discarded (females) and sorting must occur at the larval stage110,219 precluding the option of shipping pre-sorted embryos. Systems that selectively kill females are therefore more desirable. In an effort to make males incapable of siring daughters, transgenic lines have been developed which encode

HEGs capable of selectively destroying X chromosome-containing sperm106 (discussed previously in

Section 1.5A). While early versions of this system were plagued by infertility caused by sperm-mediated deposition of the HEG into zygotes destroying the maternal X chromosome106, subsequent optimization has led to systems capable of generating ~95% male-only populations with minimal fertility costs109,222.

Although significantly improved, this system is still not sufficiently stringent for most control programs, necessitating development of sexing strategies achieving nearly 100% separation. While the work outlined in this dissertation does not attempt to develop genetic sex-separation technologies, it does set out to develop innovative transgenic systems for mosquito vector control, and in doing so, has kept a constant focus on ensuring compatibility and complementarity to many of the sex-separation technologies currently under development.

27

1.11 Summary of Aims

The work described in this dissertation aims at developing novel tools to aid genetic control programs in

An. gambiae towards malaria elimination.

Over the course of this work I have aimed to develop both self-limiting population suppression, and self- sustaining population-replacement systems,– as well as a suite of supporting genetic tools to enable better field trials and fundamental studies into basic biology of the deadly primary African malaria vector,

Anopheles gambiae. Later chapters outline my studies describing work aimed at developing CRISPR-based systems to genetically sterilize males for vector control programs (Chapter 2), generating synthetic selfish genetic elements for spreading disease-refractory cargoes (Chapter 3), and validating a transgenic system for monitoring effectiveness of field trials based on the release of genetically modified (GM) mosquitoes

(Chapter 4).

In Chapter 2 we outline the development of CRISPR technology in An. gambiae to achieve ablation of the adult male germline, resulting in spermless males for use in SIT vector control programs. The genetic system we generate facilitates generation of large quantities of sterile males in mass, and results in 95% male infertility. We explore the causes of residual male fertility and assess that failure of CRISPR to mutate the target site as well as the induction of fertility-maintaining mutations both result in maintenance of male spermatogenesis in a minority of males. Following optimization, this novel genetic system could be adapted for use in SIT in An. gambiae.

In Chapter 3 we make progress toward the generation of CRISPR-enabled, evolutionarily stable gene drives for population replacement, and develop a suite of novel genetic tools to facilitate their creation and testing. We develop a novel method – based on the cellular phenomenon of Interlocus Gene

Conversion – for gene knock-in of complex templates into genetically intractable loci to create gene drive

28 docking lines. We demonstrate successful survival of individuals carrying recoding of an endogenous haplolethal ribosome gene, and characterize them as Minute ribosome mutants. In attempts to create a functioning drive, we demonstrate that strict regulation of Cas9 expression is critically important for designs targeting haplolethal genes, and identify areas requiring significant optimization for successful drive. Our work also suggests that gRNA-expressing transgenes may be epigenetically silenced over time, leading to loss of function, a finding with potential impact for future gene drive campaigns. Combined, our results suggest that drive designs targeting haploinsufficient essential genes in the primordial germ line may be fundamentally inviable.

Finally, in Chapter 4 we develop a novel transgenic tool to enable the non-invasive assessment of male mating competitiveness in GM field trials. We characterize a Male-Accessory Gland-specific promoter, and harness it to create a transgenic line in which males transfer a fluorescent mating plug to females. We demonstrate that these males have mating competitiveness comparable to wild type individuals and show no additional fitness defects, proving that this tool could be included in GM mosquito trials to facilitate the identification of successful mating events and determine the mating competitiveness of released males.

29

1.12 References 1 Gabrieli, P., Smidler, A. & Catteruccia, F. Engineering the control of mosquito-borne infectious diseases. Genome biology 15, 535, doi:10.1186/s13059-014-0535-7 (2014).

2 Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife, e03401, doi:10.7554/eLife.03401 (2014).

3 Whitfield, J. Portrait of a serial killer. Nature (2002).

4 Goldberg, A. Plagues [TV Documentary]. PBS(1988).

5 Kwiatkowski, D. P. How malaria has affected the human genome and what human genetics can teach us about malaria. Am J Hum Genet 77, 171-192, doi:10.1086/432519 (2005).

6 Organization, W. H. Vector-borne diseases, (2017).

7 WHO. World malaria report 2017. Geneva, Switzerland: World Health Organization 8 Bloom, D., Bloom, L. R. & Weston, M. in Geneva: World Economic Forum Global Health Initiative.

9 Foundation", B. a. M. G. What We Do; Malaria Stategy Overview,

10 WHO Malaria fact sheet #94, (December 2011).

11 Organization, W. H. World Malaria Report 2016 (World Health Organization, Geneva). (2016).

12 Bhatt, S. et al. The effect of malaria control on Plasmodium falciparum in Africa between 2000 and 2015. Nature 526, 207-211, doi:10.1038/nature15535 (2015).

13 Shaw, W. R. & Catteruccia, F. Vector biology meets disease control: using basic research to fight vector-borne diseases. Nature Microbiology 4, 20-34, doi:10.1038/s41564-018-0214-7 (2019).

14 Organization, W. H. Global report on insecticide resistance in malaria vectors: 2010–2016. (2018).

15 Lengeler, C. Insecticide treated bednets and curtains for malaria control. Cochrane database of systematic reviews (1998).

16 Who. World malaria report 2013. (World Health Organization, 2014).

17 N’Guessan, R., Corbel, V., Akogbéto, M. & Rowland, M. Reduced efficacy of insecticide-treated nets and indoor residual spraying for malaria control in pyrethroid resistance area, Benin. Emerging infectious diseases 13, 199 (2007).

18 Temu, E. A. et al. Pyrethroid resistance in Anopheles gambiae, in Bomi County, Liberia, compromises malaria vector control. PloS one 7, e44986 (2012).

30

19 Mulamba, C. et al. Widespread pyrethroid and DDT resistance in the major malaria vector Anopheles funestus in is driven by metabolic resistance mechanisms. PloS one 9, e110058 (2014).

20 Toé, K. H. et al. Increased pyrethroid resistance in malaria vectors and decreased bed net effectiveness, Burkina Faso. Emerging infectious diseases 20, 1691 (2014).

21 Mnzava, A. P. et al. Implementation of the global plan for insecticide resistance management in malaria vectors: progress, challenges and the way forward. Malaria journal 14, 173 (2015).

22 Ranson, H. Current and Future Prospects for Preventing Malaria Transmission via the Use of Insecticides. Cold Spring Harbor perspectives in medicine 7, doi:10.1101/cshperspect.a026823 (2017).

23 Bayoh, M. N. et al. Persistently high estimates of late night, indoor exposure to malaria vectors despite high coverage of insecticide treated nets. Parasites & vectors 7, 380 (2014).

24 Russell, T. L., Beebe, N. W., Cooper, R. D., Lobo, N. F. & Burkot, T. R. Successful malaria elimination strategies require interventions that target changing vector behaviours. Malaria journal 12, 56 (2013).

25 Glunt, K. D. et al. Long-lasting insecticidal nets no longer effectively kill the highly resistant Anopheles funestus of southern Mozambique. Malaria journal 14, 298 (2015).

26 Rehwagen, C. WHO recommends DDT to control malaria. Bmj 333, 622 (2006).

27 Ranson, H. & Lissenden, N. Insecticide Resistance in African Anopheles Mosquitoes: A Worsening Situation that Needs Urgent Action to Maintain Malaria Control. Trends Parasitol 32, 187-196, doi:10.1016/j.pt.2015.11.010 (2016).

28 Knox, T. B. et al. An online tool for mapping insecticide resistance in major Anopheles vectors of human malaria parasites and review of resistance status for the Afrotropical region. Parasites & Vectors 7, 76, doi:10.1186/1756-3305-7-76 (2014).

29 Edi, C. V., Koudou, B. G., Jones, C. M., Weetman, D. & Ranson, H. Multiple-insecticide resistance in Anopheles gambiae mosquitoes, Southern Cote d'Ivoire. Emerging infectious diseases 18, 1508-1511, doi:10.3201/eid1809.120262 (2012).

30 Cisse, M. B. et al. Characterizing the insecticide resistance of Anopheles gambiae in Mali. Malaria journal 14, 327, doi:10.1186/s12936-015-0847-4 (2015).

31 Mitchell, S. N. et al. Identification and validation of a gene causing cross-resistance between insecticide classes in Anopheles gambiae from Ghana. Proceedings of the National Academy of Sciences 109, 6147-6152 (2012).

32 Lobo, N. F., Achee, N. L., Greico, J. & Collins, F. H. Modern Vector Control. Biology in the Era of Eradication, 69 (2017).

31

33 Bruce-Chwatt, L. J. & De Zulueta, J. The rise and fall of malaria in Europe: a historico- epidemiological study. (Published for the Regional Office for Europe of the World Health …, 1980).

34 Shah, S. The fever: How malaria has ruled humankind for 500,000 years. (Sarah Crichton Books, 2010).

35 Perosa, A., Murray, P. & Murray, P. Febris: a poetic myth created by Poliziano. Journal of the Warburg and Courtauld Institutes, 74-95 (1946).

36 Majori, G. Short history of malaria and its eradication in Italy with short notes on the fight against the infection in the mediterranean basin. Mediterranean journal of hematology and infectious diseases 4, e2012016-e2012016, doi:10.4084/MJHID.2012.016 (2012).

37 Sepulcri, P. La malaria nel Veneto: storia, epidemiologia, l'opera dell'Istituto antimalarico. (Fantoni, 1963).

38 Humphreys, M. Water won't run uphill: the New Deal and malaria control in the American South, 1933-1940. Parassitologia 40, 183-191 (1998).

39 Cumbler, J. T. Northeast and midwest United States: an environmental history. (Abc-clio, 2005).

40 Kukla, J. Kentish agues and American distempers: the transmission of malaria from England to Virginia in the seventeenth century. Southern studies 25, 135-147 (1986).

41 Yalcindag, E. et al. Multiple independent introductions of Plasmodium falciparum in South America. Proceedings of the National Academy of Sciences 109, 511-516 (2012).

42 Sinka, M. E. et al. The dominant Anopheles vectors of human malaria in the Americas: occurrence data, distribution maps and bionomic precis. Parasit Vectors 3, 72, doi:10.1186/1756-3305-3-72 (2010).

43 McNeill, J. R. Mosquito empires: ecology and war in the Greater Caribbean, 1620-1914. (Cambridge University Press, 2010).

44 Bukkuri, A. The History of Malaria in the United States: How it Spread, How it was Treated, and Public Responses. Vol. 2 (2018).

45 Taylor, D. South Carolina Naturalists: An Anthology, 1700-1860. (Univ of South Carolina Press, 1998).

46 Bell, A. M. Mosquito soldiers: malaria, , and the course of the American Civil War. (LSU Press, 2010).

47 Control, C. f. D. & Prevention. Elimination of malaria in the United States (1947–1951). February 8, 2010 (2010).

48 Carson, R. . (Houghton Mifflin Harcourt, 2002).

32

49 Inouye, D. W. Mosquitoes: more likely nectar thieves than pollinators. Nature 467, 27 (2010).

50 Gorham, J. R. Orchid pollination by Aedes mosquitoes in Alaska. American Midland Naturalist, 208-210 (1976).

51 Fang, J. Ecology: a world without mosquitoes. Nature News 466, 432-434 (2010).

52 Collins, C. M., Bonds, J. A. S., Quinlan, M. M. & Mumford, J. D. Effects of the removal or reduction in density of the malaria mosquito, Anopheles gambiae s.l., on interacting predators and competitors in local ecosystems. Medical and veterinary entomology 33, 1-15, doi:10.1111/mve.12327 (2019).

53 Oye, K. A. et al. Regulating gene drives. Science 345, 626-628, doi:10.1126/science.1254287 (2014).

54 Stewart, L. & Fugate, S. Stop screwworms: selections from the Screwworm Eradication Collection Special Collections National Agricultural Library. (Department of Agriculture, Washington, DC (EUA). Agricultural Research Service, 2000).

55 Krafsur, E. S., Whitten, C. J. & Novy, J. E. Screwworm eradication in North and Central America. Parasitology Today 3, 131-137 (1987).

56 Wyss, J. H. Screwworm eradication in the Americas. Annals of the New York Academy of Sciences 916, 186-193 (2000).

57 Franz, G. Genetic sexing strains in Mediterranean fruit fly, an example for other species amenable to large-scale rearing for the sterile insect technique. In: Sterile Insect Technique. ed.: Springer; 2005. p. 427-451.

58 Klassen, W. & Curtis, C. F. History of the Sterile Insect Technique. In: Sterile Insect Technique: Principles and Practice in Area-Wide Integrated Pest Management. ed. Dordrecht: Springer Netherlands; 2005. p. 3-36.

59 Harris, A. F. et al. Successful suppression of a field mosquito population by sustained release of engineered male mosquitoes. Nature biotechnology 30, 828-830, doi:10.1038/nbt.2350 (2012).

60 Malavasi, A. Project Aedes transgenic population control in Juazeiro and Jacobina Bahia, Brazil. BMC Proceedings 8, O11-O11, doi:10.1186/1753-6561-8-S4-O11 (2014).

61 CRISPR everywhere. Nature 531, 155, doi:10.1038/531155a (2016).

62 Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nature biotechnology 34, 78-83, doi:10.1038/nbt.3439 (2015).

63 Kyrou, K. et al. A CRISPR-Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature biotechnology 36, 1062-1066, doi:10.1038/nbt.4245 (2018).

33

64 Gantz, V. M. & Jasinskiene, N. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. Proceedings of the …, doi:10.1073/pnas.1521077112 (2015).

65 Champer, J., Buchman, A. & Akbari, O. S. Cheating evolution: engineering gene drives to manipulate the fate of wild populations. Nature Reviews Genetics 17, 146 (2016).

66 Knipling, E. Possibilities of insect control or eradication through the use of sexually sterile males. Journal of economic entomology 48, 459-462 (1955).

67 Balestrino, F. et al. Gamma ray dosimetry and mating capacity studies in the laboratory on Aedes albopictus males. J Med Entomol 47, 581-591 (2010).

68 Helinski, M. E. & Knols, B. G. Mating competitiveness of male Anopheles arabiensis mosquitoes irradiated with a partially or fully sterilizing dose in small and large laboratory cages. J Med Entomol 45, 698-705 (2008).

69 Yamada, H., Vreysen, M. J., Gilles, J. R., Munhenga, G. & Damiens, D. D. The effects of genetic manipulation, dieldrin treatment and irradiation on the mating competitiveness of male Anopheles arabiensis in field cages. Malaria journal 13, 318, doi:10.1186/1475-2875-13-318 (2014).

70 Andreasen, M. H. & Curtis, C. F. Optimal life stage for radiation sterilization of Anopheles males and their fitness for release. Medical and veterinary entomology 19, 238-244, doi:10.1111/j.1365-2915.2005.00565.x (2005).

71 Helinski, M. E., Parker, A. G. & Knols, B. G. Radiation biology of mosquitoes. Malaria journal 8, S6 (2009).

72 Proverbs, M. Induced sterilization and control of insects. Annual review of entomology 14, 81- 102 (1969).

73 Barclay, H. Mathematical models for the use of sterile insects. In: Sterile Insect Technique. ed.: Springer; 2005. p. 147-174.

74 Black, W. C. t., Alphey, L. & James, A. A. Why RIDL is not SIT. Trends Parasitol 27, 362-370, doi:10.1016/j.pt.2011.04.004 (2011).

75 Phuc, H. K. et al. Late-acting dominant lethal genetic systems and mosquito control. BMC biology 5, 11, doi:10.1186/1741-7007-5-11 (2007).

76 Thomas, D. D., Donnelly, C. A., Wood, R. J. & Alphey, L. S. Insect population control using a dominant, repressible, lethal genetic system. Science 287, 2474-2476, doi:DOI 10.1126/science.287.5462.2474 (2000).

77 Fu, G. et al. Female-specific flightless phenotype for mosquito control. Proceedings of the National Academy of Sciences of the United States of America 107, 4550-4554, doi:10.1073/pnas.1000251107 (2010).

34

78 Labbé, G. M., Scaife, S., Morgan, S. A., Curtis, Z. H. & Alphey, L. Female-specific flightless (fsRIDL) phenotype for control of Aedes albopictus. PLoS neglected tropical diseases 6, e1724 (2012).

79 Marinotti, O. et al. Development of a population suppression strain of the human malaria vector mosquito, Anopheles stephensi. Malaria journal 12, 142 (2013).

80 Carvalho, D. O. et al. Suppression of a Field Population of Aedes aegypti in Brazil by Sustained Release of Transgenic Male Mosquitoes. PLoS neglected tropical diseases 9, e0003864, doi:10.1371/journal.pntd.0003864 (2015).

81 Oxitec to Apply New Generation of Self-Limiting Mosquito Technology to Malaria-Spreading Mosquitoes [press release]. 2018).

82 Knipling, E. F. Sterile-Male Method of Population Control: Successful with some insects, the method may also be effective when applied to other noxious animals. Science 130, 902-904 (1959).

83 Alphey, L. Genetic control of mosquitoes. Annual review of entomology, doi:10.1146/annurev- ento-011613-162002 (2014).

84 Catteruccia, F. et al. Stable germline transformation of the malaria mosquito Anopheles stephensi. Nature 405, 959-962, doi:10.1038/35016096 (2000).

85 Adelman, Z. N. Genetic control of malaria and dengue. (Academic Press, 2015).

86 Wang, S. & Jacobs-Lorena, M. Genetic approaches to interfere with malaria transmission by vector mosquitoes. Trends in biotechnology 31, 185-193 (2013).

87 Burt, A. & Trivers, R. Genes in Conflict: The Biology of Selfish Genetic Elements: Harvard Univ. Press, Cambridge, MA (2006).

88 Burt, A. & Koufopanou, V. Homing endonuclease genes: the rise and fall and rise again of a selfish element. Current opinion in genetics & development 14, 609-615 (2004).

89 Lyttle, T. W. Segregation distorters. Annual review of genetics 25, 511-581 (1991).

90 Charlesworth, B. & Langley, C. H. The population genetics of Drosophila transposable elements. Annual review of genetics 23, 251-287 (1989).

91 Beeman, R., Friesen, K. & Denell, R. Maternal-effect selfish genes in flour beetles. Science 256, 89-92 (1992).

92 Chen, C.-H. et al. A synthetic maternal-effect selfish genetic element drives population replacement in Drosophila. science 316, 597-600 (2007).

93 Werren, J. H. Biology of wolbachia. Annual review of entomology 42, 587-609 (1997).

94 Gershenson, S. A new sex-ratio abnormality in Drosophila obscura. Genetics 13, 488 (1928).

35

95 Daniels, S. B., Peterson, K. R., Strausbaugh, L. D., Kidwell, M. G. & Chovnick, A. Evidence for horizontal transmission of the P transposable element between Drosophila species. Genetics 124, 339-355 (1990).

96 Colleaux, L. et al. Universal code equivalent of a yeast mitochondrial intron reading frame is expressed into E. coli as a specific double strand endonuclease. Cell 44, 521-533 (1986).

97 Boveri, T. Uber Differenzierung der Zellkerne wahrend der Furchung des Eies von Ascaris megalocephala. Anat. Anz. 2, 688-693 (1887).

98 Rhoades, M. Preferential segregation in maize. Genetics 27, 395 (1942).

99 Fredga, K., GROPP, A., WINKING, H. & FRANK, F. Fertile XX-and XY-type females in the wood lemming Myopus schisticolor. Nature 261, 225 (1976).

100 Craig, G., Hickey, W. & VandeHey, R. An inherited male-producing factor in Aedes aegypti. Science 132, 1887-1889 (1960).

101 Wood, R., Cook, L., Hamilton, A. & Whitelaw, A. Transporting the marker gene re (red eye) into a laboratory cage population of Aedes aegypti (Diptera: Culicidae), using meiotic drive at the MD locus. Journal of medical entomology 14, 461-464 (1977).

102 Sinkins, S. P. & Gould, F. Gene drive systems for insect disease vectors. Nature Reviews Genetics 7, 427 (2006).

103 Burt, A. Site-specific selfish genes as tools for the control and genetic engineering of natural populations. Proceedings of the Royal Society B: Biological Sciences 270, doi:10.1098/rspb.2002.2319 (2003).

104 Scott, T. W., Takken, W., Knols, B. G. & Boëte, C. The ecology of genetically modified mosquitoes. Science 298, 117-119 (2002).

105 Windbichler, N. et al. Homing endonuclease mediated gene targeting in Anopheles gambiae cells and embryos. Nucleic acids research 35, 5922-5933, doi:10.1093/nar/gkm632 (2007).

106 Windbichler, N., Papathanos, P. A. & Crisanti, A. Targeting the X chromosome during spermatogenesis induces Y chromosome transmission ratio distortion and early dominant embryo lethality in Anopheles gambiae. PLoS genetics 4, e1000291, doi:10.1371/journal.pgen.1000291 (2008).

107 Windbichler, N. et al. A synthetic homing endonuclease-based gene drive system in the human malaria mosquito. Nature 473, doi:10.1038/nature09937 (2011).

108 Li, J. et al. Genome-block expression-assisted association studies discover malaria resistance genes in Anopheles gambiae. Proceedings of the National Academy of Sciences 110, 20675- 20680 (2013).

109 Galizi, R. et al. A synthetic sex ratio distortion system for the control of the human malaria mosquito. Nature communications 5, 3977, doi:10.1038/ncomms4977 (2014).

36

110 Bernardini, F. et al. Site-specific genetic engineering of the Anopheles gambiae Y chromosome. Proceedings of the National Academy of Sciences of the United States of America 111, 7600- 7605, doi:10.1073/pnas.1404996111 (2014).

111 Hamilton, W. D. Extraordinary Sex Ratios. Science 156, 477-488, doi:10.1126/science.156.3774.477 (1967).

112 Newton, M., Wood, R. & Southern, D. A cytogenetic analysis of meiotic drive in the mosquito, Aedes aegypti (L.). Genetica 46, 297-318 (1976).

113 Hickey, W. A. & Craig, G. B., Jr. Genetic distortion of sex ratio in a mosquito, Aedes aegypti. Genetics 53, 1177-1196 (1966).

114 Wood, R. J. & Newton, M. E. Sex-ratio distortion caused by meiotic drive in mosquitoes. The American Naturalist 137, 379-391 (1991).

115 Lyttle, T. W. Experimental population genetics of meiotic drive systems I. Pseudo-Y chromosomal drive as a means of eliminating cage populations of Drosophila melanogaster. Genetics 86, 413-445 (1977).

116 Huang, Y., Magori, K., Lloyd, A. L. & Gould, F. INTRODUCING DESIRABLE TRANSGENES INTO INSECT POPULATIONS USING Y‐LINKED MEIOTIC DRIVE—A THEORETICAL ASSESSMENT. Evolution: International Journal of Organic Evolution 61, 717-726 (2007).

117 Chan, Y. S. et al. The Design and In Vivo Evaluation of Engineered I-OnuI-Based Enzymes for HEG Gene Drive. PloS one 8, e74254, doi:10.1371/journal.pone.0074254 (2013).

118 Thyme, S. B. et al. Reprogramming homing endonuclease specificity through computational design and directed evolution. Nucleic acids research 42, 2564-2576 (2013).

119 Takeuchi, R., Choi, M. & Stoddard, B. L. Redesign of extensive protein–DNA interfaces of meganucleases using iterative cycles of in vitro compartmentalization. Proceedings of the National Academy of Sciences 111, 4061-4066 (2014).

120 Kim, Y.-G., Cha, J. & Chandrasegaran, S. Hybrid restriction enzymes: zinc finger fusions to Fok I cleavage domain. Proceedings of the National Academy of Sciences 93, 1156-1160 (1996).

121 Smidler, A. L., Terenzi, O., Soichot, J., Levashina, E. A. & Marois, E. Targeted mutagenesis in the malaria mosquito using TALE nucleases. PloS one 8, e74511, doi:10.1371/journal.pone.0074511 (2013).

122 Aryan, A., Anderson, M., Myles, K. & Adelman, Z. TALEN-Based Gene Disruption in the Dengue Vector Aedes aegypti. PloS one 8, doi:10.1371/journal.pone.0060082 (2013).

123 Aryan, A., Myles, K. & Adelman, Z. Targeted genome editing in Aedes aegypti using TALENs. Methods 69, doi:10.1016/j.ymeth.2014.02.008 (2014).

37

124 DeGennaro, M., McBride, C. S., Seeholzer, L. & Nakagawa, T. orco mutant mosquitoes lose strong preference for humans and are not repelled by volatile DEET. Nature, doi:10.1038/nature12206 (2013).

125 McMeniman, C. J., Corfas, R. A., Matthews, B. J., Ritchie, S. A. & Vosshall, L. B. Multimodal integration of carbon dioxide and other sensory cues drives mosquito attraction to humans. Cell 156, doi:10.1016/j.cell.2013.12.044 (2014).

126 Volohonsky, G. et al. Tools for Anopheles gambiae Transgenesis. G3 (Bethesda, Md.), doi:10.1534/g3.115.016808 (2015).

127 Simoni, A. et al. Development of synthetic selfish elements based on modular nucleases in Drosophila melanogaster. Nucleic acids research 42, doi:10.1093/nar/gku387 (2014).

128 Jinek, M. et al. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337, 816-821, doi:10.1126/science.1225829 (2012).

129 Gasiunas, G., Barrangou, R., Horvath, P. & Siksnys, V. Cas9-crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proceedings of the National Academy of Sciences of the United States of America 109, E2579-2586, doi:10.1073/pnas.1208507109 (2012).

130 Cho, S. W., Kim, S., Kim, J. M. & Kim, J. S. Targeted genome engineering in human cells with the Cas9 RNA-guided endonuclease. Nature biotechnology 31, 230-232, doi:10.1038/nbt.2507 (2013).

131 Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819-823, doi:10.1126/science.1231143 (2013).

132 Jinek, M. et al. RNA-programmed genome editing in human cells. eLife 2, e00471, doi:10.7554/eLife.00471 (2013).

133 Mali, P. et al. RNA-guided human genome engineering via Cas9. Science 339, 823-826, doi:10.1126/science.1232033 (2013).

134 Deltcheva, E. et al. CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III. Nature 471, 602-607, doi:10.1038/nature09886 (2011).

135 Barrangou, R. et al. CRISPR provides acquired resistance against viruses in prokaryotes. Science 315, 1709-1712, doi:10.1126/science.1138140 (2007).

136 Adli, M. The CRISPR tool kit for genome editing and beyond. Nature communications 9, 1911 (2018).

137 Esvelt, K. M. & Smidler, A. L. (Google Patents, 2016).

138 Clayton, A. M., Dong, Y. & Dimopoulos, G. The Anopheles innate immune system in the defense against malaria infection. Journal of innate immunity 6, 169-181 (2014).

38

139 Blair, C. D. & Olson, K. E. Mosquito immune responses to arbovirus infections. Current opinion in insect science 3, 22-29 (2014).

140 Dong, Y. et al. Engineered anopheles immunity to Plasmodium infection. PLoS pathogens 7, e1002458 (2011).

141 Antonova, Y., Alvarez, K. S., Kim, Y. J., Kokoza, V. & Raikhel, A. S. The role of NF-κB factor REL2 in the Aedes aegypti immune response. Insect biochemistry and molecular biology 39, 303-314 (2009).

142 Kokoza, V. et al. Blocking of Plasmodium transmission by cooperative action of Cecropin A and Defensin A in transgenic Aedes aegypti mosquitoes. Proceedings of the National Academy of Sciences 107, 8111-8116 (2010).

143 Kim, W. et al. Ectopic expression of a cecropin transgene in the human malaria vector mosquito Anopheles gambiae (Diptera: Culicidae): effects on susceptibility to Plasmodium. Journal of medical entomology 41, 447-455 (2004).

144 Oliveira Gde, A., Lieberman, J. & Barillas-Mury, C. Epithelial nitration by a peroxidase/NOX5 system mediates mosquito antiplasmodial immunity. Science 335, 856-859, doi:10.1126/science.1209678 (2012).

145 Corby-Harris, V. et al. Activation of Akt signaling reduces the prevalence and intensity of malaria parasite infection and lifespan in Anopheles stephensi mosquitoes. PLoS pathogens 6, doi:10.1371/journal.ppat.1001003 (2009).

146 Luckhart, S. et al. Sustained activation of Akt elicits mitochondrial dysfunction to block Plasmodium falciparum infection in the mosquito host. PLoS pathogens 9, e1003180, doi:10.1371/journal.ppat.1003180 (2013).

147 Summers, K., Howells, A. & Pyliotis, N. Biology of eye pigmentation in insects. In: Advances in insect physiology. 16. ed.: Elsevier; 1982. p. 119-166.

148 Yamamoto, D. S., Sumitani, M., Hatakeyama, M. & Matsuoka, H. Malaria infectivity of xanthurenic acid-deficient anopheline mosquitoes produced by TALEN-mediated targeted mutagenesis. Transgenic research 27, 51-60 (2018).

149 Zieler, H., Keister, D. B., Dvorak, J. A. & Ribeiro, J. M. A snake venom phospholipase A(2) blocks malaria parasite development in the mosquito midgut by inhibiting ookinete association with the midgut surface. The Journal of experimental biology 204, 4157-4167 (2001).

150 Moreira, L. A. et al. Bee venom phospholipase inhibits malaria parasite development in transgenic mosquitoes. The Journal of biological chemistry 277, 40839-40843, doi:10.1074/jbc.M206647200 (2002).

151 Smith, R. C., Kizito, C., Rasgon, J. L. & Jacobs-Lorena, M. Transgenic mosquitoes expressing a phospholipase A(2) gene have a fitness advantage when fed Plasmodium falciparum-infected blood. PloS one 8, e76097, doi:10.1371/journal.pone.0076097 (2013).

39

152 de Lara Capurro, M. et al. Virus-expressed, recombinant single-chain antibody blocks sporozoite infection of salivary glands in Plasmodium gallinaceum-infected Aedes aegypti. The American journal of tropical medicine and hygiene 62, 427-433 (2000).

153 Isaacs, A. T. et al. Engineered resistance to Plasmodium falciparum development in transgenic Anopheles stephensi. PLoS pathogens 7, e1002017, doi:10.1371/journal.ppat.1002017 (2011).

154 Sumitani, M. et al. Reduction of malaria transmission by transgenic mosquitoes expressing an antisporozoite antibody in their salivary glands. Insect molecular biology 22, 41-51, doi:10.1111/j.1365-2583.2012.01168.x (2013).

155 Isaacs, A. T. et al. Transgenic Anopheles stephensi coexpressing single-chain antibodies resist Plasmodium falciparum development. Proceedings of the National Academy of Sciences of the United States of America 109, E1922-1930, doi:10.1073/pnas.1207738109 (2012).

156 Ghosh, A. K., Ribolla, P. E. & Jacobs-Lorena, M. Targeting Plasmodium ligands on mosquito salivary glands and midgut with a phage display peptide library. Proceedings of the National Academy of Sciences of the United States of America 98, 13278-13281, doi:10.1073/pnas.241491198 (2001).

157 Vega-Rodriguez, J. et al. Multiple pathways for Plasmodium ookinete invasion of the mosquito midgut. Proceedings of the National Academy of Sciences of the United States of America 111, E492-500, doi:10.1073/pnas.1315517111 (2014).

158 Fiorenza, M. T., Bevilacqua, A., Bevilacqua, S. & Mangia, F. Growing dictyate oocytes, but not early preimplantation embryos, of the mouse display high levels of DNA homologous recombination by single-strand annealing and lack DNA nonhomologous end joining. Developmental biology 233, 214-224, doi:10.1006/dbio.2001.0199 (2001).

159 Preston, C. R., Flores, C. C. & Engels, W. R. Differential usage of alternative pathways of double- strand break repair in Drosophila. Genetics 172, 1055-1068, doi:10.1534/genetics.105.050138 (2006).

160 Chan, Y. S., Huen, D. S., Glauert, R., Whiteway, E. & Russell, S. Optimising homing endonuclease gene drive performance in a semi-refractory species: the Drosophila melanogaster experience. PloS one 8, e54130, doi:10.1371/journal.pone.0054130 (2013).

161 Saleh-Gohari, N. & Helleday, T. Conservative homologous recombination preferentially repairs DNA double-strand breaks in the S phase of the cell cycle in human cells. Nucleic acids research 32, 3683-3688, doi:10.1093/nar/gkh703 (2004).

162 Champer, J. et al. Reducing resistance allele formation in CRISPR gene drive. Proceedings of the National Academy of Sciences 115, 5522-5527, doi:10.1073/pnas.1720354115 (2018).

163 Unckless, R. L., Clark, A. G. & Messer, P. W. Evolution of resistance against CRISPR/Cas9 gene drive. Genetics 205, 827-841 (2017).

164 Smith, J. M. & Price, G. R. The logic of animal conflict. Nature 246, 15 (1973).

40

165 Gantz, V. M. et al. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. PNAS Plus (2015).

166 Papathanos, P. A., Windbichler, N., Menichelli, M., Burt, A. & Crisanti, A. The vasa regulatory region mediates germline expression and maternal transmission of proteins in the malaria mosquito Anopheles gambiae: a versatile tool for genetic control strategies. BMC molecular biology 10, 65, doi:10.1186/1471-2199-10-65 (2009).

167 Hammond, A. M. et al. The creation and selection of mutations resistant to a gene drive over multiple generations in the malaria mosquito. PLoS genetics 13, e1007039, doi:10.1371/journal.pgen.1007039 (2017).

168 Champer, J. et al. Novel CRISPR/Cas9 gene drive constructs reveal insights into mechanisms of resistance allele formation and drive efficiency in genetically diverse populations. PLoS genetics 13, e1006796 (2017).

169 Hammond, A. M. et al. Improved CRISPR-based suppression gene drives mitigate resistance and impose a large reproductive load on laboratory-contained mosquito populations. bioRxiv, 360339 (2018).

170 Kyrou, K. et al. A CRISPR–Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature biotechnology 36, 1062, doi:10.1038/nbt.4245 (2018).

171 Nash, A. et al. Integral gene drives for population replacement. Biol Open 8, doi:10.1242/bio.037762 (2019).

172 Marshall, J. M., Buchman, A. & Akbari, O. S. Overcoming evolved resistance to population- suppressing homing-based gene drives. Scientific reports 7, 3776 (2017).

173 Oberhofer, G., Ivy, T. & Hay, B. A. Behavior of homing endonuclease gene drives targeting genes required for viability or female fertility with multiplexed guide RNAs. Proceedings of the National Academy of Sciences 115, E9343-E9352, doi:10.1073/pnas.1805278115 (2018).

174 Mali, P. et al. CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering. Nature biotechnology 31, 833-838, doi:10.1038/nbt.2675 (2013).

175 Zhang, X.-H., Tee, L. Y., Wang, X.-G., Huang, Q.-S. & Yang, S.-H. Off-target effects in CRISPR/Cas9- mediated genome engineering. Molecular Therapy-Nucleic Acids 4, e264 (2015).

176 Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nature biotechnology 31, 827 (2013).

177 Pattanayak, V. et al. High-throughput profiling of off-target DNA cleavage reveals RNA- programmed Cas9 nuclease specificity. Nature biotechnology 31, 839 (2013).

178 Fu, Y. et al. High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells. Nature biotechnology 31, 822-826, doi:10.1038/nbt.2623 (2013).

41

179 Kosicki, M., Tomberg, K. & Bradley, A. Repair of double-strand breaks induced by CRISPR–Cas9 leads to large deletions and complex rearrangements. Nature biotechnology 36, 765 (2018).

180 Ran, F. A. et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154, 1380-1389, doi:10.1016/j.cell.2013.08.021 (2013).

181 Sundaresan, R., Parameshwaran, H. P., Yogesha, S. D., Keilbarth, M. W. & Rajan, R. RNA- Independent DNA Cleavage Activities of Cas9 and Cas12a. Cell reports 21, 3728-3739, doi:10.1016/j.celrep.2017.11.100 (2017).

182 Ingvarsson, P. K. & Taylor, D. R. Genealogical evidence for epidemics of selfish genes. Proceedings of the National Academy of Sciences of the United States of America 99, 11265- 11269, doi:10.1073/pnas.172318099 (2002).

183 Charlesworth, B. Stabilizing selection, purifying selection, and mutational bias in finite populations. Genetics 194, 955-971 (2013).

184 Barzel, A., Obolski, U., Gogarten, J. P., Kupiec, M. & Hadany, L. Home and away- the evolutionary dynamics of homing endonucleases. BMC evolutionary biology 11, 324, doi:10.1186/1471-2148- 11-324 (2011).

185 Gogarten, J. P. & Hilario, E. Inteins, introns, and homing endonucleases: recent revelations about the life cycle of parasitic genetic elements. BMC evolutionary biology 6, 94, doi:10.1186/1471- 2148-6-94 (2006).

186 Gogarten, J. P., Senejani, A. G., Zhaxybayeva, O., Olendzenski, L. & Hilario, E. Inteins: structure, function, and evolution. Annual review of microbiology 56, 263-287, doi:10.1146/annurev.micro.56.012302.160741 (2002).

187 Liu, X. Q. & Hu, Z. A DnaB intein in Rhodothermus marinus: indication of recent intein homing across remotely related organisms. Proceedings of the National Academy of Sciences of the United States of America 94, 7851-7856 (1997).

188 Perler, F. B., Olsen, G. J. & Adam, E. Compilation and analysis of intein sequences. Nucleic acids research 25, 1087-1093 (1997).

189 Thiery, O., Borstler, B., Ineichen, K. & Redecker, D. Evolutionary dynamics of introns and homing endonuclease ORFs in a region of the large subunit of the mitochondrial rRNA in Glomus species (arbuscular mycorrhizal fungi, Glomeromycota). Molecular phylogenetics and evolution 55, 599- 610, doi:10.1016/j.ympev.2010.02.013 (2010).

190 Goddard, M. R. & Burt, A. Recurrent invasion and extinction of a selfish gene. Proceedings of the National Academy of Sciences of the United States of America 96, 13880-13885 (1999).

191 Koufopanou, V., Goddard, M. R. & Burt, A. Adaptation for horizontal transfer in a homing endonuclease. Molecular biology and evolution 19, 239-246, doi:10.1093/oxfordjournals.molbev.a004077 (2002).

42

192 Cho, Y. & Palmer, J. D. Multiple acquisitions via horizontal transfer of a group I intron in the mitochondrial cox1 gene during evolution of the Araceae family. Molecular biology and evolution 16, 1155-1165, doi:10.1093/oxfordjournals.molbev.a026206 (1999).

193 Palmer, J. D. et al. Dynamic evolution of plant mitochondrial genomes: mobile genes and introns and highly variable mutation rates. Proceedings of the National Academy of Sciences of the United States of America 97, 6960-6966 (2000).

194 Yamamoto, D. S., Sumitani, M., Kasashima, K., Sezutsu, H. & Matsuoka, H. Inhibition of Malaria Infection in Transgenic Anopheline Mosquitoes Lacking Salivary Gland Cells. PLoS pathogens 12, e1005872, doi:10.1371/journal.ppat.1005872 (2016).

195 Halic, M. & Moazed, D. Transposon silencing by piRNAs. Cell 138, 1058-1060, doi:10.1016/j.cell.2009.08.030 (2009).

196 Kalmykova, A. I., Klenov, M. S. & Gvozdev, V. A. Argonaute protein PIWI controls mobilization of retrotransposons in the Drosophila male germline. Nucleic acids research 33, 2052-2059, doi:10.1093/nar/gki323 (2005).

197 Al-Kaff, N. S. et al. Transcriptional and posttranscriptional plant gene silencing in response to a pathogen. Science 279, 2113-2115 (1998).

198 Baulcombe, D. RNA silencing in plants. Nature 431, 356 (2004).

199 Tabara, H. et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123-132 (1999).

200 Lankenau, S., Corces, V. G. & Lankenau, D.-H. The Drosophila micropia retrotransposon encodes a testis-specific antisense RNA complementary to reverse transcriptase. Molecular and cellular biology 14, 1764-1775 (1994).

201 Sijen, T. & Plasterk, R. H. Transposon silencing in the Caenorhabditis elegans germ line by natural RNAi. Nature 426, 310 (2003).

202 Vagin, V. V. et al. A distinct small RNA pathway silences selfish genetic elements in the germline. Science 313, 320-324, doi:10.1126/science.1129333 (2006).

203 Franz, A. W. et al. Stability and loss of a virus resistance phenotype over time in transgenic mosquitoes harbouring an antiviral effector gene. Insect molecular biology 18, 661-672 (2009).

204 Franz, A. W. et al. Fitness impact and stability of a transgene conferring resistance to dengue-2 virus following introgression into a genetically diverse Aedes aegypti strain. PLoS neglected tropical diseases 8, e2833, doi:10.1371/journal.pntd.0002833 (2014).

205 Hadjiargyrou, M. & Delihas, N. The intertwining of transposable elements and non-coding RNAs. International journal of molecular sciences 14, 13307-13328, doi:10.3390/ijms140713307 (2013).

43

206 Shah, N., Dorer, D. R., Moriyama, E. N. & Christensen, A. C. Evolution of a large, conserved, and syntenic gene family in insects. G3 (Bethesda) 2, 313-319, doi:10.1534/g3.111.001412 (2012).

207 Deutschbauer, A. M. et al. Mechanisms of haploinsufficiency revealed by genome-wide profiling in yeast. Genetics 169, 1915-1925, doi:10.1534/genetics.104.036871 (2005).

208 Cooper, G. M., Hausman, R. E. & Hausman, R. E. The cell: a molecular approach. Vol. 2 (ASM press Washington, DC, 2000).

209 Lambertsson, A. The minute genes in Drosophila and their molecular functions. Advances in genetics 38, 69-134 (1998).

210 Marygold, S. J. et al. The ribosomal protein genes and Minute loci of Drosophila melanogaster. Genome biology 8, R216, doi:10.1186/gb-2007-8-10-r216 (2007).

211 Saeboe-Larssen, S., Lyamouri, M., Merriam, J., Oksvold, M. P. & Lambertsson, A. Ribosomal protein insufficiency and the minute syndrome in Drosophila: a dose-response relationship. Genetics 148, 1215-1224 (1998).

212 Estep, A. S., Sanscrainte, N. D. & Becnel, J. J. DsRNA-mediated targeting of ribosomal transcripts RPS6 and RPL26 induces long-lasting and significant reductions in fecundity of the vector Aedes aegypti. J Insect Physiol 90, 17-26, doi:10.1016/j.jinsphys.2016.05.001 (2016).

213 Rothschild, J. B., Tsimiklis, P., Siggia, E. D. & Francois, P. Predicting ancestral segmentation phenotypes from drosophila to anopheles using in silico evolution. PLoS genetics 12, e1006052 (2016).

214 Manoukis, N. C. et al. Structure and dynamics of male swarms of Anopheles gambiae. Journal of medical entomology 46, 227-235 (2009).

215 Focks, D. A. An improved separator for the developmental stages, sexes, and species of mosquitoes (Diptera: Culicidae). J Med Entomol 17, 567-568 (1980).

216 Google. Debug Fresno, our first U.S. field study, (July 14 2017).

217 Zacarés, M. et al. Exploring the potential of computer vision analysis of pupae size dimorphism for adaptive sex sorting systems of various vector mosquito species. Parasites & vectors 11, 656 (2018).

218 Senecio-Robotics. Senecio Compact State-of-The-Art Robotic Sex Sorting,

219 Marois, E. et al. High-throughput sorting of mosquito larvae for laboratory studies and for future vector control interventions. Malaria journal 11, 302, doi:10.1186/1475-2875-11-302 (2012).

220 Catteruccia, F., Benton, J. P. & Crisanti, A. An Anopheles transgenic sexing strain for vector control. Nature biotechnology 23, 1414-1417, doi:10.1038/nbt1152 (2005).

44

221 Bernardini, F. et al. Cross-species Y chromosome function between malaria vectors of the Anopheles gambiae species complex. Genetics 207, 729-740 (2017).

222 Galizi, R. et al. A CRISPR-Cas9 sex-ratio distortion system for genetic control. Scientific reports 6, 31139 (2016).

45

CHAPTER 2.

CRISPR-mediated germline knockouts for genetic sterilization of male Anopheles gambiae.

Andrea Smidler1,2, Doug Paton1, George Church2, Kevin Esvelt3, William R. Shaw1,†, Flaminia Catteruccia1,†

1Department of Immunology and Infectious Diseases, Harvard T.H. Chan School of Public Health, Boston,

MA 02115, USA

2Department of Genetics, Harvard Medical School, Boston, MA 02115, USA

3Media Lab, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

† These authors contributed equally to this work

Respective Contributions: This chapter was written by Andrea Smidler with editorial input from Flaminia

Catteruccia. ALS conceived of the study and performed all transgene design, cloning, transgenesis, mutant analysis, husbandry, data analysis, and drafted the manuscript. ALS and KE together performed transgene design. ALS and WRS executed the study. DP performed . GMC contributed to the study and provided CRISPR expertise. FC contributed to the design of the study, edited the manuscript, and provided entomological and reproductive expertise. Portions of this chapter were based off a published work performed in collaboration with other colleagues:

1 Werling, K. et al. Steroid Hormone Function Controls Non-competitive Plasmodium Development in Anopheles. Cell, doi:10.1016/j.cell.2019.02.036.

46

ABSTRACT

Mosquito control has consistently proven to be the most effective way to prevent the transmission of mosquito-borne diseases. In the fight against malaria, control of the Anopheles mosquito with the use of long-lasting insecticide treated bed nets (LLINs) and indoor residual spraying (IRS) has helped reduce the number of new cases by almost 40% since the turn of the century. However, LLIN- and IRS strategies are being rendered increasingly ineffective by the spread of insecticide resistance in mosquitoes, making development of novel vector control technologies important for continued control of this disease. Insect control programs targeting other species have relied on releasing sterilized males who cause infertility in the females with whom they mate. This technique, termed Sterile Insect technique (SIT) has, however, failed in Anopheles species, largely due to strong fitness defects in released males caused by the sterilization process. Here we outline a novel CRISPR-based genetic sterilization system for the selective ablation of male sperm cells in Anopheles gambiae. Specifically, robust mosaic mutagenesis of the ZGP

(AGAP006241) gene in F1 individuals resulting from a cross of a germline-expressing Cas9 line and ZPG- targeting gRNAs, resulted in complete sterility in 95% of mutant males. Mating of zpg males to wild type females caused high levels of female infertility suggesting potential application for vector control strategies although analysis of fertile mutant zpg individuals showed fertility-maintaining mutations. With further optimization, this genetic system could be adopted as the basis for sterile male release campaigns to combat the spread of malaria.

47

INTRODUCTION

Strategies that target insect vectors are central to the control of vector-borne diseases. Among control strategies targeting mosquito fertility, Sterile Insect Technique (SIT), a strategy based on the mass release of irradiated males to sterilize females upon mating2, has been successfully implemented against a variety of burdensome agricultural insect pests3-5 (previously discussed in Section 1.4A). Iterative mass release of sterilized males in the wild results in infertility in females, causing population suppression. However adapting this technology to mosquitoes has been challenging, mostly because the use of chemo- and irradiation-based sterilization reduces the mating competitiveness of males6-10.

To circumvent the potential problems linked to sterility by irradiation or chemosterilization, strategies based on genetic manipulation of insect fertility are being developed. Lethal transgenic systems have been developed in the arbovirus vectors, Aedes (discussed in Section 1.4B), which preserve male fertility but kill their offspring at later, post-embryonic developmental stages11,12. Among the most promising strategies are gene drives based on CRISPR/Cas9 genome editing technology13,14 which are capable of spreading infertility through natural mosquito populations by destroying endogenous genes required for fertility while positively biasing their own inheritance (previously discussed in Section 1.5C). The potential of gene drives to induce sterility in An. gambiae populations has been recently demonstrated under laboratory conditions, through the biased inheritance of female-sterilizing and androgenizing factors13,14, however there are multiple issues related to the use of these genetic systems. Firstly, they are prone to rapid deactivation by the evolution of resistance mutations that render them not functional14. Secondly, the self-autonomous propagation properties of a successful gene drive makes it difficult if not impossible to control once released in the wild15, raising concerns for their spread across national boundaries16 and supporting the need for safer genetic sterilization systems for the control of Anopheles mosquitoes. In this work we aimed to develop safe, controllable, and non-invasive CRISPR-based technologies for genetic control of wild An. gambiae populations.

48

The gene Zero Population Growth (ZPG) has been shown to be required for germ cell development in

Drosophila17,18 and mosquitoes19. It is a germline-specific gap junction innexin which plays a crucial role in early germ cell differentiation and survival17. ZPG is composed of four transmembrane domains which participate in forming pores on the surface of cells in the developing germline, through which close-range signaling occurs with the adjacent somatic support cells to promote homeostatic germ cell development17.

Silencing ZPG by transient RNAi injections in An. gambiae resulted in sterile males with phenotypically atrophied testes. Importantly, these males initiated typical post-mating responses in females following copulation19, making this an ideal gene target for genetic sterilization. These males were also shown to have similar mating competitiveness to wild type (wt) males19. However, the embryo-microinjection procedure required to generate these males is laborious and is not conducive to large-scale studies.

Therefore, here we set out to develop a transgenic CRISPR system to mutate ZPG and generate large numbers of sterile An. gambiae males. Crossing germline Cas9 and ZPG-targeting gRNA strains resulted in synchronous RNA-guided dominant biallelic knockouts of ZPG in the developing germline, fully preventing sperm development in 95% of males. Moreover, our system generated heritable gene knockouts in F2 progeny, further supporting the use of this type of system to generate stable knockout lines for reverse genetic studies20.

49

RESULTS

2.1 Generation of transgenic CRISPR-expression lines

To generate CRISPR spermless males we created two transgenic lines, one expressing Cas9 specifically in the germline and another expressing three gRNAs targeting ZPG (AGAP006241) as previously described1.

The Cas9 line (termed VasCas9) is characterized by a selectable neuronal 3xP3-DsRed marker, and expresses Cas921 via the Vas2 promoter22 which enables robust expression in germ cells of both males and females (Figure 2.1A). (NB: This transgenic is used throughout the work described in this dissertation). A second transgenic line, termed gZPG, expresses of three gRNAs (a-c) by the Pol III U6 promoter

(AGAP013557)13 designed to target two sequences (the first exon and the 3’UTR) in the endogenous ZPG locus, to generate large deletions and premature stop codons. This transgene is visually identifiable by a

3xP3-EYFP selectable marker which enables fluorescent identification in all individuals throughout the lifecycle, and additional Vasa2-EYFP marker for fluorescent visualization of the germline (Figure 2.1B).

Figure 2.1 | Transgenic map and fluorescence properties of VasCas9 and gZPG transgenic lines, previously described in1. [A] The VasCas9 germline expressing transgenic. It is similarly designed to previous germline endonuclease expression systems developed for gene knockout in An. gambiae23. The canonical Vasa2 promoter22 (indigo) composed of 2,300 bp upstream of the Vasa gene (AGAP008578) expresses S. pyogenes Cas9 (salmon)24, followed by an Sv40 (tan) transcription terminator. It is cloned into the pDSAR transgenesis plasmid25, providing a neuronal 3xP3-DsRed selectable marker, and inserted into the neutral X1 transgenic docking site by phiC31 transgenesis25. The larval fluorescence pattern is shown at right. [B] The gZPG gRNA-expressing transgenic line. It uses the Pol III U6 promoter composed of 322 bp upstream of AGAP013557 (navy) to express a cassette of two tethered gRNAs, gRNAa and gRNAc, and a second U6 promoter expressed a third gRNA, gRNAb (blue). A germline fluorescent marker is included composed of the Vas2 promoter (indigo) for EYFP expression (chartreuse), which can be easily visualized in the developing male germline of late age larvae (white arrows). A 3xP3-EYFP fluorescent selectable marker is included for neuronal transgenic identification (pale blue, and chartreuse) and the fluorescent pattern of transgenic larvae is shown at right. It is cloned into the pDSAY An. gambiae transgenesis plasmid25, and inserted into the neutral X13 docking site by phiC31 trasngenesis25.

50

2.2 ZPG mutants fail to develop normal testes

Crossing gZPG males to VasCas9 females yields F1 progeny with the {+/gZPG; +/VasCas9} genotype harboring a single copy of each transgene, identified by dual 3xP3-EYFP and 3xP3-DsRed fluorescence.

These F1 individuals undergo significant mosaic mutagenesis which results in failure to develop normal testes, which can be observed in the pupal stage as the absence of fluorescence from the Vas2-EYFP reporter in {+/gZPG;+/VasCas9} males (Figure 2.2A). Dissecting these tissues from 2-day old adult males reveals atrophied testes with no visible sperm in contrast to wt controls (Figure 2.2B).

Figure 2.2| The spermless male testicular phenotype observed by pupal fluorescence and following dissection. [A] Absence of developed testes indicated by absence of testis-derived Vas2-EYFP fluorescence in the ventral pupal tail. A comparison of the fluorescence pattern of individuals with the {+/gZPG;+/VasCas9} genotype (left) and control siblings not inheriting the VasCas9 transgene with the {+/gZPG;+/+} genotype (right). Individuals with both transgenes characterized by both the 3xP3-EYFP and 3xP3-DsRed neuronal selection markers (diamond- like stripe) fail to manifest any observable testicular-derived Vas2-EYFP expression (blue arrows). Control siblings inheriting just the gZPG transgene, evidenced by the absence of the 3xP3-DsRed selectable marker (dotted outline), demonstrate significant Vas2-EYFP testicular fluorescence from the cassette on gZPG (white arrows). [B] Following dissection of two-day old males, the absence of sperm in the testes (ii, blue arrows) can be observed in {+/gZPG;+/VasCas9} males compared to controls (i, white arrows). The male accessory glands are labeled as ‘MAGs’ for clarity. Panel (vi) is a visualization of the absence of sperm in the testes shown in (ii), and is a composite of the inset Differential Inference Contrast (DIC) (iii), DAPI stained (iv) and Vas2-EYFP marker (v).

51

Sequencing the germline of {+/gZPG;+/VasCas9} individuals reveals a variety of CRISPR-induced mutations, mostly large deletions between the gRNA target sites (Figure 2.3) but also some insertions as well as small deletions. Notably, many mutations appear localized over gRNA-b, suggesting advantageous cleavage properties. Often we observed different mutations within the same individual, suggesting that active mosaic mutagenesis is occurring throughout the development of the individual mosquito.

Figure 2.3 | A composite image of mutations observed in {+/gZPG;+/VasCas9} male testes, with sequencing reads aligned (red, bottom) to annotated gene sequence (blue, top). Exons introns and 3’UTR are annotated in navy, dashed-lines, and grey pale blue respectively. Portions of the sequencing image are removed for clarity, with vertical grey bars denoting number of bases removed. gRNA binding sequence shown in orange and yellow with gRNAa and gRNAb targeting Exon1 (respectively) and gRNAc targeting 3’UTR.

2.3 Male ZPG mutants are highly sterile

The absence of visible sperm in {+/gZPG; +/VasCas9} males (Figure 2.2B) suggests they may be sterile, making them candidates for use in SIT programs. To test this, we released 26 {+/gZPG;+/VasCas9} males

(3-day old) into a cage with 175 wt virgin females, and allowed them to mate for 2 nights. Females were then blood fed and allowed to lay eggs. Out of the 4,132 eggs laid only 126 larvae hatched (3% fertility), suggesting males were highly sterile. To determine whether every male had some degree of fertility or these larvae were sired by a few fully fertile males, we performed forced mating assays with individual spermless (n=26) and wt (n=17) males to wt females and assayed for fertility phenotypes. While the vast majority of females (96%, n=25/26) were completely sterile, some showed normal fertility levels and produced broods with the expected transgene ratio, demonstrating that some {+/gZPG; +/VasCas9} males maintain normal levels of fertility (Figure 2.4A). Additional control experiments confirmed that sterility is a product of both transgenes being inherited and expressed within a single male (Figure 2.4B).

52

Figure 2.4 | Genetically spermless males cause significant infertility in females following copulation independent of each transgene alone. [A] Percent fertility of individual females force-mated to {+/gZPG;+/VasCas9} (n =25) or wild type ({+/+} males (n=17). Of the {+/gZPG; +/VasCas9} males assayed, 96% were completely sterile (n = 24/25) with a single escapee fertile male siring a brood with normal percent fertility (blue asterisk). The broods sired by wild type males were 87% fertile on average (SD ± 0.2), and was statistically different from the genetically sterile males (p <0.0001, Mann Whitney). [B] Infertility is not caused by the individual gZPG or VasCas9 transgenes alone. Percent fertility of individual females force-mated to {+/gZPG; +/VasCas9} (n=15) , {VasCas9/VasCas9} (n=12), or {gZPG/gZPG} (n=8) was scored. Of the {+/gZPG; +/VasCas9} males assayed, 93% were completely sterile (n = 14/15) with a single escapee fertile male siring a brood with normal percent fertility (blue asterisk). The broods sired by {VasCas9/VasCas9} and {gZPG/gZPG} were 93% (SD ± 0.1) and 79% (SD ± 3.3) fertile respectively, which were both significantly different than the fertility observed in {+/gZPG;+/VasCas9} ( p > 0.0001 and p>0.05 respectively, Multi-Way AVOVA).

2.4 Characterization of fertile ZPG mutant males

To determine the cause of the residual fertility in otherwise genetically sterile males, we sequence- analyzed the residual testicular tissue or offspring of {+/gZPG; +/VasCas9} males to determine the nature of the mutations present. This analysis revealed multiple possible causes for residual male fertility, including lack of mutations and mutations which did not cause frame shifts or stop codons. Some

{+/gZPG;+/VasCas9} males showed a dimorphic phenotype in their testes, with one spermless and one phenotypically normal testis, suggesting mutagenesis was occurring separately within each tissue. In one individual, sequencing each testis separately revealed that the atrophied testis harbored an 8bp deletion

53 under gRNA-C, while the phenotypically normal testis had no identifiable CRISPR mutations. Sequencing the somatic tissues of this male revealed it was heterozygous for a naturally occurring 2bp Asp53Gly polymorphism within the gRNA-C binding site, suggesting that gRNA-C binding over this sequence may have been prevented in the fertile testis and had instead successfully occurred in the other tissue. At large, however, these findings suggest that some males remain fertile because of fundamental failure to mutate both copies of ZPG, and that using gRNAs targeting sites devoid of naturally occurring polymorphisms may enable complete male sterilization. Redesign of the system to target highly conserved regions within this gene could also minimize such escapees.

Interestingly, sequencing progeny from escapee {+/gZPG;+/VasCas9} fertile males and wild type females revealed larvae heterozygous for a 69 bp deletion between gRNA-A and gRNA-C roughly corresponding to the putative 4th transmembrane domain of ZPG17. Because this deletion would likely not affect localization of the other transmembrane domains nor would it affect the extracellular cysteines critical for innexin pore formation (Reviewed in26), we postulate that this ZPG mutant might permit sufficient protein function required for fertility, although further characterization would be required to confirm this.

However, this does present further proof that germline CRISPR-expression systems can generate heritable mutations for stable mutant lines, with applications for enabling reverse genetic studies20.

2.5 Female ZPG mutants display severe ovarian atrophy

While not otherwise discussed in this work, ZPG mutant sisters display significant ovarian phenotypes and merit brief discussion. Female {+/gZPG; +/VasCas9} individuals undergo active mosaic mutagenesis during development resulting in ovaries with varied phenotypes (Figure 2.5). In females 48h hours post blood- feeding, significantly less yolk and lipid deposition is observed compared to control females, and often varies for each follicle (Figure 2.5A). On closer inspection, some ovaries are partially ablated Figure 2.5B), while others experience complete ablation of one lobe and development of the other (Figure 2.5C,D).

54

Consistent the role of the ovaries in synthesis of the important insect hormone, 20-hydroxyecdysone

(20E)27,28, females with the {+/gZPG; +/VasCas9} genotype contain significantly less 20E than controls 28h following blood-feeding regardless of phenotype (Figure 2.5E), and when binned by phenotype, females with no ovarian development in both lobes have almost none1. Taken together, these females provide a valuable biological tool for study of mosquito reproduction and in the role of ovarian tissue signaling on important physiological processes.

Figure 2.5 | ZPG mutant females have significantly ablated ovarian development and are defective in 20E synthesis. [A] Ovary pairs from seven mutant ZPG females adjacent to wild type ovaries 48h post-blood feeding. Mutant ZPG ovaries display varied phenotypes ranging in lipid and yolk deposition levels, with some lobes completely failing development. Select ovaries, B, C and D inset expanded for analysis. [B] Mutant ZPG ovaries display polymorphic phenotypes consistent with differential mutagenesis within different segments of ovaries. Here the bottom half of one lobe is completely ablated (blue arrow), with upper half of lobe developed. [C,D] Here the right lobe in each ovary pair is completely ablated undergoing no yolk or lipid deposition in contrast to the opposing lobe. [E] Females with the ovaryless genotype display significantly decreased 20E levels 28h post-bloodfeeding compared to controls (p<0.01, Unpaired T-test, Mean with SEM)

55

DISCUSSION

Here we outline a system for generating genetically sterilized male Anopheles gambiae for biological study and as candidates for SIT release. We show that crossing transgenics expressing Cas9 in the germline, to transgenics expressing gRNAs targeting ZPG, produces a majority of sterile males in the F1 progeny. These sterile males have atrophied testes, no observable sperm, and harbor numerous CRISPR-generated mutant alleles that arise by active mosaic mutagenesis during development. We show that approximately

95% of these males completely sterilize the females they whom they mate. Coupled with previous research that suggests ZPG knockdown males have comparable mating fitness to wild type males19, this suggests that – following optimization of the system – ZPG mutant males could be utilized in SIT programs.

We further show that ZPG fertile males can arise by multiple mechanisms including failure to mutate potentially due to inhibited gRNA binding in the presence of SNPs, and the acquisition of de novo CRISPR- generated fertility-maintaining mutations. Finally, we demonstrate that this system can be used to create heritable knockouts in a gene of interest, evidenced by heterozygous mutant offspring sired by Zpg mutant males, a fundamental tool for basic biological studies in this important vector species 23.

Anophelines mate in large swarms where competition between males is fierce29. While previous work in the lab demonstrate that ZPG knockdown males without testes maintained similar fitness properties19, the transgenes used here may effect males fitness in ways dsRNA knockdown did not, so future analysis of male mating competitiveness is critical. Therefore, this genetic system requires similar validation to confirm them as better alternatives to chemo- and irradiation-based sterilization procedures.

This system requires optimization before release, as the 5% of males escaping sterilization is an unacceptably large proportion for mass release programs. To increase sterility, gRNAs could be targeted to less polymorphic sites, and in patterns that prevent function-maintaining mutations. Further, gRNAs targeting additional germline-essential genes could be supplemented, reducing mutagenic escape by

56 orders of magnitude with each additional gene targeted. Such genes could be any anopheline homologs to those required for the Drosophila male germline, including Tudor (AGAP008268), B2-tubulin

(AGAP008622), or perhaps even Vasa itself (AGAP00857). However before this can be undertaken, understanding the properties required for optimal gRNA cleavage in anophelines is critical. Others have shown that gRNAs vary in their mutagenic potential30, an observation also supported by our findings. This could be caused by multiple factors in gRNA design whose elucidation is beyond the scope of this work, but it underscores the need for further research into anopheline gRNA optimization.

A major hurdle for development of genetic sterilization systems is that maintenance of purely sterile, male-only, lines is impossible by nature of their inability to breed. Therefore, all such systems require some degree of inducibility to trigger female killing and to suppress sterility until immediately prior to release. RIDL systems in Aedes achieve this by induction of a lethal transgene following release, which is suppressed by addition of tetracycline during rearing12. In our system, because Δzpg knockout males and females are infertile, inducibilty is achieved by crossing together two different transgenic populations, which alone do not have fertility defects, making mass rearing more efficient. However, our system does not currently address the requirement for female removal. Female killing was demonstrated in similar systems developed contemporaneously in Drosophila31, and following the recent elucidation of sex determination factors in anophelines32,33, introducing these types of modifications to our ZPG mutant system could be implemented with relative ease.

Beyond potential vector control applications, this ZPG mutant system could be used to explore a variety of other biological questions of interest. Firstly, the role of sperm in initiating female-post mating responses is still largely unexplored. An. gambiae females display two major responses after copulation; the stimulation of oviposition following blood-feeding, and the induction of refractoriness to further mating; both initiated following sexual transfer of factors from the male to the female atrium during copulation34. Although a previous study showed that sperm is not involved in triggering these female

57 responses19, the use of transgenic spermless males may help identify more subtle effects linked to sperm transfer and storage. Indeed, in Drosophila sperm is needed to extend the mating refractoriness period to up to a week due to male-transferred proteins bound to sperm tails which are released slowly over time35-37. Application of this genetic sterilization system could therefore be used to study the effect of sperm on similar post-mating responses in female mosquitoes, opening up an intriguing avenue of research of some importance to the vector control community.

Our study shows that with correct target selection and precise temporal expression, CRISPR can generate mosaic biallelic knockout of entire tissues, in this case generating genetically spermless males. In this way our system is somewhat analogous to some ‘floxed’ murine models38, revealing the potential for this type of system to be a similarly powerful tool to study basic insect biology. Its adaptation to knockout other organs could enable salivary gland-less mosquitoes39, those lacking haemocytes40, or male accessory glands, which could revolutionize the basic research in this important insect pest, and may lead to insights across a spectrum of topics of interest to entomologists.

Our work presents a novel tool for inducible genetic sterilization in the important An. gambiae disease vector, making SIT programs to suppress this mosquito possible. This system facilitates efficient mass rearing, and robust sterilization phenotypes upon induction. While it requires significant optimization to maximize male sterility, and modification to kill females, it provides a valuable proof-of-principle that similar transgenic sterilization systems hold promise to enable genetic SIT in this deadly vector.

FUTURE DIRECTIONS

For successful field releases, near complete sterilization of males must be achieved. To accomplish this, we attempted development of a RIDL-inspired, germline-specific, tetracycline-inducible genetic rescue system. Though not otherwise discussed in this work, the transgenic, TOZR (Tet-Off ZPG Rescue), was

58 developed to provide a suppressible copy of ZPG when in a Δzpg mutant background, and was composed of the ZPG promoter and mRNA separated by a Tet-Off switch41. However we discovered TOZR did not respond to tetracycline supplementation in contrast to similar systems developed in other tissues42, suggesting that the privileged germline tissues43,44 may be impermeable to tetracycline regulation, narrowing the potential suite of tools available to develop effective inducible genetic sterilization .

Therefore to meet the needs of modern SIT, we theoretically developed a transgenic crossing system to mass produce genetically sterilized – but not transgenic – male-only populations. This high-throughput system builds upon the ZPG mutant system described above, but does not rely on flawless mosaic gRNA mutagenesis, relies on proven RIDL technologies, kills all females, and generates wholly non-transgenic males with guaranteed sterility. We have termed it the Inducible Eunuch system, and it uses specific crosses between two lines to make the sterile male-only populations required for SIT (Figure 2.6). The two lines, αline and βline, both carry biallelic Δzpg, a transgenic rescue copy of ZPG, a 3xP3-DsRed marker, and tet-suppressible RIDL, in a balancer site maintaining 50% transgene frequency in the stock population.

The αline specifically carries a 3xP3-EYFP marker and an additional RIDL construct on the Y chromosome, while the βline carries the same transgene on the X. Both stock lines would be stable and have fertility maintained by the heterozygous dominant ZPG-rescue masking the underlying Δzpg phenotype.

Fluorescent sorting larvae45 of the αline for a specific subset of females, and the βline for a subset of males would generate the parental genotypes and sexes needed to initiate the final sterile male cross. Crossing these males and females together in the absence of tetracycline would result in all daughters dying from the X-lined induction of RIDL, and death of all fertile genotypes by induction of the RIDL cassette associated with the ZPG-rescue, leaving only the Δzpg non-transgenic sons alive for release. The small nonsense Δzpg mutation carried by these males would make them indistinguishable from naturally occurring mutants, making possible the argument that they are a ‘milder’ GMO. This may provide an ideal system for SIT-like campaigns in mosquitoes and other insects in the future. Having developed a method

59 to generate ZPG knockouts, and having developed docking lines which can act as balancers (outlined in

Chapter 3), construction of such a system may be possible in Anopheles gambiae in the near future.

Figure 2.6 | Schematic for the Inducible Eunuch system. This system allows for mass rear and crossing of two lines with endogenous ZPG knocked out, which results in killing of all transgenics and females, leaving only non- transgenic sterile males for release. Fluorescent sorting larvae of the αline with a single copy of 3xP3-DsRED and non-3xP3-EYFP by COPAS45 would yield only ZPG-rescue heterozygous females, and sorting larvae from βline with the 3xP3-DsRED and 3xP3-EYFP pattern would yield only heterozygous males. This single high-throughput sorting step would generate the parental genotypes and sexes needed to initiate the final sterile male cross. Crossing these males and females together in the absence of tetracycline would result in all females dying by nature of RIDL construct induction from the X chromosome, and all those fertile individuals inheriting a ZPG-rescue cassette would die by nature of adjacent RIDL induction. This would leave the only surviving genotype as males, who have not inherited any transgenes, and who therefore manifest the Δzpg phenotype.

60

METHODS

Generation of transgenic mosquito lines gRNA design: Design of gRNAs for these lines was previously reported in1. Briefly, the ZPG locus (AGAP

006241) was PCR amplified and sequenced across multiple individuals within our An. gambiae lines to identify any SNPs present. Putative gRNA candidates were identified by in silico tools available through the Broad Institute (https://portals.broadinstitute.org/gpp/public/analysis-tools/sgrna-design) and Zhang

Laboratory at MIT (http://crispr.mit.edu46). Three gRNA targets were chosen to maximize the probability of mutagenesis early in the coding sequence, with the additional aim of achiaeving large deletions. Two gRNA candidates were chosen, gRNAa and gRNAc, targeting the sequences (5’

GCGGCTTCACTGTCGTGTGACGG 3’) and (5’ CCGATCGACTGCGTGATCGGATC 3’) within Exon 1 located 71 b and 150 bp from the stop codon respectively. They were further chosen for their localization over semi- unique restriction enzyme sites AleI and PvuI respectively to enable PCR-based identification of mutants, as previously described in23. gRNAb (5’ CCAAGTGTTTGCATTCCTGGCGG 3’) was designed to target the

3’UTR sequence to facilitate generation of large deletions. gRNAs under the control of the U657 promoter concurrently developed for use by others13 (322 bp upstream of AGAP013557) were ordered as gBlocks

(Integrated DNA technologies, Skokie, IL) in two cassettes. gRNAa and gRNAb were synthesized as a tethered pair connected by a 21 bp sequence ( 5’ TTCACTGTGCGCATTATATAT 3’) predicted to not interfere with gRNA folding secondary structure (RNAfold)47. The gRNAc was synthesized as an isolated gBlock.

Plasmid Construction: As described previously1, plasmids were constructed using standard molecular biological techniques GoldenGate cloning48,49, into the An. gambiae transgenesis plasmids pDSAY and pDSAR25. VasCas9: To build VasCas9, the 2.3 kB Vasa2 promoter22 was PCR amplified from genomic DNA using primers (5’ CAGGTCTCAATCCCGATGTAGAACGCGAG 3’) and (5’

CGGTCTCACATATTGTTTCCTTTCTTTATTCACCGG 3’) and was cloned immediately upstream of SpCas9

61 amplified from plasmid PX165 (Addgene #48137)21 using primers (5’

CAGGTCTCATATGGACTATAAGGACCACGACGGAG 3’) and (5’

CAGGTCTCAAAGCTTACTTTTTCTTTTTTGCCTGGCC 3’). These fragments were Goldengate cloned into the multiple cloning site of the pDSAR vector which provides an Sv40 terminator for protein transcription termination, an attB site to facilitate phiC31 transgenesis into well established An. gambiae transgenesis docking lines containing an attP, and a 3xP3-DsRed fluorescence selectable marker 25. gZPG: To build gZPG, the two previously discussed gRNA-containing gBlocks were Golden gate cloned into the multiple cloning site of the pDSAY transgenesis plasmid25. To facilitate in vivo validation of the presence or absence of a germline, a Vas2-EYFP fluorescence cassette was further cloned into the unique AscI site on the pDSAY plasmid backbone by Golden gate ligation. For this cassette, the Vas2 promoter was PCR amplified using the primers (5’ CGGTCTCACGCGCGATGTAGAACGCGAGCAAA 3’) and (5’

CGGTCTCACATATTGTTTCCTTTCTTTATTCACCGG 3’) and EYFP was PCR amplified with ( 5’

CAGGTCTCAATGGTGAGCAAGGGCG 3’) and (5’ CAGGTCTCAAAGCTTACTTGTACAGCTCGTCCATGCC 3’).

Complete plasmids were sequence verified by Macrogen Sequencing services (Rockville, MD, USA).

Transgenesis: Transgenesis procedures were carried out effectively as described in 1,50,51. The gZPG construct (350 ng/µl) was co-injected with a phiC31-integrase expressing helper plasmid (80 ng/µl) into the posterior end of >3h-old aligned X13 docking line25 An. gambiae embryos (n=1663), and the VasCas9 plasmid (350 ng/µl) was similarly injected into X1 docking line25 embryos (n=2585). Survivors were reared to adulthood and outcrossed in bulk to large cages of wild type G3 virgin adults (n>1000) of the opposite sex. New transformants were identified and isolated as newly hatched larvae in the subsequent F1 generation by fluorescence. First generation F1 transformants were outcrossed to wild type G3 to introduce genetic diversity before intercrossing in the subsequent F2 generation to establish homozygous lines. Homozygous lines were established by identification of homozygous individuals by fluorescence intensity and subsequent PCR verification.

62

Generation of spermless {+/gZPG;+/VasCas9} males

To generate spermless males in bulk, {gZPG/gZPG} males were crossed to {VasCas9/VaCas9} females.

Females were chosen to give the VasCas9 transgene because Vas2 maternal deposition of Cas9 protein into the developing embryo causes earlier and more significant mutagenesis resulting in a higher penetrance of the spermless phenotype. Genotype of the resulting offspring were confirmed by dual

{3xP3-EYFP and 3xP3-DsRed fluorescence}, and males were sex separated during the pupal stage to guarantee virginity.

Microscopy

Imaging of transgenic larvae and ventral pupal tails was carried out under a Leica M80 fluorescence dissecting microscope following immobilization on ice and positioning by paintbrush. Imaging of microscopic testes structure was carried out under a Zeiss Inverted Observer Z1.

Mutation Analysis

Male {+/gZPG;+/VasCas9} mutant testes or surviving larvae were analysed for mutations by PCR and sequenced for mutations. DNA extraction was carried out using the Qiagen DNeasy Blood & Tissue Kit, and PCR was carried out using a variety of primers flanking the ZPG locus ranging in fragment size from

700 bp to 5 kB, including the forwrd primers [5’ CGTTTTCTTCACTCTCGGCACG 3’], [5’

GCAGCTTCTGGTAGTCGATGTCG 3’], and [5’ CCATTCGTTTGTTGCTGAAAGC 3’], and reverse primers [5’

GACCAGAAGCCGGAAAAGATC 3’], [5’ GAGGAACGCGGGTTTTTTTG 3’], and [5’

GTGAAATGTTTGGGCCCGATC 3’]. PCR products were cloned into the CloneJet PCR Cloning Kit

(ThermoFisher Scientific) to isolate PCR products corresponding to individual alleles, and plated on

Ampicillin (100 µg/mL) LB media plates. Individual colonies were either picked, cultured in liquid media, plasmid miniprepped (SpinSmart Plasmid Miniprep DNA Purification kit, Denville Scientific) and sequenced using the universal pJET2.1F or pJET2.1R primers (Macrogen USA), or the entire agar plate was

63 sent for direct colony sequencing (Macrogen USA). Resulting sequencing reads were aligned to an annotated Snapgene 3.2.1 file of the ZPG gene sequence.

Infertility mating assays

Bulk mating: 30 Male {+/gZPG;+/VasCas9} were sexed as pupae and allowed to eclose into a 25cm x 25cm

BugDorm cage (MegaView Science co, Taiwan). Four drowned as pupae, leaving 26 living males for the experiment. Female wild type pupae of the G3 line were sexed on the same day and allowed to eclose in a separate cage. The absence of contaminating males was confirmed the next morning, and 176 females were removed and mouth aspirated into the cage containing the {+/gZPG;+/VasCas9} males. They were allowed to mate ad litibum for 4 nights, and were bloodfed on day 5 until significant diuresis was observed.

An oviposition site consisting of a Whatman® filter paper cone (90mm, Grade 2, Sigma-Aldrich) within a urinalysis cup containing 80 mL DI water was inserted into the cage on day 7. The oviposition cup was removed day 8, and larvae were counted and scored for transgene frequency day 9. Eggs were counted on day 11 and 12.

Individual forced-mating assays: 5 day-post eclosure virgin males of their respective phenotypes and blood-fed virgin wild type females of the G3 line were force-mated to guarantee paternity (method available at https://www.beiresources.org/MR4Home.aspx). Male carcasses were saved for subsequent mutation analysis. Successful mating was validated by verification of the mating plug in the female atrium by autoflourescence visible through the female cuticlce (previously demonstrated in 34) and females were isolated to oviposit within individual paper cups lined with filter paper and filled with 1 cm DI water. The number of eggs laid and larvae hatched were counted from each female’s brood, and larvae screened for transgene fluorescence to verify paternity. Escapee larvae sired by genetically sterilized

{+/gZPG;+/VasCas9} fathers were collected for subsequent sequence analysis.

64

REFERENCES

1 Werling, K. et al. Steroid Hormone Function Controls Non-competitive Plasmodium Development in Anopheles. Cell, doi:10.1016/j.cell.2019.02.036.

2 Knipling, E. F. Sterile-male method of population control. Science 130, 902-904 (1959).

3 Krafsur, E. S., Whitten, C. J. & Novy, J. E. Screwworm eradication in North and Central America. Parasitol Today 3, 131-137 (1987).

4 Lindquist, D. A., Abusowa, M. & Hall, M. J. The screwworm fly in Libya: a review of its introduction and eradication. Medical and Veterinary Entomology 6, 2-8 (1992).

5 Hendrichs, J., Franz, G. & Rendon, P. Increased effectiveness and applicability of the sterile insect technique through male‐only releases for control of Mediterranean fruit during fruiting seasons. Journal of Applied Entomology 119, 371-377 (1995).

6 Abdel-Malek, A., Tantawy, A. & Wakid, A. Studies on the Eradication of Anopheles pharoensis by the Sterile-Male Technique Using Cobalt-60. III. Determination of the Sterile Dose and Its Biological Effects on Different Characters Related to‟ Fitness” Components. Journal of economic entomology 60, 20-23 (1967).

7 Sharma, V., Razdan, R. & Ansari, M. Anopheles stephensi: effect of gamma-radiation and chemosterilants on the fertility and fitness of males for sterile male releases. Journal of economic entomology 71, 449-452 (1978).

8 Reisen, W. K. Lessons from the past: an overview of studies by the University of Maryland and the University of California, Berkeley. Ecological aspects for application of genetically modified mosquitoes, 25-32 (2003).

9 Ageep, T. B. et al. Participation of irradiated Anopheles arabiensis males in swarms following field release in Sudan. Malaria journal 13, 484, doi:10.1186/1475-2875-13-484 (2014).

10 Munhenga, G. et al. Mating competitiveness of sterile genetic sexing strain males (GAMA) under laboratory and semi-field conditions: Steps towards the use of the Sterile Insect Technique to control the major malaria vector Anopheles arabiensis in South Africa. Parasites & Vectors 9, 122, doi:10.1186/s13071-016-1385-9 (2016).

11 Fu, G. et al. Female-specific flightless phenotype for mosquito control. Proceedings of the National Academy of Sciences of the United States of America 107, 4550-4554, doi:10.1073/pnas.1000251107 (2010).

12 Phuc, H. K. et al. Late-acting dominant lethal genetic systems and mosquito control. BMC biology 5, 11, doi:10.1186/1741-7007-5-11 (2007).

13 Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nature biotechnology 34, 78-83, doi:10.1038/nbt.3439 (2015).

65

14 Kyrou, K. et al. A CRISPR–Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature biotechnology 36, 1062, doi:10.1038/nbt.4245 (2018).

15 Noble, C., Adlam, B., Church, G. M., Esvelt, K. M. & Nowak, M. A. Current CRISPR gene drive systems are likely to be highly invasive in wild populations. eLife 7, doi:10.7554/eLife.33423 (2018).

16 Oye, K. A. et al. Regulating gene drives. Science, doi:10.1126/science.1254287 (2014).

17 Tazuke, S. I. et al. A germline-specific gap junction protein required for survival of differentiating early germ cells. Development 129, 2529-2539 (2002).

18 Mukai, M. et al. Innexin2 gap junctions in somatic support cells are required for cyst formation and for egg chamber formation in Drosophila. Mechanisms of development 128, 510-523, doi:10.1016/j.mod.2011.09.005 (2011).

19 Thailayil, J., Magnusson, K., Godfray, H., Crisanti, A. & Catteruccia, F. Spermless males elicit large-scale female responses to mating in the malaria mosquito Anopheles gambiae. Proceedings of the National Academy of Sciences 108, doi:10.1073/pnas.1104738108 (2011).

20 Dong, Y., Simões, M. L., Marois, E. & Dimopoulos, G. CRISPR/Cas9 -mediated gene knockout of Anopheles gambiae FREP1 suppresses malaria parasite infection. PLoS pathogens 14, e1006898, doi:10.1371/journal.ppat.1006898 (2018).

21 Ran, F. A. et al. Genome engineering using the CRISPR-Cas9 system. Nat Protoc 8, 2281-2308, doi:10.1038/nprot.2013.143 (2013).

22 Papathanos, P. A., Windbichler, N., Menichelli, M., Burt, A. & Crisanti, A. The vasa regulatory region mediates germline expression and maternal transmission of proteins in the malaria mosquito Anopheles gambiae: a versatile tool for genetic control strategies. BMC molecular biology 10, 65, doi:10.1186/1471-2199-10-65 (2009).

23 Smidler, A. L., Terenzi, O., Soichot, J., Levashina, E. A. & Marois, E. Targeted mutagenesis in the malaria mosquito using TALE nucleases. PloS one 8, e74511, doi:10.1371/journal.pone.0074511 (2013).

24 Ren, X. et al. Optimized gene editing technology for Drosophila melanogaster using germ line- specific Cas9. Proceedings of the National Academy of Sciences of the United States of America 110, 19012-19017, doi:10.1073/pnas.1318481110 (2013).

25 Volohonsky, G. et al. Tools for Anopheles gambiae Transgenesis. G3 (Bethesda, Md.), doi:10.1534/g3.115.016808 (2015).

26 Dahl, G. & Muller, K. J. Innexin and pannexin channels and their signaling. FEBS letters 588, 1396-1402 (2014).

66

27 Riehle, M. A. & Brown, M. R. Insulin stimulates ecdysteroid production through a conserved signaling cascade in the mosquito Aedes aegypti. Insect biochemistry and molecular biology 29, 855-860 (1999).

28 Brown, M. R. et al. Identification of a steroidogenic neurohormone in female mosquitoes. The Journal of biological chemistry 273, 3967-3971 (1998).

29 Manoukis, N. C. et al. Structure and dynamics of male swarms of Anopheles gambiae. Journal of medical entomology 46, 227-235 (2009).

30 Li, M., Akbari, O. S. & White, B. J. Highly Efficient Site-Specific Mutagenesis in Malaria Mosquitoes Using CRISPR. G3 (Bethesda) 8, 653-658, doi:10.1534/g3.117.1134 (2018).

31 Kandul, N. P. et al. Transforming insect population control with precision guided sterile males with demonstration in flies. Nat Commun 10, 84, doi:10.1038/s41467-018-07964-7 (2019).

32 Criscione, F., Qi, Y. & Tu, Z. GUY1 confers complete female lethality and is a strong candidate for a male-determining factor in Anopheles stephensi. eLife 5, e19281 (2016).

33 Krzywinska, E., Dennison, N. J., Lycett, G. J. & Krzywinski, J. A maleness gene in the malaria mosquito Anopheles gambiae. Science 353, 67-69 (2016).

34 Gabrieli, P. et al. Sexual transfer of the steroid hormone 20E induces the postmating switch in Anopheles gambiae. Proceedings of the National Academy of Sciences of the United States of America 111, 16353-16358, doi:10.1073/pnas.1410488111 (2014).

35 Liu, H. & Kubli, E. Sex-peptide is the molecular basis of the sperm effect in Drosophila melanogaster. Proceedings of the National Academy of Sciences 100, 9929-9933 (2003).

36 Chapman, T. et al. The sex peptide of Drosophila melanogaster: female post-mating responses analyzed by using RNA interference. Proceedings of the National Academy of Sciences 100, 9923- 9928 (2003).

37 Peng, J. et al. Gradual release of sperm bound sex-peptide controls female postmating behavior in Drosophila. Current Biology 15, 207-213 (2005).

38 Leneuve, P. et al. Cre‐mediated germline mosaicism: a new transgenic mouse for the selective removal of residual markers from tri‐lox conditional alleles. Nucleic acids research 31, e21-e21 (2003).

39 Myat, M. M., Isaac, D. D. & Andrew, D. J. Early genes required for salivary gland fate determination and morphogenesis in Drosophila melanogaster. Adv Dent Res 14, 89-98, doi:10.1177/08959374000140011501 (2000).

40 Terriente-Felix, A. et al. Notch cooperates with Lozenge/Runx to lock haemocytes into a differentiation programme. Development 140, 926-937, doi:10.1242/dev.086785 (2013).

41 Gossen, M. et al. Transcriptional activation by tetracyclines in mammalian cells. Science 268, 1766-1769 (1995).

67

42 Lycett, G. J., Kafatos, F. C. & Loukeris, T. G. Conditional expression in the malaria mosquito Anopheles stephensi with Tet-On and Tet-Off systems. Genetics 167, 1781-1790, doi:10.1534/genetics.104.028175 (2004).

43 Kaur, G., Mital, P. & Dufour, J. M. Testisimmune privilege - Assumptions versus facts. Animal reproduction 10, 3-15 (2013).

44 Milholland, B. et al. Differences between germline and somatic mutation rates in humans and mice. Nature Communications 8, 15183, doi:10.1038/ncomms15183 (2017).

45 Marois, E. et al. High-throughput sorting of mosquito larvae for laboratory studies and for future vector control interventions. Malaria journal 11, 302, doi:10.1186/1475-2875-11-302 (2012).

46 Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nature biotechnology 31, 827 (2013).

47 Hofacker, I. L. et al. Fast folding and comparison of RNA secondary structures. Monatshefte für Chemie/Chemical Monthly 125, 167-188 (1994).

48 Engler, C. & Marillonnet, S. Combinatorial DNA assembly using Golden Gate cloning. Methods Mol Biol 1073, 141-156, doi:10.1007/978-1-62703-625-2_12 (2013).

49 Geissler, R. et al. Transcriptional activators of human genes with programmable DNA-specificity. PloS one 6, e19509, doi:10.1371/journal.pone.0019509 (2011).

50 Fuchs, S., Nolan, T. & Crisanti, A. Mosquito transgenic technologies to reduce Plasmodium transmission. Methods Mol Biol 923, 601-622, doi:10.1007/978-1-62703-026-7_41 (2013).

51 Pondeville, E. et al. Efficient PhiC31 integrase-mediated site-specific germline transformation of Anopheles gambiae. Nat Protoc 9, 1698-1712, doi:10.1038/nprot.2014.117 (2014).

68

CHAPTER 3. Advancement towards developing evolutionarily stable gene drives for Anopheles gambiae population replacement

Andrea L. Smidler1,2, Eryney A. Marrogi2, Enzo Mameli1, George M. Church2, Kevin Esvelt3,†, Flaminia

Catteruccia1,†

1Department of Immunology and Infectious Diseases, Harvard T.H. Chan School of Public Health, Boston,

MA 02115, USA

2Department of Genetics, Harvard Medical School, Boston, MA 02115, USA

3Media Lab, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

† Co-corresponding, please direct genetics inquiries to KME, and entomologic inquiries to FC

Respective Contributions: This chapter was written by Andrea Smidler, with editorial input from Flaminia

Catteruccia. ALS and KME conceived of the study and designed drive architectures. ALS performed all transgene design, cloning, transgenesis, husbandry, data analysis, and drafted the manuscript. ALS and

EAM executed the study and Western blots. EM assisted in embryonic microinjections. GMC contributed to the study, conceived the CrIGCkid phenomenon, and provided CRISPR expertise. FC contributed to the design of the study and provided entomological and reproductive expertise. Portions of this chapter include ideas from published Research Articles co-written by Andrea Smidler in collaboration with colleagues. These articles are included in full in the Appendix for reference.

APPENDIX D

1 Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife, e03401, doi:10.7554/eLife.03401 (2014).

69

ABSTRACT

Artificial selfish-genetic elements enabled by CRISPR/Cas9 promise to spread of disease-refractory genes into wild mosquito populations, providing a novel tool to combat malaria. All gene drives described to date have either already selected for resistance mutations, or are likely to do so in the near future, necessitating development of improved designs. While some drive designs seek to eliminate local populations, replacement drives instead disperse neutral cargoes through them, imposing fewer fitness costs and increasing the probability of success. In this work, we develop a suite of tools to construct, and attempt generation of, evolutionary stable gene drives for population replacement. We target the haplolethal and haplosufficient RpL11 and Rpt1 genes for insertion – but not disruption – with a drive. As these genes are vastly intolerant to missense and nonsense mutations, successful drive homing is the only survivable outcome for the cell, preventing and deleting resistance alleles in the process. We first make gene drive docking lines in this genetically intractable locus, recoding RpL11 and Rpt1 during knockin, making them resistant to self-cleavage and forcing HDR to delete resistance alleles. For this, we modify the HACK system developed in Drosophila for use as a novel knockin method in Anopheles gambiae using

CRISPR DNA cleavage to stimulate interlocus gene conversion (IGC) from transgenic donors, and term the technology CrIGCkid. Using CrIGCkid we integrate multiple gene drive docking lines, and characterize the requirements for reproducibility, however these docking sites may cause aberrant lethality following gene drive integration, and their phenotypes may preclude drive in females. We further develop multiple germline promoters for Cas9 expression, and attempt ectopic drive of the docking sites with them, revealing the critically narrow expression profile required for drive-enabling Cas9 in this system; as lethality, infertility, or failure to drive was observed from all promoters tested. Our findings suggest that multi-gRNA drive designs may not stimulate HDR in the manner initially predicted, and that targeting haplolethal genes during spermatogenesis may be fundamentally incompatible with fertility as putative driving individuals are highly infertile and the only viable larvae recovered had escaped mutagenesis.

70

Finally, we observe that phenotypes reliant on robust gRNA expression lessen over time, consistent with silencing of transgenes expressing long noncoding RNAs observed by others, raising doubts about the long-term viability of CRISPR gene drives. Taken together, our results stress the numerous difficulties associated with the design and construction of evolutionarily stable gene drives, and demonstrates that constructing successful gene drives still has many hurdles to overcome.

INTRODUCTION

A large fraction of the research on Anopheles innate immunity aims to develop tools to make mosquitoes incapable of malaria transmission. Once transmission-blocking genetic factors have been identified, selfish genetic elements can be used to spread them through wild populations rendering entire species refractory to pathogen transmission (Section 1.5). These gene drive systems would replace disease- susceptible populations with disease-refractory ones through the biased inheritance of an anti-malarial cargo, aiding vector control campaigns. The advent of CRISPR/Cas9 technology (Section 1.5C) has dramatically increased the chances of developing and releasing population replacement gene drives. As observed in recent studies2,3, CRISPR/Cas9 gene drives however face numerous hurdles, as for instance they are plagued by nearly immediate evolution of genetic resistance to drive spread3 (Section 1.6). These include de-novo mutations under the gRNA binding site which occur as the result of DNA break repair by

End Joining (EJ) instead of the Homology Directed Repair (HDR) pathway required for successful drive homing. In the presence of these resistant mutations (drive-resistant alleles, DRAs), gene drives cannot home and are ultimately outcompeted to extinction in the population3. While recent drive designs have attempted to circumvent the issue of DRAs by homing into and knocking-out a conserved splice site critical for female development4 (Section 1.7), such designs are not suitable for population replacement strategies which require maintenance of essential gene function. And in light of modeling suggesting four

71 gRNAs are required for population replacement5, designs using less should be considered susceptible to

DRA-based inhibition and not evolutionarily stable6.

To enable more effective CRISPR-mediated gene drives and delay the emergence of DRAs, we designed a gene drive system inserted within, but not disrupting the function of, a haplolethal and a haplosufficient essential gene. Because these genes do not tolerate mutations, successful drive copying would be the only viable outcome for the cell – and that any nonlethal DRAs (r1DRAs) created in the process would be deleted (Section 1.9).

RESULTS

3.1 Gene drive design

Here we develop evolutionarily stable drives for population replacement in An. gambiae based on designs previously proposed1 (Appendix D), and proven in yeast7. Our evolutionarily stable drive design is inserted within – but does not disrupt the function of – a haplolethal gene (Figure 3.1). We delete a C-terminal segment and 3’UTR, as well as the intervening intergenic sequence, and replace it with our gene drive architecture. This includes reconstitution of protein coding sequence with a recoded DNA sequence8, and surrogate replacement 3’UTRs from a paralog (Figure 3.1A). Replacing the endogenous coding sequence with a divergent recoded one maintains endogenous protein function9, prevents cutting of the drive- containing chromosome, and prevents HDR within the region by forcing it to occur beyond the recoding boundary; enabling deletion of r1DRAs during homing (Figure 3.1C). Reintroduction of a paralogous surrogate 3’UTR maintains homeostatic protein regulation while minimizing homology to the homolog.

The CRISPR-encoding gene drive components including germline Cas9, gRNAs, and any associated genetic cargo, would be immediately adjacent (Figure 3.1B), occupying the area corresponding to the intergenic sequence deleted from wild type. The gRNAs target the wild type haplolethal C-terminal sequence at multiple, non-overlapping, sites (Figure 3.1D) while being unable to cleave the drive-containing

72 chromosome due to recoding. If DNA breaks repair by any number of EJ mechanisms most mutations disrupt protein function and cause cell death (Figure 3.1F), prohibiting transmission of r2DRAs to the next generation. In this system, because most haplolethal genes cannot tolerate any missense or nonsense mutations, r1DRAs are limited to SNPs in wobble bases, necessitating simultaneous r1DRAs at all four gRNA targets for full resistance; a highly improbable event. So long as cleavage by one gRNA initiates HDR, successful homing – which must occur beyond the recoding (Figure 3.1C, magenta arrows) – deletes all r1DRAs that have arisen (Figure 3.1E). In effect, this design makes most DRAs lethal, and deletes the nonlethal during homing, making successful drive the only outcome permissive of cell survival. To our knowledge, such a design is the only homing-based architecture capable of evolutionarily stable population replacement proposed to date1. Other replacement systems that increase evolutionary stability by targeting cellular EJ-repair mechanisms for suppression10, or use single gRNAs6, will still permit

DRA emergence at some frequency.

Figure 3.1 | An evolutionarily stable CRISPR gene drive design for population replacement. At top are the two homologs, one drive-containing and one wild type. [A] DNA sequence recoding (red bars) replaces C-terminal sequence of an essential gene (mint). [B] Cas9 and multiple gRNAs are encoded on the drive (Cas9, salmon; gRNAs, red shades). [C] Boundaries of homology indicating where HDR can occur (also indicated left). [D] The haplolethal wild type target recognized by gRNAs and Cas9(salmon PacMan). [E] Outcome of successful drive homing into the homolog results in a homozygous cell. [F] Outcome of EJ repair results in lethal mutation.

73

3.2 Gene drive docking site locus identification

Our design necessitates insertion into the C-terminal end of a haplolethal gene for selection pressure against DRA formation. Further enhancement is achievable if both drive ends are inserted into haplolethal genes, necessitating identification of a locus with two such genes in a tail-to-tail orientation. Ideal haplolethal genes are those involved in ribosome protein synthesis (Section 1.9), however we failed to identify any such gene pairs in the annotated An. gambiae genome11. Expanding the search to include conserved haplosufficient genes, we identified the RpL11-Rpt1 locus on chromosome 2L with the 60S ribosomal subunit (AGAP011173) facing the 26S proteasome subunit T1 (AGAP011174). Though less ideal than adjacent haplolethal genes, haplosufficient Rpt1 would still provide selection against DRAs.

3.3 Designing two docking sites in RpL11-Rpt1 – benefits and drawbacks of each design

To develop evolutionarily stable gene drives within this locus, we pursued drive insertion as a two-step process (Figure 3.2). Step 1 creates a gene drive docking site by HDR-based knockin, replacing the endogenous 3’ coding sequences and UTRs with recoding, a fluorescent selectable marker, and two attPs for subsequent phiC31 drive transgenesis12 (Figure 3.2, Step 1), permitting study of recoding’s effect in isolation from a drive. Step 2 inserts the CRISPR gene drive into the docking site by phiC31 RCME transgnesis13,14, exchanging the docking site fluorescent marker for the drive (Figure 3.2, Step 2).

74

Figure 3.2 | Constructing gene drives into RpL11-Rpt1 dRPLTu shown for example. Step 1 introduces recoding (red bars), replacement 3’UTRs (violet, orange), a flourescent marker (cyan) and attP sites (black) for transgenesis, by CRISPR-mediated HDR (magenta arrows). The corresponding wild type sequence is deleted from the chromosome during insertion. Step 2 introduces the CRISPR-encodinging gene drive by phiC31 transgenesis (navy arrows) via the attP and attB sites (black) inserting the drive and exchanging the corresponding flourecent reporters.

A number of assumptions were made when developing this drive design. The first assumption, shared by many scientists developing gene drives, is that expression of CRISPR constructs is maintained across many generations (i.e. the behavior of a line at generation 1 (G1) should be similar to its behavior at G10). The second assumption is that the recoding and surrogate 3’UTRs can fully recapitulate endogenous gene function. The third assumption is that CRISPR-expressing gene drives inserted into the docking sites trasngenically (Figure 3.2 Step 2) should not show position effects15. The fourth assumption is that multi- gRNA designs can stimulate robust HDR (although this assumption has recently been questioned16). The fifth assumption is that targeting ribosome and proteasome genes in the primordial germ line (the cell type targeted for drive) is compatible with fertility. And the final assumption is that, because successful drive homing is the only survivable outcome for the cell, CRISPR expression in any cell types incapable of

HDR will cause cell death and may affect mosquito fitness accordingly. Through the course of this work, we demonstrate evidence that many of these assumptions may be incorrect.

75

3.3 A. Designing dRPLT – a gene drive docking site with Rpt1 knocked-out. We developed two gene drive docking lines in RpL11-Rpt1 following the basic design outlined above, with the properties of each summarized in Table 3.1. The first, dRPLT (docking site in RpL11-Rpt1), lacks a 3’UTR replacement for

Rpt1, rendering heterozygotes functionally hemizygous for the proteasome subunit. Rpt1 hemizygocity makes the remaining copy haploinsufficient and thereby of DRA-intolerant, however also rendering dRPLT

– and inserted drives – homozygotes inviable. Drive could still be tested by use of a single-sex-specific promoter for Cas9 expression, permitting heterozygous dominant trait fixation instead of homozygous allelic fixation, and would maintain the most stringent selective pressures against DRA-formation.

However, homozygous lethality of this line would make it cumbersome to rear, difficult to manipulate, and potentially unfit for field release, motivating development of a second alternative docking line, termed dRPLTu, containing a complete 3’ UTR (dRPLT phenotype characterization follows in Section 3.5A).

Table 3.1 | Summary and comparison of characteristics of the dRPLT and dRPLTu gene drive docking sites.

76

3.3 B. Designing dRPLTu – a gene drive docking site maintaining Rpt1 function. The second docking line, dRPLTu (docking site in RpL11-Rpt1 with UTR), is identical to dRPLT but exchanges a m2Turquoise fluorescent selectable marker (henceforth CFP) for EYFP, and adds a replacement 3’UTR from Rpt3. dRPLTu was designed to maintain wild type-like Rpt1 function, permitting establishment of homozygous lines. Drives in this site could spread to allelic fixation, but will permit the emergence of r2DRAs in Rpt1.

3.4 Knockin of dRPLT docking line Into RpL11-Rpt1 by HDR

We first pursued knockin of the dRPLT docking line template into embryos by standard HDR-based knockin methods17,18. The HDR-donor plasmid, HDRdRPLTdonor, contains design features outlined in Section 3.3A, and is mapped in Figure 3.2 and Appendix A. Injecting HDRdRPLTdonor and in vitro-synthetized gRNAs targeting RpL11 and Rpt1 into 2,732 VasCas9 embryos (Section 2.1) resulted in high embryonic lethality

(8.9% survival compared to standard 15-33% survival with other transgenes), and failed to result in knockin progeny. Further PCR on carcasses of injected individuals revealed no evidence of dRPLT- integrated cells. These results suggest RpL11-Rpt1 is at least partially genetically intractable to delivery of

DNA by this method, and knockin by direct embryo injection may not be achievable. Indeed An. gambiae transgenesis injections occur into the syncytial blastoderm such that most nuclei undergo mutagenesis while few knockin by HDR, killing the embryo, and necessitating development of single-cell knockin technologies. Therefore to knockin the docking lines we designed a new method based on the cellular phenomenon of interlocus gene conversion (IGC), inspired by a similar system developed in Drosophila19, to achieve integration of a distal donor transgene into our target locus.

3.5 Developing transgenic Interlocus Gene Conversion (IGC) for knockin into genetically intractable loci

Inspired by a press release from Editas Medicine20 which reported a novel CRISPR-induced DNA repair mechanism of beta with hemoglobin delta, we decided to recapitulate this gene conversion phenomenon transgenically for knockin of HDRdRPLTdonor.

77

To harness interlocus gene conversion (IGC) for knockin in An. gambiae, we designed a transgenic crossing scheme to provide the donor template, gRNAs, and Cas9 transgenically in the adult germline with fluorescent markers orientated for identification of gene conversion events in the offspring (Putative mechanism, Figure 3.3). We term this system CrIGCkid (CRISPR-mediated Interlocus Gene Conversion- derived Knockin; /ˈkrɪkɪd/). For this we modified the HDRdRPLTdonor to contain the Rpt1 3’UTR surrogate for dRPLTu (termed HDRdRPLTudonor), and added to both donors 8 gRNAs targeting RpL11-Rpt1, a second fluorescent marker, act-DsRed, outside the homology arms of the donor cassette.* These transgenes were subcloned into piggyBac transposons21 for semi-random transgenesis throughout the genome, and were termed IGCdRPLTpB and IGCdRPLTupB respectively (Figure 3.3, Figure 3.4 respectively, Appendix A).

We hypothesized that crossing IGCdRPLTpB- or IGCdRPLTupB-containing transgenics to VasCas9 may enable knockin of the donor template – including recoding and associated 3xP3-FP† marker – into RpL11-Rpt1 in the adult germline by an IGC-like mechanism. Successful repair of RpL11-Rpt1 by IGC would copy the 3xP3-

FP marker into the locus – leaving the act-DsRed marker outside of the homology arms unintegrated – allowing for fluorescent screening of knockin larvae by a 3xP3-FP-positive, act-DsRed-negative pattern.

These larvae would be the desired dRPLT and dRPLTu lines, and could be confirmed by PCR. However

CrIGCkid from donor transgenes on the same chromosome as RpL11-Rpt1 (chromosome 3) would be difficult to screen, as knockin larvae would still be positive for act-DsRed from the linked donor. However,

PCR could identify such ‘hidden’ docking site larvae, and would provide insight into the dynamics of

CRISPR-induced IGC gene knockin. Transgenesis injections established four unique IGCdRPLTpB transgenic families and five donor IGCdRPLTupB families (named IGCdRPLTupB1, IGCdRPLTupB2 IGCdRPLTupB3,

IGCdRPLTupB4, and IGCdRPLTupB5) to generate dRPLT and dRPLTu respectively via CrIGCkid.

* The act-DsRed uses the Actin5c promoter for intense expression in the gut and the rest of body. † † Because dRPLT and dRPLTu use different fluorescent proteins for selection (3xP3-EYFP, 3xP3-CFP, respectively) but are similar in function, when referring to fluorescence of either line, it will be shortened to 3xP3-FP.

78

Figure 3.3 | Putative mechanism for IGC to generate gene drive docking lines, dRPLT shown. In the pre- meiotic germline of {+/IGCdRPLTpB; +/VasCas9} individuals, Cas9 expresses via the Vasa2 promoter from chr.2L [A], and gRNAs express from IGCdRPLTpB elsewhere in the genome [B] shown here on chr. 3. The endogenous RpL11-Rpt1 locus [C] is targeted by the Cas9/gRNA complex which creates dsDNA breaks [D] stimulating the cell to repair by HDR. While many HDR events repair from the homolog (not shown), some repair via HDR-based interlocus gene conversion [E](magenta arrows) using the homology donor template [F] provided on the IGCdRPLTpB transgene. This results in gene conversion of the gene drive docking site dRPLT into RpL11-Rpt1 [G], thereby integrating all recoding, 3’UTRs, attP sequences and fluorescent markers within the homology donor. Once integrated, the dRPLT [G] EYFP fluorescent marker becomes unlinked from the act-DsRed on IGCdRPLTpB [B] enabling visualization of knockin in the progeny following germline meiotic segregation of the chromosomes in the parental adult germline.

3.5 A. Generating dRPLT by CrIGCkid. Crossing females from the four IGCdRPLTpB families to VasCas9 males yielded F1 offspring with the {+/IGCdRPLTpB ; +/VasCas9} genotype capable of CrIGCkid in the germline.

Outcrossing these CrIGCkid male offspring to wild type females yielded six larvae from one family with

3xP3-EYFP unlinked from act-DsRed, suggesting IGC-mediated knockin of dRPLT had occurred. PCR confirmed knockin, and the dRPLT docking line was founded (Section 3.6A). We performed this replicate five generations (G5) after establishment of IGCdRPLTpB transgenics, however subsequent replicates at

G11 failed to reproduce CrIGCkid, preventing further characterization of the CrIGCkid of dRPLT.

79

3.5 B. Generating dRPLTu by CrIGCkid. Before initiating dRPLTu knockin by CrIGCkid, we determined the

IGCdRPLTupB insertion sites of the five families (Table 3.2) to characterize the influence of donor locus on

CrIGCkid efficiency.

Table 3.2 | Summary of IGCdRPLTupB insertion sites

Summary of IGCdRPLTupB insertion sites for five iso-female families used in CrIGCkid characterization assays. Line name shown at left, number of transgene insertions in middle column, and chromosome arm and band shown at right. If inverse PCR confirmed insertion sequence, but did not align to the annotated genome it is denoted as “Unknown”, and is likely inserted into repetitive regions of X chromosome.

For CrIGCkid of dRPLTu, we crossed G5 {+/IGCdRPLTupB ; +/VasCas9} males from each family to wt females and screened for knockin larvae. Two families, IGCdRPLTupB4 and IGCdRPLTupB5, produced larvae suggestive of CrIGCkid by fluorescence. From these individuals we established the dRPLTu docking line

(Section 3.6B) and validated dRPLTu knockin into RpL11-Rpt1 by PCR. PCR confirming insertion of each end of dRPLTu into RpL11-Rpt1 was performed with multiple primer sets, with one side shown below

(Figure 3.4). These findings demonstrate that CrIGCkid is a reproducible process for knockin into otherwise genetically intractable genes, and can be used to generate docking sites in haplolethal genes – a critical step for making evolutionarily stable population replacement gene drives in many organisms.

80

A

B

Figure 3.4 | PCR confirmation of dRPLTu knockin into RpL11-Rpt1 following IGC. [A] Map of PCR products generated with primer binding sites (arrows) and predicted fragment sizes shown. Wild type (unintegrated) RpL11-Rpt1 locus on top; PCR 1 (pale blue, 1.5 kB) specific to this locus with primers binding specifically to endogenous 3’ coding sequences removed during recoding (red bars). Middle, dRPLTu integrated into RpL11- Rpt1, PCR 2 product (navy, 2.2 kB) specific to attP and to genomic sequence beyond homology arm. PCR 3 (Carolina blue, 0.9 kB) specific to IGCdRPLTupB (bottom) encompasses a portion of the gRNA-encoding cassette and act-DsRed reporter. [B] PCR results on 7-day old adult mosquitoes. Three homozygotes, eight heterozygotes and three wild type siblings shown with donor transgene, wild type, and water controls.

In these experiments we observed that families of males undergoing CrIGCkid demonstrated decreased fertility (Figure 3.6A, Percent fertility) suggesting CrIGCkid may be genetically burdensome. The first three broods from ‘CrIGCkiding’ ({+/IGCdRPLTupB4; +/VasCas9}) males were wholly infertile followed by fluorescent identification of single CrIGCkid-positive larvae in each of two subsequent broods (Figure

3.6A, Percent fertility). PCR analysis on the final brood revealed that 77% (n = 27/35) of larvae co- inheriting IGCdRPLTupB4 also harbored a hidden dRPLTu site not visible due to fluorescence overlap (Figure

3.6A, Percent ‘hidden’ dRPLTu larvae). CrIGCkid of dRPLTu from IGCdRPLTupB5 could not be formally

pB quantified in this replicate (Figure 3.6A, nXdRPLTu), but was confirmed by PCR. Both lines carry IGCdRPLTu

81 on chromosome 3L, 15.8 Mb and 9.7Mb away from RpL11-Rpt1 respectively, suggesting CrIGCkid occurs in a proximity dependent manner in concordance with IGC observed in other organisms (Reviewed in22).

Observation of more ‘hidden’ than visible dRPLTu larvae suggests CrIGCkid occurs more frequently intrachromosomally than with the homolog. To further characterize CrIGCkid in these males, we performed further replicates at G12. In G12 no CrIGCkid was observed from IGCdRPLTupB4-containing males, and none were identified by PCR (n=68) suggesting either CrIGCkid is a stochastic process occurring rarely, or that IGCdRPLTupB4 function had degraded over time. Meanwhile IGCdRPLTupB5-containing males sired many dRPLTu larvae in each of three broods (Figure 3.6A, Percent visibly identifiable dRPLTu larvae). PCR for hidden dRPLTu revealed its presence in 23% (n = 33/141) of IGCdRPLTupB5-containing individuals. Estimating the total hidden dRPLTu larvae in the brood (537 larvae x 0.23 = 124), reveals a similar number to those observed directly by fluorescence (n=115) (Figure 3.6A, Percent projected

‘hidden’ dRPLTu larvae). The close linkage of the IGCdRPLupB5 donor to the RpL11-Rpt1 target (9.7 Mb) therefore suggests it was capable of templating CrIGCkid interchromosomally, capable of CrIGCkid followed by homing of dRPLTu into the homolog, or is separated from RpL11-Rpt1 by a recombination hotspot.

The absence of CrIGCkid from each brood suggests it may be stochastic occurring in a rare subset of male germlines. To explore this, at G16 we crossed cohorts of ten males undergoing CrIGCkid from the

IGCdRPLTupB4 and IGCdRPLTupB5 families to 200 wt females. No visibly identifiable or hidden dRPLTu sites were identified among all larvae scored (n > 5,000) from five cohorts using IGCdRPLTupB4 and four cohorts using IGCdRPLTupB5. This suggests that either CrIGCkid is sufficiently rare that it was unobserved, or – because these experiments occurred at IGCdRPLTupB G16 while VasCas9 function remained unchanged over many years – that the gRNA-expressing transgenes facilitating knockin may be losing function over time. Further characterization of CrIGCkid for gene knockin is beyond the scope of this work at this time.

82

A Characterizing CrIGCkid in males Figure 3.6 | Characterization of CrIGCkid to create dRPLTu in [A] males and [B] females. Black and grey denote results from a replicate at G5 and blue denotes a second replicate at G12. Percent fertility: Each point is a single egg-lay (one brood) from a single cage of {+/IGCdRPLTupB; +/VasCas9} males [A] or females [B], outcrossed to reciprocal wt mates. Open points are broods without egg laid and filled grey points are broods with only infertile eggs (G5 only). Closed points are percent fertility if >1 larvae hatched. Number in parentheses is number eggs laid. Brood data was considered invalid and discarded if all larvae could not be scored living. Percent visible dRPLTu larvae: Percent larvae with the visible dRPLTu phenotype of {3xP3-CFP- positive; act-DsRed-negative} among total, pie charts summed across all broods. Total summed larvae hatched and scored as n=# below, number of broods summed in parentheses, ‘n=’ within pie chart is number of visible dRPLTu larvae, ‘n/a’ is no data collected. nXdRPLTu is total visible dRPLTu larvae identified in invalid broods with total B Characterizing CrIGCkid in females broods in parentheses. Percent ‘hidden’ dRPLTu larvae: Percent of IGCdRPLTupB- positive larvae PCR-verified to have co- inherited a dRPLTu integration. Number of broods analyzed in parentheses. Results from G5 or G12 in grey or blue respectively. ‘n/a’ denotes no data collected. Percent projected ‘hidden’ dRPLTu larvae. For replicates with visible dRPLTu larvae, and for which ‘hidden’ dRPLTu larvae were PCR- screened (G5 IGCdRPLTupB4 and G12 IGCdRPLTupB5 grey outlined or blue outlined pie chart respectively), percent calculated ‘hidden’ dRPLTu (77% and 23% respectively) was multiplied by the total number of IGCdRPLTupB-positive larvae observed (36 and 537 respectively) to approximate the number of ‘hidden’ dPRLTu larvae present in the total population (mauve, Projected {+/IGCdRPLTupB ; +/dRPLTu}).

83

3.5 C. CrIGCkid does not occur in females, and infertility phenotypes lessen over time. To determine if

CrIGCkid could also occur in females, we analyzed the offspring of G5 {+/IGCdRPLTupb ; +/VasCas9} females from the five families. We discovered most families were highly infertile, with the exception of line

IGCdRPLTupB3, with IGCdRPLTupB4 females laying no eggs across all four broods (Figure 3.6B, Percent fertility). Among the larvae produced by other families, none displayed the fluorescence pattern indicative of CrIGCkid (Figure 3.6B, Percent visible dRPLTu larvae). Later replicates at G12 on

IGCdRPLTupB4- and IGCdRPLTupB5-containing females confirmed no knockin larvae by florescence nor PCR

(n = 0/45 and n = 0/62; IGCdRPLTupB4 and IGCdRPLTupB5 respectively), however overall fertility had increased (p<0.05 and ns respectively, Mann-Whitney) (Figure 3.6B, Percent fertility, blue points).

Together these results suggest CrIGCkid does not occur in females, at least not in a manner which permits larval survival, and increasing fertility further points towards attenuated IGCdRPLTupB transgene function.

3.6 Characterization of dRPLT and dRPLTu docking lines

3.6 A. dRPLT is homozygous lethal, experiences proteasome dysfunction, and is likely a Rpt1 knockout.

Omission of a 3’UTR during Rpt1 recoding should cause dRPLT recessive lethality due to its requirement for sub-cellular localization23. Indeed, intercrossing {+/dRPLT} yielded no homozygous offspring. To monitor lethality, we crossed {+/dRPLT} to a second transgenic within dRPLT, mNosGDdRPLT (discussed in section 3.7B, maintains all properties of dRPLT), harboring a 3xP3-DsRed marker enabling identification of homozygotes by {3xP3-EYFP/3xP3-DsRed} fluorescence. The crosses yielded the expected frequencies of larval genotypes (Figure 3.7A | p > 0.05 Chi-squared test), however all {dRPLT/mNosGDdRPLT} individuals died within 24 hours, phenocopying Drosophila Δrpt124. Western blots confirmed proteasome dysfunction by significant poly-ubiquitin accumulation23,25,26 in homozygotes compared to controls (Figure 3.7B).

Together these findings confirm knockout of Rpt1 and knockin of dRPLT.

84

Figure 3.7 | Fitness characterization of dRPLT. [A] Genotypes of one-day-old offspring from {+/dRPLT} x {+/mNosGDdRPLTu} cross, scored by 3xP3-EYFP or 3xP3-DsReD fluorescence. Nonsignificantly different from expected Mendelian ratios of 1:1:1:1 (Chi-squared test p > 0.05). [B] Western blot for proteasome dysfunction in wild type {+/+}, heterozygous {+/dRPLT} or {+/ mNosGDdRPLTu}, and homozygous {dRPLT/mNosGDdRPLTu} siblings. Supernatant and pellet shown, probing for poly-ubiquitin. 30 larvae per sample, with murine tissue culture positive control.

3.6 B. dRPLTu homozygotes are viable but display Minute-like fertility phenotypes. In contrast to dRPLT, dRPLTu homozygous larvae survive to adulthood (Figure 3.4). Western blot has not yet confirmed proteasome activity, but homozygous viability suggests the replacement 3’UTR from Rpt3 is sufficient to maintain most proteasome function (Table 3.1). However we observed that many homozygotes pupated later than siblings, and either failed to eclose from the pupal case (n = 7/34) or drowned (n= 21/34), with only 17% surviving (n = 6/34). These observations are consistent with Drosophila Minute phenotypes characterized by delayed pupation, pupal death, and failure to eclose27.

To further explore Minute phenotypes in dRPLT and dRPLTu, we performed larval competition assays between docking site homozygotes, heterozygotes and wt siblings. In all experiments wt siblings pupated significantly earlier than docking site siblings when reared with together (Figure 3.8A | p < 0.001, and p <

0.001 , Wilcoxon rank), suggesting both lines experience the classic pupation delay observed in other

Minutes27. However, while fewer {+/dRPLT} larvae survived to pupation than expected (Figure 3.8A | p <

0.01, Chi-square test), dRPLTu larvae had normal survival rates, suggesting that heterozygous proteasome dysfunction – rather than ribosome dysfunction – makes dRPLT larvae more susceptible to larval competition.

85

Figure 3.8 | dRPLT and dRPLTu individuals display many hallmark Minute phenotypes. [A] dRPLT and dRPLTu individuals form pupae later than wild type siblings. Graphs show sum of three biological replicates, each with 50 wt or 50 {+/dRPLT} siblings, or 25 wt, 50 {+/dRPLTu}, and 25 {dPRLTu/dRPLTu} siblings. Date of pupae formation was scored each day by fluorescence until all pupated, and the percent pupae formed each day of total calculated and plotted. Mean and SD shown, significance calculated as inverse survival curve by Wilcoxon rank (p < 0.001, p < 0.001). Total number pupated by genotype: dRPLT differs significantly from expected 50/50 ratio (Chi-Squared, p < 0.005). dRPLTu does not differ significantly from 25/75 fluorescence ratio (Chi-squared p > 0. 01). [B] Midgut (white arrow) and ovaries (grey arrows) of {dRPLTu/dRPLTu} 10-day old female mated with males ad libitum, 48 h after second blood feed. Blood-meal is fully digested but ovaries are undeveloped. [C] Closer examination of {dRPLTu/dPRLTu} 48 h after second blood-meal reveals slight but stunted follicular development. [D] Differences in body size of wild type female {+/+} adjacent to {dPRLTu/dPRPLTu} sister.

These lines showed other defects generally associated with Minutes. While attempting to establish a stable {dRPLTu/dRPLTu} line, we observed that homozygous females could mate and blood-feed, but never laid eggs after many (n>3) blood meals. Dissection 48 hours post-blood feed revealed small and undeveloped ovaries when they should be voluminous (Figure 3.8B, example normal 48h post-blood feed ovaries shown in Figure 2.5A). Further examination revealed that follicle growth was stunted (Figure

86

3.8C), displaying a string-of-pearls-like phenotype similar to those observed in Drosophila Minutes28 caused by insufficient protein synthesis during the ribosome-intense process of oogenesis28. While this phenotype requires further characterization, it is likely due to insufficient ribosome activity, and will prevent establishment of true-breeding {dRPLTu/dRPLTu} lines. Finally, a clear macroscopic difference in adult body size and length was observed (Figure 3.8D), further supporting the hypothesis that dRPLTu homozygotes are Minutes. While Western blots to determine if homozygotes experience decreased RpL11 expression were inconclusive, collectively the phenotypes observed strongly point towards a ribosome deficiency due to some aspect of recoding or misregulation by the RpL32 replacement 3’UTR.

3.7 Gene drives in dRPLT and dRPLTu may be lethal

3.7 A. Inserting B2GD into dRPLT or dRPLTu results in death of all transgenic individuals. With creation of the drive docking lines complete, we began transgenic insertion of gene drives. We constructed a spermatogenic specific gene drive plasmid using the β2-tubulin promoter29,30 for Cas9 expression and the eight gRNAs proven effective to facilitate CrIGCkid, with further 3xP3-DsRed fluorescence and attB sites for insertion onto the docking sites. The plasmid is termed B2GD (β2-enabled Gene drive), and the resulting transgenic is B2GDdRPLT (β2-tubulin enabled Gene drive in the dRPLT site)(Figure 3.9).

Injections into >7,000 {+/dRPLT} mixed embryos gave four B2GDdRPLT larvae, but all died in development.

Such low transgenesis rates and death of all nascent larvae is rare, raising suspicions of a biological culprit.

To determine if the homozygous viable dRPLTu site could tolerate transgenesis, we injected 1009 embryos with B2GD and recovered three B2GDdRPLTu nascent transgenics, however followed again by death. Despite low numbers, failure of nascent transgenics to reach adulthood is rare in our hands (Table 3.3), suggesting that gene drive transgenesis into both docking sites is incompatible with life.

87

Figure 3.9 | Gene drive and distal gene drive transgene maps developed (Appendix A). Orientations of inserts are correct, size not to scale. Gene drives (GDs) are transgenes for which insertion into the docking lines was attempted (shown in dRPLT). Distal Gene drives (dGDs) are in random loci with piggyBac terminal repeats shown (forest green). RpL11(mint) and Rpt1(navy) shown, recoding (red bars) and RpL11 replacement 3’UTR (violet) consistent with prior images. Cas9 (salmon), recombined phiC31 attR and attL sequences (black), and gRNAs are shown in blocks of four (red). Promoters for Cas9 begin with ‘p’ and color-coded by gene (mNos (rose), Vasa (indigo), β2 (lime), ZPG 2 kB (rust), ZPG 1 kB (copper)). LacZ multiple cloning site for anti-malarial cargoes in blue. 3xP3-DsRed selectable markers shown in pale blue and red-orange, and act-DsRed selectable markers shown in grey and red-orange. Vasa protein fusion is shown in pale indigo with the XTEN protein linker shown in aqua.

Table 3.3 | Number embryos injected, transgenics recovered and survived to adulthood

Survival rates of select phiC31 transgenesis developed during the course of this work. Transgene names at left are reported for clarity but not all discussed here.

88

3.7 B. Lethality of gene drives in docking sites may be due to aberrant Cas9 overexpression. We hypothesize that RpL11 or Rpt1 regulatory elements may be aberrantly affecting Cas9 expression when in the docking site. In our system, Cas9 expression in any tissue incapable of HDR is expected to be lethal, however direct assay of Cas9 levels on nascent transgenics is impossible given low transgenesis rates and early death of B2GDdRPLT and B2GDdRPLTu individuals, necessitating analysis by indirect methods.

To determine if dRPLT can tolerate insertion of drive-like transgenes independent of confounding Cas9 expression, we inserted a modified gene drive capable of gRNA expression but incapable of Cas9 translation. Enabled by a truncated (1,439 bp) Nanos germline promoter (AGA006098 this Nanos gene drive is termed mNosGD, (Figure 3.9, previously referenced in Section 3.6A) and expresses the same eight gRNAs used in B2GD (Section 3.7A) and for CrIGCkid (Section 3.5B). It was transgenically inserted into dRPLT, sequence verified to contain no polymorphisms, confirmed to display phenotypes consistent with transgenic gRNA expression (Figure 3.11), and has been maintained as heterozygotes for 44+ generations.

Thus demonstrating dRPLT can tolerate insertion of gene drive-like gRNA-encoding transgenes incapable of Cas9 protein expression.

Figure 3.10 | Western blot analysis for Cas9 expression in drive lines. 7 5-10 day old male lower abdomens and female ovaries dissected and probed for Cas9 (160 kDA) or VasaFusionCas9 225 kDA). Controls: actin loading ctl. (42 kDA), Wild type G3 (black) negative ctl. and VasCas9 (red) positive ctl. Drives: mNosGDdRPLT (rose) gene drive at G40, and distal drives B2dGDpB (lime), ZPG2dGDpB (rust), and VFdGDpB (indigo), ZPG1dGDpB not shown.

89

At G6 we confirmed mNosGDdRPLT females express germline Cas9 mRNA but no protein by Western (not shown), however Western blot at G40 either hints at possible expression or a blot artifact (Figure 3.10).

While different versions of the Nanos promoter can enable drive31, the absence of drive-enabling Cas9 expression in mNosGDdRPLT was confirmed as no biased inheritance observed in G1 or G2 (Figure 3.11).

Subsequent replicates at approximately G40 confirmed the continued absence of biased inheritance, suggesting the Western artifact not likely Cas9 expression in this line (Figure 3.10).

Figure 3.11 | No biased inheritance of mNosGDdRPLT. Scatter plot of transgene frequency in progeny of adults tested for drive. Female and male mosquitoes heterozygous for the mNosGDdRPLT allele, outcrossed to wild type, were scored for biased inheritance of mNosGDdRPLT by fluorescence in the progeny. Male {+/mNosGDdRPLT) is multiple broods from the single founder male. Female {+/mNosGDdRPLT} were analyzed for drive in G2 and points single broods from isolated females. A replicate at approximately G40 is a bulk brood from 6 females (n = 593 larvae scored). The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown.

To determine if B2GD CRISPR-expressing cassettes are lethal independent of insertion in the docking site, we inserted β2-Cas9 and gRNA RpL11-Rpt1-targeting expression cassettes randomly throughout the genome by piggyBac transgenesis32. If non-lethal, these transgenics should be easily established, and if capable of drive they should facilitate biased inheritance of dRPLT or dRPLTu when co-inherited in the same individual. Much like a daisy drive33, we term these distal gene drives (dGD) in this work, with the one enabled by β2-Cas9 expression termed B2dGDpB (Figure 3.9). Injecting B2dGDpB into >700 embryos yielded a surplus of nascent transgenic larvae (>90 larvae), expressing polymorphic fluorescence

90 phenotypes indicative of variable insertion sites, and Cas9 expression was verified by Western (Figure

3.10). This suggests that transgenic expression of B2GD CRISPR-components targeting RpL11-Rpt1 are nonlethal when inserted within a variety of genomic loci, but are lethal when directly inserted into dRPLT or dRPLTu.

To test the ability of B2dGD to bias inheritance of dRPLTu, we scored {+/B2dGDpBmix;+/dRPLTu} males and females from bulk and single-family populations for drive, but none revealed biased inheritance (Figure

3.12). The β2 promoter has been previously demonstrated to enable spermatogenic endonuclease expression30, but its expression begins in early spermatocytes34 possibly too late after meiosis I to stimulate interchromosomal homing, however IFA on our lines will follow to confirm this. Importantly confirmation of Cas9 expression suggests that mosquito male germlines can tolerate some RpL11-Rpt1- targeting CRISPR expression, likely in later stages of spermatogenesis.

Figure 3.12 | B2dGDpB does not facilitating drive of dRPLTu in males. Female and male mosquitoes heterozygous for the B2dGDpB and dRPLTu transgenes {+/B2dGDpB;+/dRPLTu} were outcrossed to wild type and analyzed for the biased inheritance of dRPLTu in the progeny. Scatter plot of the transgene frequency observed in the progeny with each filled dot representing the frequency of a single brood of either a single female founder ({+/B2dGDpB;+/dRPLTu} ♀) or a cohort of brothers from an iso-female line ({+/B2dGDpB;+/dRPLTu}♂). Closed triangles show three broods from the same cage of iso-female brothers. Closed triangle with an astersk denotes a single brood demonstrating statistically significant drive (60.2% allele frequency, p < 0.01, O. vs E Two-tailed), but subsequent broods failed to recapitulate. Open diamonds show drive frequencies from broods sired by mixed population adult males (non-iso-female) {+/B2dGDpBmix;+/dRPLTu}. The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown.

91

Together these data are enigmatic; B2dGDpB transgenics expresses CRISPR and are viable from many genomic positions but do not enable drive of dRPLTu, while their functional clone, B2GD, kills all nascent transgenic individuals when inserted into dRPLT or dRPLTu, and further experiments confirm dRPLT can tolerate insertion of non-Cas9-translating drive-like transgenes. Taken together these suggests that inserting transgenes capable of any Cas9 protein expression into the docking lines causes lethality when in the presence of ubiquitously expressed transgenic gRNAs, supporting the hypothesis that regulatory sequences controlling the locus may be able to cause aberrant Cas9 expression and associated lethality.

With this in mind, we set out to identify and characterize promoter candidates capable of drive-enabling

Cas9 expression, independent of the confounding effects associated with insertion into the docking lines, using the distal gene drive (dGD) system introduced earlier (Section 3.7B).

3.8 Developing drive-enabling germline-Cas9 promoters independent of docking sites.

Due to lethality associated with inserting Cas9-expressing transgenes into dRPLT and dRPLTu, we set out to characterize germline promoters for Cas9 expression from more neutral loci. These experiments could facilitate initial characterization of the DRA-inhibiting properties of these drive designs, and would provide a starting point for subsequent fine-tuning of Cas9 expression within dRPLT or dRPLTu.

3.8 A. Vasa2 transgenic promoter is too leaky for drive. The Vasa2 promoter has been used for Cas9 expression in successful gene drives in both An. gambiae35 and An. stephensi17. However this blood-meal inducible transgenic promoter is leaky into adjacent tissues36 and causes maternal deposition into the developing embryo37. While not an issue for other drive designs, Cas9 expression in any tissues incapable of HDR would cause significant fitness defects or death in our system, making Vasa2 a poor candidate for our drive designs. We verified that Vasa2-Cas9 embryo deposition is lethal by injecting a Vasa2-Cas9 gene drive construct (Figure 3.9, VasGD) into embryos, which recovered only 1.7% survivors (n=17/1010), the lowest percent survival of embryo microinjections to date in our hands. To verify Cas9 expression from

92 the VasGD transgenesis plasmid was inducing lethality, we co-injected a dsCas9-expressing RNAi plasmid to silence Cas9, which resulted in a significant survival increase (9.5 % survival, n = 105 survivors /1108 injected, p < 0.0001, Fisher’s exact, Two-sided). These data therefore confirm that a Vasa2-enabled gene drive would likely be embryonic lethal in our drive design.

We next attempted to rescue the non-translating Cas9 in mNosGDdRPLT with the Vasa2-Cas9 expressed by the VasCas9 line by crossing these transgenic lines together. In the event of drive, individuals with both transgenes ({+/mNosGDdRPLT ; +/Vas-Cas9}) should show biased inheritance of mNosGDdRPLT. In a first experiment using G20 of mNosGDdRPLT, we discovered that driving females died within 24 hours post- bloodfeeding (n = 6), while control sisters survived and laid normal broods (Figure 3.13, n = 9 females,

Chi-square test), suggesting Vasa2-Cas9 is leaky following blood feeding and causes lethal RpL11-Rpt1 mutagenesis. Interestingly, subsequent replicates at approximately G40 produced viable females that however showed complete infertility (n=10) compared to control sisters which laid normal broods (n=12)

(Figure 3.13). The lessening of the lethal phenotype over time suggests that transgenic expression weakens with subsequent generations, though it is still sufficient to cause complete infertility. We cannot test this hypothesis directly due to absence of RNA samples from earlier generations.

We then scored for biased inheritance of mNosGDdRPLT from {+/mNosGDdRPLT;+/Vas-Cas9} driving males and nondriving control brothers. From bulk cage outcrosses we discovered that control males sired larvae with expected transgene frequencies (n = 1321 total larvae, 49.0% transgene frequency) and normal fertility levels (89.9% fertility), while driving males sired no larvae despite 482 eggs laid (Figure 3.13).

These results confirm that Vasa2-Cas9 drives are incompatible with our system design, suggesting more fine-tuned promoters need to be developed.

93

Figure 3.13 | Rescuing mNosGDdRPLT Cas9 function with Vasa-Cas9 results in death or infertility of all females and infertility in males. Female and male mosquitoes heterozygous for the mNosGDdRPLT and Vasa-Cas9 transgenes ({+/mNosGDdRPLT;+/Vas-Cas9} genotype) were outcrossed to wild type and analyzed for the biased inheritance of mNosGDdRPLT in the progeny compared to sibling controls {+/+;+/Vas-Cas9}. Both transgenes are unlinked but share a common 3xP3-DsRed selectable marker, therefore drive would be indicated by >75% 3xP3- DsRed-positive larvae from {+/mNosGDdRPLT;+/VasCas9} parents, and >50% 3xP3-DsRed-positive larvae from {+/+;+/VasCas9} control siblings. Females were analyzed at approximately G20 and G40 of mNosGDdRPLT . Scatter plot of the transgene frequency observed in progeny with each filled dot representing the transgene frequency of a single brood from a single female (♀) or from a cohort of 25 brothers (♂). Control {+/+;+/Vas-Cas9} points shown in grey, and driving genotype {+/mNosGDdRPLT;+/Vas-Cas9} in pink. Females dead following blood-feeding are marked by circled x’s (~G20, n=6), and open circles denote infertility of the brood for both males and females. Male infertility was characterized by complete infertility despite n=482 eggs laid, while female infertility (Gen. 40, n=10) manifested as no laid eggs (n=9) with a single female laying infertile eggs (n=1). The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown for groups with > 1 data point >0.

In a subsequent brood from this experiment driving ({+/mNosGDdRPLT;+/Vas-Cas9}) males were incompletely infertile, siring 19 larvae from n>600 infertile eggs. Interestingly, the larvae were exclusively nontransgenic, inheriting neither CRISPR-expressing transgene VasCas9 nor mNosGDdRPLT. These larvae were not the result of contamination, suggesting a rare event had occurred likely in the germline of a single {+/mNosGDdRPLT; +/Vas-Cas9} male. Together these suggest a Vasa2-Cas9 gene drive would cause significant male infertility, and rare viable sperm do not inherit either CRISPR-expressing transgenes.

3.8 B. A Vasa protein fusion to fine-tune Cas9 expression. To overcome this issue, we attempted to precisely express Cas9 only in the primordial germ cells using the Vasa2 promoter to express a fusion between the full-length Vasa protein and Cas9 by an Xten linker38, termed VFdGDpB (Vasa Fusion distal

94 gene drive, Figure 3.9). The rational for this design was based on knowledge in other Dipterans where

Vasa localization to the PGC nuage is dependent on its physical interactions with other proteins such as

Oscar39. However IFAs could not verify if the Vasa-Cas9 fusion protein was correctly localized to the germ cells, despite Western confirming its expression. Driving {+/VFdGDpB; +/dRPLTu} males demonstrated no drive of dRPLTu across two broods (Figure 3.14, n = 149/294, 50.7% ; n = 221/411 53.8%; both ns, Two- tailed Binomial) in concordance with no VasaFusionCas9 in male germlines (Figure 3.10). However females also did not display drive (Figure 3.14, n = 398/853, 46.7%, ns Two-tailed binomial) despite

VasaFusionCas9 expression (Figure 3.10), and despite having normal levels of fertility. Therefore we set out to develop alternative germline promoters with lower baseline expression which may be capable of drive.

Figure 3.14 | A Vasa fusion distal drive, VFdGDpB is fertile but fails to facilitate drive of dRPLTu. Female and male mosquitoes heterozygous for the VFdGDpB and dRPLTu transgenes ({+/VFdGDpB; +/dRPLTu} genotype) were outcrossed to wild type and analyzed for the biased inheritance of dRPLTu in the progeny. Scatter plot of the percent progeny inheriting a copy of dRPLTu. Two consecutive broods shown from a single founder male, and a single brood shown from a cohort of six females. The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown for groups with >1 data point.

3.8 C. Two versions of the ZPG promoter for gene drive Cas9 expression. To develop alternate germline promoters for Cas9 drive expression, we developed distal gene drives using the 5’ and 3’ regulatory sequences of the Zero population growth (ZPG) Innexin2 gene (AGAP006241). A first transgenic construct using 2 kB of the 5’ promoter sequence and 1 kB of 3’UTR sequence amplified from the genome was termed ZPG2dGDpB (for ZPG 2 kB promoter distal Gene drive). However this promoter failed to achieve

95

Cas9 expression in adult germlines (Figure 3.10), and did not facilitate dRPLTu drive in bulk female experiments nor from iso-female families (Figure 3.15, p>0.05, Chi-squared). Males from the families in which females demonstrated slight but insignificant drive were either infertile, or showed a bias against dRPLTu inheritance. Taken together, these suggest that the ZPG2 does not facilitate drive-enabling Cas9 expression.

Figure 3.15 | A ZPG2 distal drive, ZPG2dGDpB fails to facilitate drive of dRPLTu. Female and male mosquitoes heterozygous for the ZPG2dGDpB and dRPLTu transgenes {+/ ZPG2dGDpB;+/dRPLTu} were outcrossed to wild type and analyzed for the biased inheritance of dRPLTu in the progeny. Scatter plot of the dRPLTu transgene frequency observed in the progeny with each filled point representing the frequency of a single brood of either a single female founder ({+/ZPG2dGDpB;+/dRPLTu} ♀) or a cohort of brothers from an iso-female line ({+/ZPG2dGDpB;+/dRPLTu}♂) and were chosen due to >55% dRPLTu frequency in female founders. Closed dots denote a single brood analyzed from a single family. Closed up triangles, down triangles, and squares specify multiple broods analyzed from three distinct families. Down triangle brood with 66.7% dRPLTu frequency only representative of 12 larvae, and biased inheritance not recapitulated in subsequent broods. Open triangles on 0 denote infertile broods from respective families. Open diamonds denote large bulk experiments from a mixed insertion population of {+/ZPG2dGDpBmix;+/dRPLTu} with ZPG2dGDpB at varying genomic loci. The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown.

More recently, an alternative ZPG promoter was published capable of robust and tight Cas9 expression for gene drive4,31. This ZPG promoter was ~1 kB long, while the 3’UTR used for termination was nearly identical to that used in ZPG2dGDpB.. Therefore we designed a distal drive using these regulatory sequences ordered as clonal IDT gBlocks4 and developed a transgenic line termed ZPG1dGDpB (ZPG 1 kB promoter distal gene drive). In bulk outcrosses of 10 driving {+/ZPG2dGDpB;+/dRPLTu} females, all broods

(n=4) laid no eggs despite having undergone oogenesis, with the exception of one female who laid an infertile egg batch. In light of similar observations in G40 female {+/mNosGDdRPLT;+/Vas-Cas9} drive rescue experiments (Figure 3.13), these results suggest that RpL11-Rpt1-targeting, drive-enabling Cas9-

96 expression in the germline causes female infertility, consistent with Minute-like infertility phenotypes associated with ribosome depletion27,28,40. In the absence of fine-tuning of Cas9 expression, drive designs targeting haplolethal genes in the primordial germline may not be viable in females, especially given significant female-infertility phenotypes known to occur in female Minute mutants.

Figure 3.16 | A ZPG1 distal drive, ZPG1dGDpB causes complete female infertility, and incomplete male infertility which decreases in subsequent egg-lays. Female and male mosquitoes heterozygous for the ZPG1dGDpB and dRPLTu transgenes {+/ZPG2dGDpB;+/dRPLTu} were outcrossed to wild type and analyzed for the biased inheritance of dRPLTu in the progeny shown as scatter plot. Open points denote complete infertility of the brood with each filled point representing the frequency of dRPLTu-positive larvae in a single brood. A cohort of (n=10) female {+/ZPG2dGDpB;+/dRPLTu} were analyzed in bulk over the course of four consecutive egg-lays. Eggs laid were n=33, n=152, and n=15 for the first three egg-lays respectively, all infertile. A cohort of 35 mixed insertion site male {+/ ZPG2dGDpBmix;+/dRPLTu}♂ were analyzed over three consecutive egg-lays. Closed dots denote the frequency of dRPLTu-positive larvae in a fertile brood, all broods were highly infertile, n >1000 eggs laid. Brood 1 contained a single dRPLTu-positive larvae among many wt (n= 133), three of which inherited the ZPG2dGDpB CRISPR-encoding transgene (not plotted). Brood 2 was wholly infertile. Brood 3 maintained normal inheritance ratios of dRPLTu (plotted) and ZPG2dGDpB (not plotted). The 50% expected Mendelian ratio is marked by a dotted line, with Mean and (± SEM) shown for groups with >1 data point >0.

To analyze distal drive properties of dRPLTu by ZPG1dGDpB in males, we scored for dRPLTLu biased inheritance in the offspring of 35 driving {+/ZPG2dGDpBmix; +/dRPLTu} males. Brood 1 was highly infertile

(n=134 larvae, >1000 eggs laid), however among hatched larvae, only one inherited dRPLTu and three inherited ZPG2dGDpB, suggesting a bias against inheritance of dRPLTu or the CRISPR-encoding parental distal drive; ‘anti-drive’ (Figure 3.16). These findings phenocopy those observed in the offspring of a single brood of {+/mNosGDdRPLT;+/Vas-Cas9} males (discussed in Section 3.8A), in which males capable of drive were largely infertile, but when fertile, they only sired larvae without CRISPR-encoding transgenes. Brood

97

2 was completely infertile despite >1000 eggs laid, while Brood 3 was similarly fertile to Brood 1, but was not characterized by negative biases in transgene frequency. The sequential nature of these brood phenotypes therefore possibly suggests occurrence of a time-dependent anti-drive or fertility effect.

Sequence analysis on the fertile larvae sired in these broods revealed no evidence of DRA mutations under any of the eight gRNA binding sites, suggesting those larvae completely escaped CRISPR mutagenesis.

DISCUSSION

Our work makes important advancements towards the development of evolutionarily stable gene drives for population replacement in the important An. gambiae malaria vector, and suggests that developing such homing-based drives in haplolethal genes still has many significant hurdles to overcome.

CrIGCkid is a novel knockin technology for insertion into difficult-to-manipulate loci in anophelines; it is robust, reproducible, proximity-dependent, and occurs exclusively in males.

To make gene drive docking lines, we develop a novel method in the anopheline GM toolkit for reproducible CRISPR-mediated gene knockin, CrIGCkid, for insertion into haplolethal genes – a critical step for making evolutionarily stable gene drives for population replacement. CrIGCkid of two distinct donor templates, dRPLT and dRPLTu, over multiple replicates demonstrates that the technique is predictable and reproducible (Section 3.5B,C), and most likely occurs in a linked manner between the target and donor locus. Templating can occur from either side of the target site, with identical donors in distant loci templating CrIGCkid variably, consistent with IGC in other organisms19,22. Together these confirm the flexibility of this system, but provide general design guidelines when implementing CrIGCkid from other loci and in other organisms.

98

Within these experiments CrIGCkid is only observed in males, possibly due to the thousands of recombinatorial possibilities producible during spermatogenesis41. Dipteran male germlines can reestablish the stem cell niche from a surviving spermatogonial cell in the event of significant genetic perturbation (reviewed in 42), suggesting lethal CRISPR mutagenesis may bias for the desired knockin outcome which is resistant to further cleavage. The absence of CrIGCkid from females is likely due to the effects of CRISPR mutagenesis causing ribosome depletion which is critical to oogenesis27 explaining

Minute-like infertility phenotypes (Figure 3.8C). Despite reproducibility, CrIGCkid was still a rare event, whose identification was facilitated by Anopheles’ colossal fecundity, suggesting it may not be reproducible in species with lower reproductive outputs. Similar systems have also been successful in

Drosophila19 and are also characterized by high fecundity.

Recoding haplolethal RpL11 causes Minute-like phenotypes.

To construct our gene dive design, endogenous RpL11 required sequence recoding. Swapping the codons in this way does not alter amino acid sequence and should maintain protein function, however tRNAs vary in cellular abundance 43, necessitating codon optimization for optimal translation efficiency44. Cells require as many as 10 million ribosomes per cell45, making flawless translation of these proteins critically important. To our knowledge, Volohonski et al.12 presents the only study on anopheline codon frequency, but analyzes codons from a single haplolethal gene among a few other genes analyzed, suggesting it is not a comprehensive sampling. With the absence of an alternative, the gene recoding undertaken here was performed based on their analysis, and in the absence of formal studies on anopheline ribosomal 3’UTR regulation, a replacement for RpL11 was chosen from RpL32 based on shared functionality and stochasticity, and minimal homology. To our knowledge, while creation of the docking lines presents the first reported haplolethal gene recoding in higher eukaryotes, both however experience pupation delay, and dRPLTu homozygotes display small body size, eclosion failure, and female infertility – the classic symptoms of ribosome deficiency (Figure 3.8). Despite the absence of direct protein evidence, these

99 phenotypes convincingly suggest dRPLT and dRPLTu are anopheline Minutes, and underscore the need for research into anopheline codon optimization and ribosome 3’UTR regulation before successful development of evolutionarily stable drives can be achieved within ribosome genes. gRNA-expressing transgenes may be silencing over time.

Females undergoing CrIGCkid became less infertile with increasing age of the gRNA-expressing donor line

(Figure 3.6B), and CrIGCkid itself was never observed beyond G12. Similarly, as the mNosGDdRPLT line aged, driving {+/VasCas9; +/mNosGDdRPLT} females – which initially died 24H after blood feeding – began surviving (albeit infertile), consistent with a decrease in the function of transgenically expressed Cas9 or gRNAs (Figure 3.13). Because all experiments were carried out over four years using the same VasCas9 line, which has never otherwise varied in function, this suggests the function of the gRNA-expressing transgene was dampened over time. These findings are reminiscent of those observed in dsRNA- expressing Aedes transgenics, where the dsRNA-mediated anti-pathogen phenotype was lost over 17 generations46, therefore we postulate that the stem-loop like structure of gRNAs and dsRNAs are sufficiently alike they may be triggering similar cellular silencing mechanisms, perhaps originally evolved to silence transposon selfish-genetic elements47-50. The PIWI pathway is hypothesized to have evolved for the purpose of silencing active transposons in the germline47,51. It silences transposons in the germline52 by recognition of ‘foreign’ long noncoding RNAs, followed by post-transcriptional silencing and subsequent heterochromatinization of the genetic element, possibly taking many generations. In flies activation of one such transposon, the P-element, is sufficiently genetically burdensome that it causes complete infertility in young flies, but following PIWI activation the P-element is silenced and fertility restored in a single generation as an individual fly ages53 reminiscent of our findings in Section 3.8C. PIWI may be silencing non-coding RNA transcribing transgenes in mosquitoes, and its disruption may be required to maintain transgene function over generations. While still in need of more robust study, these clues raise serious concerns about the ability of gene drives to spread in the long term, even after

100 overcoming the hurdle of DRAs, underscoring the necessity for further research into the naturally evolved mechanisms for transposon silencing in many mosquito vector species in which gene drives are being developed.

The Goldilocks Cas9 conundrum – aberrant Cas9 expression within dRPLT and dRPLTu

All CRISPR gene drives are the same in two respects; if there is too little Cas9 in the germ line then drive cannot occur, and if Cas9 expression is localized correctly in the primordial germline, then drive can occur.

However our drive is unique in that overabundance of Cas9 in cell types incapable of HDR results in cell death by RpL11 mutagenesis.

During attempts to insert B2GD into the docking sites, all nascent transgenics died, raising suspicion of a biological culprit. Because position effect causes expression variation from an identical transgenes in different loci15, we suspected that the high transcriptional activity of the adjacent RpL11-Rpt1 genes may be affecting expression of Cas9 cassettes in the docking sites. However direct examination on nascent transgenic larvae was impossible due to transgenesis inefficiency and early death, necessitating indirect testing. To confirm docking sites could tolerate transgene insertion without lethality, we inserted the drive-like mNosGD transgene lacking Cas9 translation into dRPLT and maintained it for >40 generations

(Section 3.7B). To determine if the B2GD transgene itself was lethal, we inserted it at multiple other sites in the genome without fitness defect. While we later discovered that β2-Cas9 distal gene drives could not bias dRPLTu inheritance likely due to late Cas9 expression34 (Figure 3.10), their insertion into docking sites caused lethality, suggesting augmented lethal expression in this position. These suggest that the position effects influencing dRPLT and dRPLTu may cause aberrant Cas9 expression from integrated transgenes into intolerant tissues, emphasizing the need for developing mechanisms to fine tune CRISPR expression within these sites. While this could be achieved following introduction of insulating sequences such as gypsy insulators54, they were beyond the scope of this work at this time.

101

The Goldilocks Cas9 Conundrum – many failed promoters; the killers, the sterilizers, and the idlers

Inability to insert drives onto the docking sites necessitated assay of the system’s homing capabilities ectopically by CRISPR expression from other loci. In this manner, consistent with the Vasa2 promoter’s known drive-enabling properties17,18,37 we confirmed that Vasa-Cas9 expressed gene drives in our design would cause embryonic lethality, female infertility, and high levels of male fertility (Figure 3.13, Section

3.8A), and attempts to fine-tune expression through the use of fusion proteins proved unsuccessful

(Section 3.8B). Similarly, we confirmed the published drive-enabling ZPG1 promoter, known to be even more tightly regulated than Vas2, also causes complete female sterility, and high levels of infertility in males while not demonstrating biased inheritance (Figure 3.16). To identify a Cas9 promoter with dampened and less leaky expression we developed many distal gene drive lines which failed to show any drive phenotype (B2dGDpB, ZPG2dGDpB, VFdGDpB, Summarized in Figure 3.17), suggesting that identifying the perfect promoter for Cas9 expression may be challenging.

The Goldilocks Cas9 Conundrum – the anti-drive

Female infertility caused by ribosome depletion suggests that the highest probability of observing drive is through males. However male infertility has also been observed in some strong Drosophila Minute lines27, suggesting they too may incur sterility from RpL11 mutagenesis. In experiments carried out in driving males co-inheriting mNosGDdRPLT and VasCas9, most broods were infertile (Figure 3.13), as were most broods attempting distal drive of dRPLTu by ZPG1dGDpB. These data suggest that drive-enabling Cas9 expression in males also expressing RpL11-targeting gRNAs, results in infertility suggesting that ribosomal targeting and mutagenesis of the male primordial germline may be fundamentally incompatible with fertility without significant fine-tuning.

Taken together these data bring into focus the critically precise Cas9 expression profile required to permit drive in our system. On one end of the spectrum, we observed that Cas9 expression in later stages of

102 spermatogenesis permits fertility, while being too late to enable drive (B2dGDdRPLT Figure 3.12), and on the other end of the spectrum, Cas9 ‘overexpression’ causes complete infertility or death (Figure 3.13,

Figure 3.16). This emphasizes the precision with which Cas9 must be tuned along a narrow spectrum – one that ranges from non-driving fertility, to driving fertility, to complete infertility. In our system, if mosquitoes express known drive-enabling Cas9, and are fertile, then we should observe drive. If we fail to observe drive in these conditions, then we can glean valuable insights about our drive designs and the assumptions we made during design. We have observed a few such events, but the results were unexpected. In these experiments, not only was drive not observed, but the opposite occurred – a bias against the inheritance of CRISPR-encoding transgenes. In broods from males with mNanosGDdRPLT rescued by VasCas9, we observed a single brood with completely nontransgenic larvae (when at least half should have inherited mNanosGDdRPLT), and in a similar brood from ZPG1dGDpB driving dRPLTu we observed a strong bias against inheritance of the CRISPR-encoding distal drive in early broods. In the escaping fertile larvae from these broods, there was no evidence of any DRAs at any of the eight gRNA binding sites, suggesting these larvae had escaped mutagenesis (as is known to occur in other drives2,18).

Interestingly, as the mosquito aged, the ability of larvae to tolerate inheritance of the CRISPR-expressing transgene increased, a pattern consistent with possible PIWI silencing of the transgene as the sire aged.

These findings suggest that Cas9 expression was sufficiently fine-tuned to permit fertility in the presence of driving Cas9, but that drive did not occur. This implies that targeting RpL11 with CRISPR in the primordial germline may be fundamentally incompatible with fertility, as only those sperm which escaped mutagenesis sired viable larvae. Importantly, because successful drive homing should enable survival of the germ cells, yet this is not observed, this suggests that repair of breaks by HDR was not the predominant pathway for this drive design. Taken in conjunction with the findings of Oberhofer et al16, this raises the possibility that multi-gRNA drive designs may not stimulate HDR as originally hypothesized, and that end- joining mechanisms may precede or supersede HDR in these designs. While redesigning the drives to

103 utilize a Cas9 paired nickase strategy55 to stimulate HDR may be a possibility, it was beyond the scope of this work. Another strategy to stimulate HDR by Cas9 via fusion to geminin, which aims to localize Cas9 to post-S-phase stages of the cell cycle56 is not recommended, as drive must happen between homologs, not sister chromatids, making homing during pre-S-phase stages preferable. In all cases, these gene drives must be redesigned to stimulate robust HDR in the presence DNA-breaks from multiple gRNAs before success can be achieved.

To date, the only proposed gene drive designs designed for evolutionarily stable population replacement are those which rely on targeting haplolethal genes with multiple gRNAs to provide strong selection pressures against the evolution of DRAs1. Therefore this work raises the possibility that such homing based gene drives for population replacement may never come to fruition if i) Cas9 cannot be tuned perfectly to permit targeting of ribosome genes in the germline while maintaining fertility and drive (or if a tolerant non-ribosomal haplolethal gene cannot be identified), and ii) if multi-gRNA designs cannot be adapted to stimulate robust homing.

104

expression profiles.

-

s attemptings to demonstrate drive under differing Cas9

Summary of experimen7

| Summary| Goldilocks Cas9 Conundrum.

Figure 3.17 105

METHODS gRNA Design For CrIGCkid and gene drives

To design gRNAs against RpL11 and Rpt1 genes we verified the sequence of these genes in our laboratory stock lines by PCR and sequencing of 12 individuals. For each gene we identified two gRNA pairs targeting the 3’ region of the coding sequence (for a total of 8 gRNAs) in an paired-nickase orientation57 using the gRNA design tools outlined in Chapter 2. All 8 gRNAs used in CrIGCkid, gene drive, and distal gene drive experiments were clonal. The gRNA pairs were designed as tandem pairs linked by a 15-25bp RNA linker predicted to minimize aberrant secondary structures following analysis in RANfold webserver

(http://rna.tbi.univie.ac.at/cgi-bin/RNAWebSuite/RNAfold.cgi). The gRNAs were ordered as gBlocks with a U6 promoter for expression (outlined in Chapter 2). The pair gRNAn and gRNAo targeted the exon 3

RpL11 sequences [5’ AAACCGTAGATACCGATCGTCGG 3’] and [5’ TTGGACTTCTACGTTGTGCTTGG 3’] separated by a 24 bp random linker [5’ TCAGATAAAAAAAAAGCACCTGTT 3’]. The pair gRNAl and gRNAm target the exon 4 RpL11 sequences [5’ TCGCGTCCTCCTTGGTCAGACGG 3’] and [5’

TGGTTCCAGCAGAAGTACGACGG 3] separated by a 28 bp random sequence [5’

TCACTCTGACCAAGTCTGATGCCACCTG ‘3]. The pair gRNAs and gRNAr target the Rpt1 exon 6 sequences

[5’ TCACCTTGTTGACCGCCTCGAGG 3’] and [5’ CCAAGTTCAGCGCCACGCCGAGG 3’] separated by a random

[5’ GAGGAGGTCAAAAGTTGATCCGACTCGG 3’] linker. The pair gRNAp and gRNAq target the Rpt1 exon 6 sequences [5’ TGTCGGTCGAGCGGGACATTCGG 3’] and [5’ GAATCTTGAATATGTGCGTGCGG 3’] separated by a 24 bp [ 5’ ACTCTGTCCTGCTCCTGCACGTTCCA 3’] linker.

Recoding RpL11 and Rpt1 genes

To recode the endogenous Rpl11 and Rpt1 genes the boundaries for recoding were determined to be immediately upstream of the ‘outermost’ gRNA pairs designed. Recoding was performed manually using

106 the codon optimization table outlined in Volohonski et al.12, with as many codons as possible being replaced with the next-most common codon, and with no codons of >10% frequency being used.

Recoding of Rpl11 commences 620 bp from the start codon, pre-splices (removes) the 4th intron, with 55 bp of the remaining 178 bp (31%) being recoded. Recoding was performed to leave the original PAM sequences of gRNA binding sites intact, to enable subsequent reversal drives to use the recoded sequences as secondary gRNA binding sites. From outermost to innermost, the four gRNAs targeting

RpL11 have 65%, 65%, 75% and 75% identity to the recoded version of the gene respectively. To maintain ribosome function, a 209 bp replacement 3’UTR from RpL32 (AGAP002122) was added immediately following the stop codon. RpL32 is has a similar regulation profile and similar mRNA expression levels to

RpL11 (https://www.vectorbase.org), while having no predictable homology with the endogenous RpL11

3’UTR sequence.

Recoding of Rpt1 commences 1366 bp from the start codon and alters 91 bp of the remaining 253 bp (36% sequence recoding). Recoding was performed to leave the original PAM sequences of gRNA binding sites intact, to enable subsequent reversal drives to use the recoded sequences as secondary gRNA binding sites. The four gRNAs targeting Rpt1 have 65%, 60%, 65% and 50% identity to the recoded version of the gene respectively. A 3’UTR was omitted in dRPLT, however in dRPLTu a 198 bp replacement 3’UTR from

Rpt3 (AGAP003008) was added immediately following the stop codon to maintain wild type RPT1 subcellular localization.

Plasmid Construction

All plasmids and cloning was carried out through standard molecular biological methods including standard molecular cloning58, GoldenGate cloning59,60, or Gibson cloning 61. Synthetic sequences were ordered as IDT gBlocks, and naturally occurring sequences were PCR amplified from genomic DNA samples. For clarity here, and due to the variety of cloning methods used, sequences added to the 5’

107 terminus of primers intended for restriction enzyme digest or Gibson cloning are omitted and only the sequence annealing to the intended DNA sequence for amplification is presented.

HDRdRPLTdonor: This gene drive docking site homology donor was designed for knockin of dRPLT into

RpL11-Rpt1 in the genome by HDR. The plasmid was constructed within the pDSAY transgenesis backbone which provides a Kanamycin resistance cassette, and a 3xP3-EYFP cassette with Sv40 terminator12. First assembled were the homology arms, recoding and 3’UTRs. The homology arm encompassing Rpt1 was

PCR amplified from Anopheles gambiae genomic DNA using primers [5’ GTTGTCAACCAAAGTCTGTTGGCG

3’] and [5’ CCCTCCAGATCCGGCAGCCCGAAC 3’]. This was cloned into the AfLII site on PDSAY. Next in sequence and in frame with the Rpt1 homology fragment was cloned the codon recoding (ordered as a gBlock) with the sequence of [5’

CCGtACcCAtATtTTtAAaATcCAtGCgaGaagcATGagcGTgGAaCGtGAtATcCGGTTcGAatTGcTaGCgaGgCTcTGtC

CcAAtagcACgGGaGCtGAaATtCGaagtGTcTGCACGGAaGCaGGgATGTTcGCtATaCGGGCtaGgaGaAAaGTcGCc

ACcGAaAAaGAtTTCCTgGAaGCcGTgAAtAAaGTcATtAAatcgTAtGCgAAaTTttcgGCgACcCCccGGTAcATGACcTA tAATTAG 3’]. Within this sequence, recoded DNA bases are denoted in lowercase and original wild type base pairs in capitals. In this donor plasmid, no additional 3 UTR was added after the stop codon provided in this recoding. This was followed by a unique sequence not otherwise discussed in this work. This sequence included deliberately designed and optimized gRNA binding sites for second generation gene drives capable of targeting and homing through the recoded sequences in this first generation gene drive, to increase drive effectiveness, and included a further 100 random base pairs to enable synthesis. This sequence was ordered as a gBlock with [5’

CCCAAGACCACATAAAAATCGAGACAGAATCCCAGGCCATAAATGCCAATGGTCGGTTTCGTATGCGATAACTCAA

GTAGATGGCTACTCCCAAGACATTCGTTTCGTTCTAAGAAGCCTCGGAGCATCCAAGACTACGACCTGGAAGGTAC

GTATCCATCATATTTTTGTTGAAACCAGCCAATAAGCTGGCATCCTCCTTCGTAAGCCGG 3’] . Following this unique sequence phiC31 attP sequences were ordered as clonal gBlocks [ 5’

108

GTACTGACGGACACACCGAAGCCCCGGCGGCAACCCTCAGCGGATGCCCCGGGGCTTCACGTTTTCCCAGGTCAG

AAGCGGTTTTCGGGAGTAGTGCCCCAACTGGGGTAACCTTTGAGTTCTCTCAGTTGGGGGCGTAGGGTCGCCGAC

ATGACACAAGGGGTTGTGACCGGGGTGGACACGTACGCGGGTGCTTACGACCGTCAGTCGCGCGAGCGCGA 3’] to be compatible with attB sequences on Anopheles gambiae transgenesis plasmids. Two attP sites were cloned immediately upstream and downstream of the 3xP3-EYFP-Sv40 fluorescence cassette on pDSAY to enable later RCME with gene drive transgenes onto the docking site, and the LacZ MCS, and attB sequences were removed from pDSAY in the process. Into the AscI site on pDSAY was then cloned the

RpL11 homology and recoding sequence in the native orientation to enable compatible HDR knockin with the endogenous RpL11-Rpt1 locus. The RpL11 sequence was PCR amplified from genomic DNA using the primers [5’ TGTTATTGGAAAGTACGTATCGTT 3’] and [5’ ATCGTACTTGATGCCCAGATC 3’]. This was cloned upstream of the RpL11 recoded DNA sequence ordered as a gBlock with the final intron pre-spliced [ 5’

CCGACcATtGGcATtTAtGGccTcGAtTTtTAtGTgGTctTgGGCCGTCCAGGATAtAAtGTgGCgCAaaGgaGgCGtAAaA

TGGGtAAaATtGGgTTtATaCACCGgCTtACgAAGGAGGAtGCcATGAAaTGGTTtCAaCAaAAaTAtGAtGGcGTgATt

CTcAAttcgAAaGCcAAaTAA 3’]. Again with recoded DNA bases denoted in lowercase and original wild type basepairs in capitals. This is followed by a surrogate replacement 3’UTR from Rpl32 amplified from genomic using primers [5’ GCCCTGGTGCAAGAGAG 3’] and [5’ AAATTTGCCGTTTTAAGCGTAAT 3’]. This is followed by an additional Sv40 terminator sequence in an attempt to guarantee termination. This is followed by an additional unique sequence to provide additional gRNA binding sites for next generation drives homing through the first generation drives designed here. It was ordered as a gBlock [ 5’

AACAAACGCGAGATACCGGAAGTCTGCACCGgATaTCaCGtTCcACgctCAGCTCACGGGAGgATtTTaAAaATaTGg

GTaCGGGGTGCCACCTGTACGGCTCGTTGAAGCCGATTAGTACAATAGATTTATTCAACCCCAAAGGTCTACACTC

CCGGCCTACTCTTAGCTGATATGTCGCGTCCCggGGgGTcGCcgaaAAtTTcGTTCAATCGAGTgACtTTaTTcACgGCt

TCcAGG 3’]. This sequence immediately leads into the attP site previously cloned after the 3xP3-EYFP-

Sv40 casette discussed earlier. For a graphical plasmid map refer to Appendix A.

109

IGCdRPLTpB: This transgenic homology donor plasmid was designed to integrate the dRPLT donor sequence into the genome and to facilitate subsequent interlocus gene conversion- mediated knockin of the dRPLT docking site into endogenous RpL11-Rpt1. To build this plasmid, the entire donor cassette was

PCR amplified (encompassing both homology arms, recoding and fluorescence markers) from

HDRdRPLTdonor using the primers [5’ GTTGTCAACCAAAGTCTGTTGGCG 3’] and [5’

TGTTATTGGAAAGTACGTATCGTTG 3’]. This was cloned into the pXL-BACII-LoxP-3xP3DsRed-LoxP

(Addgene plasmid 26852) multiple cloning site to provide piggyBac terminal repeats for transgenesis.

Outside of the homology arms the 8 gRNAs were cloned in four gBlocks with gRNAp gRNAq, gRNAr, and gRNAs upstream of the RpL11 homology arm and gRNAl, gRNAm, gRNAn and gRNAo upstream of the Rpt1 homology arm. Outside of the homology arms, the 3xP3 promoter originally present on the pXL-BACII-

LoxP-3xP3DsRed-LoxP plasmid was swapped for an Actin5c promoter amplified from Drosophila melanogaster genomic DNA with primers [5’ CGCATGTGCTTGTGTGTGAG 3’] and [5’

TGTAAGCTGCAATGGAAAGAATGC 3’. On the backbone of this plasmid was cloned a Vasa-PiggyBac transposase expression cassette composed of the Vas2 promoter amplified with [5’

CAGGTCTCACATGCGATGTAGAACGCGAGCAAA 3’] and [5’ ATTGTTTCCTTTCTTTATTCACCGG 3’] expressing piggyBac transposase amplified with [5’ CAGGTCTCAATGGGCTCTAGCCTGGAC 3’] and [5’

CAGGTCTCACATGTCAGAAACAGCTCTGGCACATG 3’].

IGCdRPLTupB: This transgenic homology donor plasmid was designed to integrate the dRPLTu donor sequence into the genome and to facilitate subsequent interlocus gene conversion- mediated knockin of the dRPLT docking site into endogenous RpL11-Rpt1. This plasmid was built from the IGCdRPLTpB donor plasmid. For this the 3xP3-EYFP selectable marker cassette on IGCdRPLTpB was removed and replaced the a m2Tursuoise fluorescence sequence amplified from pDSAT12 using the primers [5’

ATGGTGAGCAAGGGCGAG 3’] and [5’ TCATCCGGACTTGTACAGC 3’]. Also onto this plasmid, a surrogate replacement 3’UTR for Rpt1 was added from Rpt3 which was amplified from genomic DNA by [5’

110

ACACAGCAGAAGCACAG 3’] and [5’ CCATTGCTTCAACTTCCG 3’] and was cloned immediately following the stop codon within Rpt1 recoding.

B2GD: This transgenesis plasmid is a gene drive designed to insert within the docking sites by phiC31 transgenesis, carries the 8 gRNAs used throughout this work and is enabled by the β2-tubulin promoter for Cas9 expression. To build this plasmid, first the pDSAR transgenesis backbone plasmid12 was modified to remove the original attB site, and two new attB sites with the sequence of [5’ tcgaCGATGTAGGTCACaGTCTCGAAGCCGCGGTGCGGGTGCCAGGGCGTGCCCTTGGGCTCCCCGGGCGCGTACT

CCACCTCACCCATCTGGTCCATCATGATGAACGGGTCGAGGTGGCGGTAGTTGATCCCGGCGAACGCGCGGCGCA

CCGGGAAGCCCTCGCCCTCGAAACCGCTGGGCGCGGTGGTCACGGTGAGCACGGGACGTGCGACGGCGTCGGC

GGGTGCGGATACGCGGGGCAGCGTCAGCGGGTTCTCGACGGTCACGGCGGGCA 3’] were cloned into the NheI and BsrGI restriction enzyme sites on the pDSAR backbone. During the insertion of one attB’s an AvrII restriction enzyme site was added, into which a Vasa2-PhiC31 integrase-expressing cassette was inserted.

The entire integrase-expression cassette was PCR amplified from integrase helper plasmid 13012 with primers [5’ CGATGTAGAACGCGAGCAAATTC 3’] and [5’ CTAGCAATCGGTGCAAGCTTTCAC 3’]. Next the gNRAl, gRNAm, gRNAn, gRNAo – containing gBlocks were cloned into the AflII enzyme site with 3’ ends oriented towards the LacZ MCS adding an SbfI site followed by reconstituting the AflII sequence on the 3’ end. The gRNAp, gRNAq, gRNAr, gRNAs- containing gBlocks were then cloned into the AscI site oriented with 3’ ends facing the LacZ MCS to reduce the probability of plasmid deletion between the gRNA blocks instead forcing shuffling events to result in an inversion and not a deletion. The B2Cas9 expression cassette was next cloned within the SbfI and AflII sites in two different versions. A version of the β2 promoter corresponding to the one outlined in29 composed of 2,234 bp upstream of the start codon were used for expression. It was amplified from genomic DNA by [5’ GAAGACGTTTGGAACGAGC 3’] and [5

TTTCGAAACTGTGAAAGCCG 3’], cloned upstream of Cas9 amplified from VasCas9 using primers [5’

ATCCCGATGTAGAACGCGAGC 3’] and [5’ TTACTTTTTCTTTTTTGCCTGGCC 3’], followed by a 1,500 bp β2

111

3’UTR amplified from genomic using primers [5 AACTAAAGCTAAATTGAACACCC 3’] and [5’

TCCGACATTAACATCTACGTGC 3’]. This plasmid was used for transgenesis injections into dRPLT. A second version of the plasmid using a shorter β2 promoter (500 bp) was used for injections into dRPLTu based on sequences characterized in 62,63, shown to be identical in function to 29 but significantly shorter in length.

This promoter sequence was amplified with primers [5’ AGCGTTCATAATTGATATAGTTTTG 3’] and [5’

AAGCTTGATATCTTTCGAAACTGTG 3’], cloned upstream of Cas9 amplified as before, followed by a 503 bp

β2 tub 3’ UTR amplified with [5’ TCCGGATGAAAGCTTAACTAAAGCTAAATTGAACACCC 3’] and [5’

CGATTTAAGGACCGATTCC 3’]. mNosGD: This transgenesis plasmid is a gene drive designed to insert within the docking sites by phiC31 transgenesis, carries the 8 gRNAs used throughout this work and is enabled by a truncated Nanos

(AGAP006098) promoter for Cas9 mRNA but not protein expression. This gene drive was cloned in the same manner as B2GD above, with the exception of enabling Cas9 expression by a Nanos promoter and

3’UTR in lieu of β2-tubulin’s. The promoter is composed of the 1,439 bp immediately upstream of the start codon and was amplified from genomic DNA by [5’ AATCGATTAACCGGCAAC 3’] and [5’

CTTGCTTTCTAGAACAAAAGG 3’]. This was cloned upstream of eSpCas964, in an attempt to improve on- target specificity, after demonstrating its ability to mutate target sites in An. gambiae embryos following in-vitro microinjection experiments. These experiments are not otherwise discussed in this work to maintain brevity and clarity, and this is the only implementation of eSpCas9 in this work. This eSpCas9 cassette was amplified with the primers [5’ ATGGACTATAAGGACCACGACGG 3’] and [5’

CTACTTTTTCTTTTTTGCCTGGC 3’]. Immediately following, 1,052 bp corresponding to the Nanos 3’UTR was

PCR amplified with [5’ TAGGACAGAGTCGTTCGTTC 3’] and [5’ GTCTTCATCCGAGAAAGGTAATG 3’].

VasGD: This transgenesis plasmid is a gene drive designed to insert within the docking sites by phiC31 transgenesis, carries the 8 gRNAs used throughout this work and is enabled by the Vas2 promoter for Cas9 expression. This gene drive was cloned in the same manner as B2GD above, with the exception of enabling

112

Cas9 expression by a Vas2 promoter and Sv40 terminator. The endogenous Vasa 3’UTR was decided to not be used as I had previously demonstrated that the Sv40 terminator in combination with this promoter was sufficient to enable robust germline endonuclease expression65, and that the combination also robustly enabled germline Cas9 expression specifically66. Therefore the Vasa2-Cas9-Sv40 cassette was amplified as one large fragment from the VasCas9 transgene using primers [ 5’

ATCCCGATGTAGAACGCGAGC 3’] and [5’ GATACATTGATGAGTTTGGACAAACC 3’] cloned into the drive plasmid.

B2dGDpB: This transgenesis plasmid is a distal gene drive designed to insert randomly throughout the genome by piggyBac transgenesis, carries the 8 gRNAs used throughout this work, and is enabled by the short β2-tubulin promoter62 and 3’UTR for Cas9 expression. This transgenic also sought to disprove fundamental B2GD lethality due to CRISPR expression alone. To construct this plasmid, first all the donor sequences between the homology arms on IGCdRPLTpB were removed, leaving the piggyBac terminal repeats, Vasa-transposase cassette, 8 gRNAs, and act-DsRed fluorescent marker. Into the gap caused by removal of the donor sequence, the B2-Cas9 expression cassette was inserted. This cassette was PCR amplified from the ‘shorter’ B2GD described above using he primers [ 5’ AGCGTTCATAATTGATATAGTTTTG

3’] and [5’ CGATTTAAGGACCGATTCC 3’]

VFdGDpB: This transgenesis plasmid is a distal gene drive designed to insert randomly throughout the genome by piggyBac transgenesis, carries the 8 gRNAs used throughout this work, and is enabled by the

VasaFusion protein system for Cas9 expression, and was constructed in the same manner as B2dGDpB. The

Vasa2 promoter was first PCR amplified with primers [5’ ATCCCGATGTAGAACGCGAGC 3’] and [5

ATTGTTTCCTTTCTTTATTCACCGG 3’] directly from VasGD. Next was cloned in a synthesized Vasa-encoding pre-spliced DNA sequence prefaced by a 3xFLAG tag followed by a Sv40 NLS ordered as a gBlock with the sequence [5’

ATGGACTATAAGGACCACGACGGAGACTACAAGGATCATGATATTGATTACAAAGACGATGACGATAAGATGGCC

113

CCAAAGAAGAAGCGGAAGGTCAGCGGAGACGGTGAATGGGATGATTGTGACGAGGGACGTAGCTTCGATCAGC

CCAAGTACGATGCTACCGAAAATAGTGTTCAGGACAACGATACTAACGGGTTCGACAACTATCAGTCCAACAATG

GTTTCGGTGACGAGTATCAAAGCAACGACAACGGCGGGTATGGTGGAGGTGACGATGGCTACGGTGGAGGTGG

CCGTGGCGGTCGAGGCGGTCGGGGTGGAGGACGTGGTCGCGGTCGAGGCCGCGGCGGACGTGATGGTGGAGG

GGGCTTCGGTGGAGGTGGATACGGTGATCGAAACGGCGATGGTGGGCGACCTGCCTATTCTGGCAACAGCGATC

CAAGCATGGACCAGGTGAAGACGGACAAGCCGAGGGAGCTGTACATTCCGCCACTTCCGACCGAGGACGAATCG

CTGATCTTCGGGTCCGGCATCAGCTCCGGGATCAATTTTGACAAGTTCGAGGAAATCCAGGTGCGCGTGTCGGGC

GAAAATCCACCGGATCACGTGGAGAGCTTCGAGCGCTCCGGTTTGCGCGAGGAGGTGATGACGAACGTGCGCAA

GTCGAGCTACACGAAACCGACGCCCATCCAGCGGTACGCCATCCCGATCATCCTGAACGGGCGCGACCTGATGGC

CTGCGCCCAGACCGGTTCGGGCAAGACGGCCGCCTTCATGCTGCCGATGATCCACCATCTGCTGGACAAGGAAGA

CTCGCTGGAGCTGCGCACGCGCAACCCGTACATCGTGATCGTGGCCCCGACCCGCGAGCTCGCGATTCAAATCCA

CGACGAGGGGCGCAAGTTCGCGCACGGCACCAAGCTGAAGGTCTGCGTATCGTACGGCGGCACAGCCGTACAGC

ACCAGCTGCAGCTGATGCGCGGTGGCTGCCACGTGCTGGTGGCAACGCCCGGTCGGCTGCTGGACTTCATCGACC

GGGGGTACGTGACGTTTGAGAACGTGAACTTCGTGGTGCTCGACGAAGCCGATCGGATGCTGGACATGGGCTTTC

TGCCCTCGATCGAGAAGGTGATGGGACACGCGACGATGCCGGAAAAGCAGCAGCGGCAGACGCTCATGTTTTCG

GCTACGTTCCCGGCCGAAATCCAGGAGCTGGCGGGCAAGTTCCTGCACAACTACATCTGCGTGTTCGTGGGCATC

GTCGGCGGAGCGTGTGCCGATGTGGAGCAGACGATCCATCTGGTGGAAAAGTTCAAGAAGCGCAAGAAGCTGG

AGGAGATCCTGAACGGCGGCAATCCCAAGGGTACGCTCGTGTTCGTGGAGACGAAGCGCAATGCGGACTATCTG

GCATCGCTGATGTCCGAGACGCAGTTCCCGACCACCTCCATCCATGGCGATCGGTTGCAGCGGGAGCGCGAGATG

GCGCTGTACGATTTCAAGTCCGGGCGGATGGACGTGCTGATCGCGACGTCGGTAGCGGCCCGCGGGCTGGACAT

TAAAAATGTGAACCACGTGGTGAACTACGATCTGCCGAAGAGCATCGATGATTACGTGCACCGTATCGGGCGAAC

CGGGCGCGTCGGTAACAAGGGCCGAGCGACTAGCTTCTACGACCCGGAGGCGGACCGGGCTATGGCGAGCGATC

TGGTGAAGATTCTCACCCAGGCCGGGCAGTCGGTGCCGGACTTTTTGAAGGATGCTGGCGGCTCCGGCTCGTACA

TGGGATCGTCCCAGTTCGGTGGCAAGGATATCCGCGACTCGTACGGTTCGCGTGTCGATGCCCAGCCGGTTGCCC

114

TCGAGCCGGAGGAAGAGTGGGAG 3’]. This was followed by an XTEN linker demonstrated previously to enable linkage of Cas9 to other large proteins67,68 with the sequence [5’

TCGGGCAGCGAGACGCCGGGCACCTCGGAGTCGGCCACCCCGGAGTCG 3’]. This is followed by linkage to

Cas9 amplified as previously described, followed by a 991 bp Vasa 3’UTR amplified from genomic by [5’

CTTGGGGTGGGGTTGTTATGTG 3’] and [5’ AGAAAATGTGGCCATTAACAGC 3’].

ZPG2dGDpB: This transgenesis plasmid is a distal gene drive designed to insert randomly throughout the genome by piggyBac transgenesis, carries the 8 gRNAs used throughout this work, and is enabled by an

~2 kB ZPG promoter sequence, and is constructed in the same manner as B2dGDpB. This promoter was previously developed in the lab for use in a sterilization rescue transgenic not otherwise discussed in this work, termed TOZR. This promoter sequence was effectively used to express the Tet-off transcriptional activator, and had been demonstrated to successfully express Tet-Off mRNA, but was never able to be validated by preliminary Western blots (not discussed). Given the lack of an alternative, verification of mRNA was considered sufficient evidence to warrant use of this germline promoter for construction of a distal gene drive. The 1,998 bp preceding the ZPG start codon were PCR amplified from genomic DNA by primers [5’ GGTGTCGAGCTTGAAGATTTC 3’] and [5’ CTCGATGCTGTATTTGTTGTTG 3’]. This promoter was cloned immediately upstream of the Cas9 previously discussed, followed by a 1,016 bp 3’ UTR sequence amplified by [5’ AGGACGGCGAGAAGTAATCATATG 3’] and [5’ GGGAATTCTAACCGTTAAATCGATTC 3’].

ZPG1dGDpB: This transgenesis plasmid is a distal gene drive designed to insert randomly throughout the genome by piggyBac transgenesis, carries the 8 gRNAs used throughout this work, and is enabled by an

~1 kB ZPG promoter sequence previously proven to enable robust drive-enabling Cas9 expression while being more fine-tuned than Vasa2 4,31. It is constructed in the same manner as B2dGDpB. The promoter and 3’UTR sequences were ordered as clonal gBlocks to those previously published. The 1,074 bp promoter gBlock was ordered with the sequence [5’

CAGCGCTGGCGGTGGGGACAGCTCCGGCTGTGGCTGTTCTTGCGAGTCCTCTTCCTGCGGCACATCCCTCTCGTCG

115

ACCAGTTCAGTTTGCTGAGCGTAAGCCTGCTGCTGTTCGTCCTGCATCATCGGGACCATTTGTATGGGCCATCCGCC

ACCACCACCATCACCACCGCCGTCCATTTCTAGGGGCATACCCATCAGCATCTCCGCGGGCGCCATTGGCGGTGGT

GCCAAGGTGCCATTCGTTTGTTGCTGAAAGCAAAAGAAAGCAAATTAGTGTTGTTTCTGCTGCACACGATAATTTT

CGTTTCTTGCCGCTAGACACAAACAACACTGCATCTGGAGGGAGAAATTTGACGCCTAGCTGTATAACTTACCTCA

AAGTTATTGTCCATCGTGGTATAATGGACCTACCGAGCCCGGTTACACTACACAAAGCAAGATTATGCGACAAAAT

CACAGCGAAAACTAGTAATTTTCATCTATCGAAAGCGGCCGAGCAGAGAGTTGTTTGGTATTGCAACTTGACATTC

TGCTGCGGGATAAACCGCGACGGGCTACCATGGCGCACCTGTCAGATGGCTGTCAAATTTGGCCCGGTTTGCGAT

ATGGAGTGGGTGAAATTATATCCCACTCGCTGATCGTGAAAATAGACACCTGAAAACAATAATTGTTGTGTTAATT

TTACATTTTGAAGAACAGCACAAGTTTTGCTGACAATATTTAATTACGTTTCGTTATCAACGGCACGGAAAGATTAT

CTCGCTGATTATCCCTCTCGCTCTCTCTGTCTATCATGTCCTGGTCGTTCTCGCGTCACCCCGGATAATCGAGAGAC

GCCATTTTTAATTTGAACTACTACACCGACAAGCATGCCGTGAGCTCTTTCAAGTTCTTCTGTCCGACCAAAGAAAC

AGAGAATACCGCCCGGACAGTGCCCGGAGTGATCGATCCATAGAAAATCGCCCATCATGTGCCACTGAGGCGAAC

CGGCGTAGCTTGTTCCGAATTTCCAAGTGCTTCCCCGTAACATCCGCATATAACAAACAGCCCAACAACAAATACA

GCATCGAG 3’], and was cloned immediately upstream of Cas9 as previously described. This is followed by a 1,037 bp 3’UTR ordered as a gBlock with the sequence [ 5’

CAGCGCTGGCGGTGGGGACAGCTCCGGCTGTGGCTGTTCTTGCGAGTCCTCTTCCTGCGGCACATCCCTCTCGTCG

ACCAGTTCAGTTTGCTGAGCGTAAGCCTGCTGCTGTTCGTCCTGCATCATCGGGACCATTTGTATGGGCCATCCGCC

ACCACCACCATCACCACCGCCGTCCATTTCTAGGGGCATACCCATCAGCATCTCCGCGGGCGCCATTGGCGGTGGT

GCCAAGGTGCCATTCGTTTGTTGCTGAAAGCAAAAGAAAGCAAATTAGTGTTGTTTCTGCTGCACACGATAATTTT

CGTTTCTTGCCGCTAGACACAAACAACACTGCATCTGGAGGGAGAAATTTGACGCCTAGCTGTATAACTTACCTCA

AAGTTATTGTCCATCGTGGTATAATGGACCTACCGAGCCCGGTTACACTACACAAAGCAAGATTATGCGACAAAAT

CACAGCGAAAACTAGTAATTTTCATCTATCGAAAGCGGCCGAGCAGAGAGTTGTTTGGTATTGCAACTTGACATTC

TGCTGCGGGATAAACCGCGACGGGCTACCATGGCGCACCTGTCAGATGGCTGTCAAATTTGGCCCGGTTTGCGAT

ATGGAGTGGGTGAAATTATATCCCACTCGCTGATCGTGAAAATAGACACCTGAAAACAATAATTGTTGTGTTAATT

116

TTACATTTTGAAGAACAGCACAAGTTTTGCTGACAATATTTAATTACGTTTCGTTATCAACGGCACGGAAAGATTAT

CTCGCTGATTATCCCTCTCGCTCTCTCTGTCTATCATGTCCTGGTCGTTCTCGCGTCACCCCGGATAATCGAGAGAC

GCCATTTTTAATTTGAACTACTACACCGACAAGCATGCCGTGAGCTCTTTCAAGTTCTTCTGTCCGACCAAAGAAAC

AGAGAATACCGCCCGGACAGTGCCCGGAGTGATCGATCCATAGAAAATCGCCCATCATGTGCCACTGAGGCGAAC

CGGCGTAGCTTGTTCCGAATTTCCAAGTGCTTCCCCGTAACATCCGCATATAACAAACAGCCCAACAACAAATACA

GCATCGAG 3’].

Embryonic microinjection and transgenesis

Procedures carried out essentially as described in12,14,69 with 350ng/µl of transgenesis plasmid injected and 80 ng/µl of helper plasmid injected if by PhiC31 transgenesis. Transgenesis into dRPLT or dRPLTLu were into the mixed offspring of {+/dRPLTx} intercrossing adults due to the homozygous lethality and infertility of the lines respectively as previously discussed. Due to inability to isolate heterozygous and homozygous docking site embryos from wild type siblings, all genotypes were indiscriminately aligned and injected, leading to reduced transgenesis efficiencies as wild type embryos would not be capable of transgenesis, and homozygotes would be unable to begat transgenic offspring. Following injections into the docking sites, individuals were outcrossed to the wild type G3 laboratory stock line in bulk, and nascent transgenic F1 offspring were identified by fluorescence. PiggyBac transgenesis based injections to establish the IGCdRPLTpB and IGCdRPLTupB donor lines were carried out directly within wild type G3 embryos. Injections to establish the piggyBac distal gene drives were either carried out within G3 embryos, and survivors immediately outcrossed to {+/dRPLTu} to assay for distal drive immediately in the

F1’s , or the reciprocal was carried out with injections occurring directly in mixed {+/dRPLTu} embryos followed by F0 survivor outcross to G3 depending on the availability of each line. Under no circumstances were injections carried out in {+/dRPLTu} embryos followed by F0 survivor outcross to {+/dRPLTu} in order to guarantee all newly established distal drive transgenics could only be heterozygous for the dRPLTu site,

117 guaranteeing not false positives for drive. Precise injection and crossing parameters are available upon request.

PCR validation of dRPLT and dRPLTu insertion:

PCR validation of docking site insertion occurred using many combinations of primers. Some amplicons were sequenced to verify the product. Primer combinations used consist of those that specifically bind within the newly introduced sequence – often recoding, a replacement UTR, an attB or a fluorescent marker – and amplify over the homology arms and into a sequence within the genome. Not all primers worked given the polymorphic nature of the reference genome to our lines, and the polymorphic nature of the 5’ regulatory sequences targeted. However all primers listed here worked to produce a valid PCR product of the predicted size in at least reaction, and not all possible combinations were tested. Primers used specific to the polymorphic genomic sequences beyond the RpL11 homology arm include [5’

GTGGAGCGTTTTTTCCACTTTCG 3’], [5’ CGGGACGGTTTATGAATGTATAACG 3’], [5’

GTAATTTGCTATGCCAATTGTCGTG 3’], [5’ CGGGGGATCCGGTTTTAATG 3’], [5

GCTACACACGACAAAAGTGCATC 3’], [5’ CAATTGTCGTGCAGTGGAGC 3’] and [5’

GCGATACTTGCTGATTTGAGTGC 3’]. PCRs specific to the integrated sequence used include those binding to the RpL11 recoding include [5’ GGCCCAAGACCACATAAAAATCG 3’], [5’ GAGGCCATAAATGCCAATGGTC

3’], [5’ GCGCCACATTATATCCTGGACG 3’], and [5’ GCGCCACATTATATCCTGGACG 3’]. Primers specific to the

Rpt1 recoding used include [5’ GATATCACGTTCCACGCTCATG 3’], [5’ GTATAGCGAACATCCCTGCTTCC 3’],

[5’ CAAGACGTCTCCGAATTCCTTC 3’], and [5’ GATATCACGTTCCACGCTCATG 3’]. A Primer specific to the

3’UTR replacement for RpL11 from RpL32 includes [5’ CTCCTCAAAACGGATCCTTTCTTC 3’]. Primers specific to the EYFP cassette within dRPLT include [5’ CTGAACTTGTGGCCGTTTACGTCG 3’], and a primer specific to m2 Turquoise is [5’ GTCGTCCTTGAAGAAGATGGTG 3’]. Primers specific to the attP or adjacent sequences include [5’ GGGTTCGAAATCGATAAGCTTGG 3’], [5’ CTTAAAGCTTATCGATACGCGTACG 3’], and

[5’ CATTTATGGCCTGGGATTCTGTC 3’]. Primers specific to the Rpt1 replacement 3’UTR from Rpt3 include

118

[5’ CAAGACGTCTCCGAATTCCTTC 3’]. Primers specific to the genomic sequences beyond the Rpt1 homology arm includes [5’ CACACCTTTACATATCGCTCGC 3’], [5’ GGCCTTCTCTGACCGATTCT 3’], and [5’

GACTGTGCAAATTTTGTCATTTGG 3’].

Larval Fluorescence Scoring

Egg broods were laid into an oviposition cup composed of a a Whatman® filter paper cone (90mm, Grade

2, Sigma-Aldrich) within a urinalysis cup containing 80 mL DI water. The small amount of water in the bottom of the cone ( 10 mL) collected and concentrated larvae for further analysis. Larvae were scored the afternoon of hatching, approximately 2 hours after first hatchlings emerged, and late hatchling larvae were scored the next morning. Larvae were removed from the cone of water in the base of the dish by a

1 mL pipette with the tip clipped to have an opening circumference between 1-3mm. The larvae were then further concentrated onto a sieve with a fine mesh to not injure hatchlings. Larvae were then pipetted onto Teflon-printed diagnostic slides, and immobilized on ice or on a Corning ® CoolBoxTM XT

Cooling workstation. Larvae were then analyzed under a Leica M80 fluorescence dissecting scope for fluorescence. Notably CrIGCkid larvae analysis occurred under a wideband GFP filter enabling simultaneous visualization of the presence for intense gut-derived act-DsRED and/or 3xP3-FP fluorescence. Larvae of desire florescence phenotypes were then isolated from siblings by serial dilution with a clipped tip 200µL pipette until isolated, and then reared apripriately.

Western blots:

Western blots were carried out essentially as described in 70. Tissue samples, quantity, age, and dissection status are provided in respective figure legends. They were dissected and prepared in Np-40 cell lysis buffer system (ThermoFisher Scientific) supplemented with 1mM PMSF in DMSO and 40µ Protease

Inhibitor . Western blots to detect Cas9 were either carried out with an α-FLAG tag antibody previously described70(an epitope included on all Cas9 constructs) when blots performed before G20, or directly

119 with an α-Cas9 antibody compatible with all Cas9’s used in this work from Cell Signaling Technologies

®(Cas9 XP ®, Rabbit mAB # 19526, 1:1000 dilution) when blotted after G20.

Westerns targeting poly-ubiquitin aggregates were carried out with 30, 1-day old, larvae of each genotype. Early replicates attempted larval treatment with the known proteasome inhibitor, Mg132, by overnight incubation of larvae in a 10mM solution in Di water. Larvae survived the night, and failed to demonstrate polyubiquitin aggregates in subsequent western. Therefore a murine tissue culture sample with known proteasome defects was used as a positive control composed of 1µg/µL C2C12 cell line trated with Bortezomib. Inspired by 25, Polyubiquitin aggregates were finally visualized following release from aggregates via treatment of the membrane with 6M GuHCl, 20mM Tris pH 7.5, 1mM PMSF and 5mM β-

ME for 30 min at 4°, before blocking and antibody incubation with 1:500 solution of the Fk2 Mono- and polyubiquitinated antibody (Enzo Life Sciences, BML-PW150-0025).

Western blots targeting RpL11 were undertaken using four different primary antibodies, all of which failed to identify the anopheles homolog of RpL11 : RpL11 polyclonal (PA5-27468, ThermoFisher Scientific),

RpL11 polyclonal (GTX101651, GeneTex), RpL11 polyclonal (aa128-178, LS-C290580, LifeSpan

BioSciences), and RpL11 Polyclonal (LS-C190025, LifeSpan BioSciences).

Larval competition assays:

For larval competition trays they were set up similarly. Hatchling larvae from intercrossing heterozygous

{+/dRPLT} or {+/dRPLTu} parents. For dPRLT, larvae were screened 20H post hatching, and three cohorts were isolated, each containing exactly 50 wt and 50 dRPLT heterozygous larvae, with great care taken to exclude homozygotes by fluorescence and size differentiation. All three cohorts were reared according to standard husbandry protocols outlined in the MR4 manual

(https://www.beiresources.org/Portals/2/VectorResources/2016%20Methods%20in%20Anopheles%20

Research%20full%20manual.pdf). Larvae were reared in adjacent trays at 28° in a -controlled

120 insectary. All trays were given as similar quantities of larval food as possible. Trays were monitored each day for development of pupae. On the day of first pupae observed for that tray, it was denoted “day one of pupation”, the pupa was removed, scored for fluorescence and sex. Each day pupae were removed and scored similarly until no living pupae or larvae were present. Larval competition trays assaying dRPLTu development were carried out in an identical manner with the exception of the starting quantities of each genotype. In this case, to more closely mimic the normal distribution of larval genotype frequencies, 25 wt, 50 heterozygous, and 25 homozygous larvae were seeded into each tray at commencement of experiment. Pupae for this experiment were scored for simply the presence or absence of fluorescence, preventing quantification of dRPLTu heterozygotes vs homozygotes at this time, however daily samples were collected to enable later analysis of the differences between these two genotypes.

121

REFERENCES

1 Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife, e03401, doi:10.7554/eLife.03401 (2014).

2 Gantz, V. M. & Jasinskiene, N. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. Proceedings of the …, doi:10.1073/pnas.1521077112 (2015).

3 Hammond, A. M. et al. The creation and selection of mutations resistant to a gene drive over multiple generations in the malaria mosquito. PLoS genetics 13, e1007039, doi:10.1371/journal.pgen.1007039 (2017).

4 Kyrou, K. et al. A CRISPR–Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature biotechnology 36, 1062, doi:10.1038/nbt.4245 (2018).

5 Marshall, J. M., Buchman, A. & Akbari, O. S. Overcoming evolved resistance to population- suppressing homing-based gene drives. Scientific reports 7, 3776 (2017).

6 Nash, A. et al. Integral gene drives for population replacement. Biol Open 8, doi:10.1242/bio.037762 (2019).

7 DiCarlo, J. E., Chavez, A., Dietz, S. L., Esvelt, K. M. & Church, G. M. RNA-guided gene drives can efficiently bias inheritance in wild yeast. RNA-guided gene drives can efficiently bias inheritance in wild yeast (2015).

8 Kuo, J. et al. Synthetic genome recoding: new genetic codes for new features. Current genetics 64, 327-333, doi:10.1007/s00294-017-0754-z (2018).

9 DiCarlo, J. E., Chavez, A., Dietz, S. L., Esvelt, K. M. & Church, G. M. Safeguarding CRISPR-Cas9 gene drives in yeast. Nature biotechnology 33, 1250, doi:10.1038/nbt.3412 (2015).

10 Yan, Y. & Finnigan, G. C. Development of a multi-locus CRISPR gene drive system in budding yeast. Scientific reports 8, 17277, doi:10.1038/s41598-018-34909-3 (2018).

11 Cook, R. et al. The generation of chromosomal deletions to provide extensive coverage and subdivision of the Drosophila melanogaster genome. Genome biology 13, doi:10.1186/gb-2012- 13-3-r21 (2012).

12 Volohonsky, G. et al. Tools for Anopheles gambiae Transgenesis. G3 (Bethesda, Md.), doi:10.1534/g3.115.016808 (2015).

13 Bateman, J. R., Lee, A. M. & Wu, C. T. Site-specific transformation of Drosophila via phiC31 integrase-mediated cassette exchange. Genetics 173, 769-777, doi:10.1534/genetics.106.056945 (2006).

122

14 Pondeville, E. et al. Efficient PhiC31 integrase-mediated site-specific germline transformation of Anopheles gambiae. Nat Protoc 9, 1698-1712, doi:10.1038/nprot.2014.117 (2014).

15 Benedict, M. Q., McNitt, L. M., Cornel, A. J. & Collins, F. H. Mosaic: a position-effect variegation eye-color mutant in the mosquito Anopheles gambiae. The Journal of heredity 91, 128-133 (2000).

16 Oberhofer, G., Ivy, T. & Hay, B. A. Behavior of homing endonuclease gene drives targeting genes required for viability or female fertility with multiplexed guide RNAs. Proceedings of the National Academy of Sciences 115, E9343-E9352, doi:10.1073/pnas.1805278115 (2018).

17 Gantz, V. M. et al. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. PNAS Plus (2015).

18 Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nature biotechnology 34, 78-83, doi:10.1038/nbt.3439 (2015).

19 Lin, C. C. & Potter, C. J. Editing Transgenic DNA Components by Inducible Gene Replacement in Drosophila melanogaster. Genetics 203, 1613-1628, doi:10.1534/genetics.116.191783 (2016).

20 Medicine, E. (ed Dan Budwick) (2015).

21 Nolan, T., Bower, T. M., Brown, A. E., Crisanti, A. & Catteruccia, F. piggyBac-mediated germline transformation of the malaria mosquito Anopheles stephensi using the red fluorescent protein dsRED as a selectable marker. The Journal of biological chemistry 277, 8759-8762, doi:10.1074/jbc.C100766200 (2002).

22 Chen, J.-M., Cooper, D. N., Chuzhanova, N., Férec, C. & Patrinos, G. P. Gene conversion: mechanisms, evolution and human disease. Nature Reviews Genetics 8, 762 (2007).

23 Wójcik, C. & DeMartino, G. N. Intracellular localization of proteasomes. The international journal of biochemistry & cell biology 35, 579-589 (2003).

24 Szabo, A. et al. Molecular characterization of the Rpt1/p48B ATPase subunit of the Drosophila melanogaster 26S proteasome. Molecular genetics and genomics : MGG 278, 17-29, doi:10.1007/s00438-007-0223-3 (2007).

25 Koerver, L. et al. The de-ubiquitylating enzyme DUBA is essential for spermatogenesis in Drosophila. Cell Death & Differentiation 23, 2019-2030, doi:10.1038/cdd.2016.79 (2016).

26 Bhattacharyya, S., Yu, H., Mim, C. & Matouschek, A. Regulated protein turnover: snapshots of the proteasome in action. Nat Rev Mol Cell Biol 15, 122-133, doi:10.1038/nrm3741 (2014).

27 Lambertsson, A. The minute genes in Drosophila and their molecular functions. Advances in genetics 38, 69-134 (1998).

28 Cramton, S. E. & Laski, F. A. string of pearls encodes Drosophila ribosomal protein S2, has Minute-like characteristics, and is required during oogenesis. Genetics 137, 1039-1048 (1994).

123

29 Catteruccia, F., Benton, J. P. & Crisanti, A. An Anopheles transgenic sexing strain for vector control. Nature biotechnology 23, 1414-1417, doi:10.1038/nbt1152 (2005).

30 Galizi, R. et al. A synthetic sex ratio distortion system for the control of the human malaria mosquito. Nature communications 5, 3977, doi:10.1038/ncomms4977 (2014).

31 Hammond, A. M. et al. Improved CRISPR-based suppression gene drives mitigate resistance and impose a large reproductive load on laboratory-contained mosquito populations. bioRxiv, doi:10.1101/360339 (2018).

32 Grossman, G. L. et al. Germline transformation of the malaria vector, Anopheles gambiae, with the piggyBac transposable element. Insect molecular biology 10, 597-604 (2001).

33 Noble, C. et al. Daisy-chain gene drives for the alteration of local populations. Proceedings of the National Academy of Sciences, 201716358, doi:10.1073/pnas.1716358116 (2019).

34 Pompon, J. & Levashina, E. A. A New Role of the Mosquito Complement-like Cascade in Male Fertility in Anopheles gambiae. PLoS biology 13, e1002255, doi:10.1371/journal.pbio.1002255 (2015).

35 Galizi, R. et al. A CRISPR-Cas9 sex-ratio distortion system for genetic control. Scientific reports 6, 31139 (2016).

36 Champer, J. et al. Novel CRISPR/Cas9 gene drive constructs reveal insights into mechanisms of resistance allele formation and drive efficiency in genetically diverse populations. PLoS genetics 13, e1006796 (2017).

37 Papathanos, P. A., Windbichler, N., Menichelli, M., Burt, A. & Crisanti, A. The vasa regulatory region mediates germline expression and maternal transmission of proteins in the malaria mosquito Anopheles gambiae: a versatile tool for genetic control strategies. BMC molecular biology 10, 65, doi:10.1186/1471-2199-10-65 (2009).

38 Guilinger, J. P., Thompson, D. B. & Liu, D. R. Fusion of catalytically inactive Cas9 to FokI nuclease improves the specificity of genome modification. Nature biotechnology 32, 577-582, doi:10.1038/nbt.2909 (2014).

39 Wang, S.-C. et al. Germ plasm localisation of the HELICc of Vasa in Drosophila: analysis of domain sufficiency and amino acids critical for localisation. Scientific reports 5, 14703, doi:10.1038/srep14703 (2015).

40 Estep, A. S., Sanscrainte, N. D. & Becnel, J. J. DsRNA-mediated targeting of ribosomal transcripts RPS6 and RPL26 induces long-lasting and significant reductions in fecundity of the vector Aedes aegypti. J Insect Physiol 90, 17-26, doi:10.1016/j.jinsphys.2016.05.001 (2016).

41 Gilles, J. R. L., Vreysen, M. J. B. & Damiens, D. Anopheles arabiensis Sperm Production After Genetic Manipulation, Dieldrin Treatment, and Irradiation. Journal of Medical Entomology 50, 314-316, doi:10.1603/me12058 (2013).

124

42 de Cuevas, M. & Matunis, E. L. The stem cell niche: lessons from the Drosophila testis. Development 138, 2861-2869 (2011).

43 Moriyama, E. N. & Powell, J. R. Codon usage bias and tRNA abundance in Drosophila. Journal of molecular evolution 45, 514-523 (1997).

44 Welch, M., Villalobos, A., Gustafsson, C. & Minshull, J. Designing genes for successful protein expression. Methods in enzymology 498, 43-66, doi:10.1016/b978-0-12-385120-8.00003-6 (2011).

45 Cooper, G. M., Hausman, R. E. & Hausman, R. E. The cell: a molecular approach. Vol. 2 (ASM press Washington, DC, 2000).

46 Franz, A. W. et al. Stability and loss of a virus resistance phenotype over time in transgenic mosquitoes harbouring an antiviral effector gene. Insect molecular biology 18, 661-672 (2009).

47 Kalmykova, A. I., Klenov, M. S. & Gvozdev, V. A. Argonaute protein PIWI controls mobilization of retrotransposons in the Drosophila male germline. Nucleic acids research 33, 2052-2059, doi:10.1093/nar/gki323 (2005).

48 Al-Kaff, N. S. et al. Transcriptional and posttranscriptional plant gene silencing in response to a pathogen. Science 279, 2113-2115 (1998).

49 Baulcombe, D. RNA silencing in plants. Nature 431, 356 (2004).

50 Tabara, H. et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123-132 (1999).

51 Ozata, D. M., Gainetdinov, I., Zoch, A., O'Carroll, D. & Zamore, P. D. PIWI-interacting RNAs: small RNAs with big functions. Nature reviews. Genetics 20, 89-108, doi:10.1038/s41576-018-0073-3 (2019).

52 Vagin, V. V. et al. A distinct small RNA pathway silences selfish genetic elements in the germline. Science 313, 320-324, doi:10.1126/science.1129333 (2006).

53 Khurana, J. S. et al. Adaptation to P element transposon invasion in Drosophila melanogaster. Cell 147, 1551-1563, doi:10.1016/j.cell.2011.11.042 (2011).

54 Carballar-Lejarazú, R., Jasinskiene, N. & James, A. A. Exogenous gypsy insulator sequences modulate transgene expression in the malaria vector mosquito, Anopheles stephensi. Proceedings of the National Academy of Sciences 110, 7176-7181, doi:10.1073/pnas.1304722110 (2013).

55 Ran, F. A. et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154, 1380-1389, doi:10.1016/j.cell.2013.08.021 (2013).

56 Yang, S., Li, S. & Li, X. J. Shortening the Half-Life of Cas9 Maintains Its Gene Editing Ability and Reduces Neuronal Toxicity. Cell reports 25, 2653-2659.e2653, doi:10.1016/j.celrep.2018.11.019 (2018).

125

57 Mali, P. et al. RNA-guided human genome engineering via Cas9. Science 339, 823-826, doi:10.1126/science.1232033 (2013).

58 Sambrook, J., Russell, D. W. & Russell, D. W. Molecular cloning: a laboratory manual (3-volume set). Vol. 999 (Cold spring harbor laboratory press New York, 2001).

59 Engler, C. & Marillonnet, S. Combinatorial DNA assembly using Golden Gate cloning. Methods Mol Biol 1073, 141-156, doi:10.1007/978-1-62703-625-2_12 (2013).

60 Geissler, R. et al. Transcriptional activators of human genes with programmable DNA-specificity. PloS one 6, e19509, doi:10.1371/journal.pone.0019509 (2011).

61 Gibson, D. G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nature methods 6, 343 (2009).

62 Windbichler, N., Papathanos, P. A. & Crisanti, A. Targeting the X chromosome during spermatogenesis induces Y chromosome transmission ratio distortion and early dominant embryo lethality in Anopheles gambiae. PLoS genetics 4, e1000291, doi:10.1371/journal.pgen.1000291 (2008).

63 Magnusson, K. et al. Transcription regulation of sex-biased genes during ontogeny in the malaria vector Anopheles gambiae. PloS one 6, e21572, doi:10.1371/journal.pone.0021572 (2011).

64 Slaymaker, I. M. et al. Rationally engineered Cas9 nucleases with improved specificity. Science (New York, N.Y.) 351, 84-88, doi:10.1126/science.aad5227 (2016).

65 Smidler, A. L., Terenzi, O., Soichot, J., Levashina, E. A. & Marois, E. Targeted mutagenesis in the malaria mosquito using TALE nucleases. PloS one 8, e74511, doi:10.1371/journal.pone.0074511 (2013).

66 Werling, K. et al. Steroid Hormone Function Controls Non-competitive Plasmodium Development in Anopheles. Cell, doi:10.1016/j.cell.2019.02.036.

67 Guilinger, J. P., Thompson, D. B. & Liu, D. R. Fusion of catalytically inactive Cas9 to FokI nuclease improves the specificity of genome modification. Nature biotechnology 32, 577-582, doi:10.1038/nbt.2909 (2014).

68 Schellenberger, V. et al. A recombinant polypeptide extends the in vivo half-life of peptides and proteins in a tunable manner. Nature biotechnology 27, 1186 (2009).

69 Fuchs, S., Nolan, T. & Crisanti, A. Mosquito transgenic technologies to reduce Plasmodium transmission. Methods Mol Biol 923, 601-622, doi:10.1007/978-1-62703-026-7_41 (2013).

70 Smidler, A. L., Scott, S. N., Mameli, E., Shaw, W. R. & Catteruccia, F. A transgenic tool to assess Anopheles mating competitiveness in the field. Parasit Vectors 11, 651, doi:10.1186/s13071- 018-3218-5 (2018).

126

CHAPTER 4.

A transgenic tool to assess Anopheles mating competitiveness in the field

Andrea L. Smidler1, Sean N. Scott1, Enzo Mameli1, W. Robert Shaw1† and Flaminia Catteruccia1†

1 Harvard T.H. Chan School of Public Health, Department of Immunology and Infectious Diseases, Boston

MA, USA

† These authors contributed equally to this work

Respective Contributions: This chapter is based off a published work performed in collaboration with colleagues. ALS, WRS, and FC conceived of the research study. ALS performed all transgene design, cloning and embryonic microinjections, and initial transgene characterization. WRS conceived and coordinated experimental design and oversaw sample preparation and collection and analyzed data. SNS performed sample preparation and performed the IFA, Western blot, and RNA timecourse experiments. EM performed electron microscopy experiments. ALS and SNS shared in line husbandry, and EM, WRS and

SNS performed mating captures. FC contributed to the design of the study. ALS, WRS and FC wrote the manuscript.

1 Smidler, A. L., Scott, S. N., Mameli, E., Shaw, W. R. & Catteruccia, F. A transgenic tool to assess Anopheles mating competitiveness in the field. Parasit Vectors 11, 651, doi:10.1186/s13071-018-3218-5 (2018).

127

ABSTRACT

Malaria parasites, transmitted by the bite of an anopheline mosquito, pose an immense public health burden on many tropical and subtropical regions. The most important malaria vectors in sub-Saharan

Africa are mosquitoes of the Anopheles gambiae complex including A. gambiae sensu stricto. Given the increasing rates of insecticide resistance in these mosquitoes, alternative control strategies based on the release of genetically modified males are being evaluated to stop transmission by these disease vectors.

These strategies rely on the mating competitiveness of release males, however currently there is no method to determine male mating success without sacrificing the female. Here we exploit the male An. gambiae reproductive characteristic of transferring seminal secretions, made in the male accessory glands

(MAGs), into the female atrium during mating to generate a transgenic system capable of monitoring copula pairs. We develop and validate the use of a MAG-specific promoter to fluorescently label the mating plug and visualize the occurrence of insemination in vivo. We used the promoter region of the major mating plug protein, Plugin, to control the expression of a Plugin-tdTomato (PluTo) fusion protein, hypothesizing that this fusion protein could be incorporated into the plug for sexual transfer to the female.

A. gambiae PluTo transgenic males showed strong red fluorescence specifically in the MAGs and with a pattern closely matching endogenous Plugin expression. Moreover, the fusion protein was integrated into the mating plug and transferred to the female atrium during mating where it could be visualized microscopically in vivo without sacrificing the female. PluTo males were equally as competitive at mating as wild type males, and females mated to these males did not show any reduction in reproductive fitness.

The validation of the first MAG-specific promoter in transgenic A. gambiae facilitates the live detection of successful insemination hours after copulations has occurred. This provides a valuable tool for the assessment of male mating competitiveness not only in laboratory experiments but also in semi-field and field studies aimed at testing the feasibility of releasing genetically modified mosquitoes for disease control.

128

INTRODUCTION

Targeting fertility for insect control has been successful in a variety of species, and a number of strategies aimed at inducing sterility in field populations are currently being developed in mosquitoes. Sterile Insect

Technique (SIT), a strategy based on the mass release of irradiated males to sterilize females upon mating

7, has been successfully implemented against populations of the screwworm Cochliomyia hominivorax and the Mediterranean fruit fly Ceratitis capitata to reduce the economic burden of these important agricultural pests 8-10. Adapting SIT and other similar sterilizing technologies (such as chemosterilization) to Anopheles vectors of human malaria has been challenging, and requires optimization to maximize competitiveness of colonized, mass-reared and sterilized males 11-15. To circumvent the potential problems linked to sterility by irradiation or chemosterilants, strategies based on genetic manipulation of insect fertility are being considered. Among the most promising strategies, gene drives are in development for

Anopheles which could spread infertility through natural mosquito populations by positively biasing their own inheritance. The potential of gene drives to induce sterility in field populations has been recently demonstrated in cages of A. gambiae16, where gene drive transgenes targeting female fertility genes spread for a few generations17. Regardless of their mode of action, all genetic control strategies require the release of male mosquitoes that have high mating competitiveness and can successfully mate with field females, making studies into male reproductive fitness a critical requisite for successful implementation.

A. gambiae are largely monandrous, which means they mate only once in their lifetime. During copulation seminal secretions produced by the male accessory glands (MAGs) are transferred to the female atrium in the form of a coagulated mating plug 18. The plug is composed of seminal secretions including proteins

19 and the steroid hormone 20-hydroxyecdysone (20E) 20, and is digested by the female over a period of

24 h following copulation 21. The loss of female receptivity to further mating is at least partially due to the

129 transfer and function of the steroid hormone 20E 22-24, which also triggers oviposition in blood fed females and affects other important aspects of the female post-mating physiology, including egg development and fertility 22,23,25. One of the most abundant mating plug proteins is Plugin, a MAG-specific glutamine-rich structural protein that becomes incorporated into the plug upon cross-linking to itself and other seminal proteins by the action of the plug-forming transglutaminase AgTG3 21,26. Transcriptional silencing of AgTG3 partially prevents plug formation and transfer, which results in severe sperm storage defects in the female, causing infertility 21. Other less characterized mating plug factors that may play a role in female post-mating physiology include short accessory gland proteins (Acps), a number of proteases and peptidases 27-29, serine protease inhibitors which play a role in mammalian fertility 30, and cysteine-rich secretory folding proteins (CRISPs) that are involved in gamete interactions 31 . While many of these proteins are characterized in other organisms, their role in male mating fitness has not yet been elucidated in A. gambiae, making development of novel MAG-specific tools important for addressing these outstanding questions. Studies on the evolutionary trajectory of mating plug formation have shown that transfer of a mating plug is a derived trait not limited to A. gambiae but widely conserved across anopheline species from Africa and South East Asia, while being absent in males of Central American species like A. albimanus 24.

The presence of a mating plug within the female atrium following mating is considered a de facto marker of successful copulation given its relevance for sperm storage 21,22,32. To facilitate measuring male mating success in vivo, we developed a novel transgenic line, Plugin-tdTomato (PluTo), for use as a marker for the occurrence of insemination in females. PluTo males express a Plugin-tdTomato fusion protein specifically in their MAGs via the Plugin promoter, without negatively affecting endogenous Plugin levels. We show that Plugin-tdTomato is incorporated into the mating plug and is transferred to the female during mating, where it is detectable by microscopic examination after copulation. Importantly, mating experiments in competition with wild type males suggest that male mating competitiveness is not affected by expression

130 of this transgene. Moreover, transfer of this fusion protein does not affect the reproductive fitness of females mated to PluTo males. These data demonstrate that this transgenic construct could be useful for determining the mating competitiveness of A. gambiae strains, as well as of other anopheline species that transfer a mating plug, in semi-field and field conditions.

RESULTS

4.1 Generation and characterization of the plugin-tdTomato transgenic line.

Plugin is specifically and abundantly expressed within the MAGs21,33. During mating, this seminal protein is transferred to females as part of the mating plug, a gelatinous structure that is composed of multiple other proteins as well as the steroid hormone 20E21. To create a transgenic line exhibiting robust MAG- specific expression, we cloned a region comprising 2,688 bp to drive expression of tdTomato 34 upstream of the Plugin start codon, likely encompassing the Plugin promoter. We then used this region fused to a

Plugin cDNA construct21 to achieve incorporation into the mating plug for transfer to females. The two protein-coding regions of the fusion construct were connected by a 6-serine residue linker (reviewed in

35) (Figure 4.1A). The transgenic cassette was cloned into the pDSAY transgenesis plasmid36 and was injected into A. gambiae embryos from the X1 docking line which contains an attP site for φC31-mediated integration site on chromosome 2L36, in conjunction with an integrase-expressing helper plasmid36. A single transgenic line was generated and named Plugin-Tomato (PluTo). Transgene insertion into the X1 docking site was confirmed by PCR.

131

Figure 4.1 | PluTo transgene design and male phenotype. [A] The PluTo transgene is composed of 2688 bp of the 5’ regulatory region immediately upstream of the start codon of AGAP009368 followed by Plugin cDNA fused to a 6-serine linker followed by tdTomato., with 3xP3-EYFP for selection. Each coding sequence has an SV40 terminator to prevent transcriptional read-through. Following insertion, the flanking recombination sites attL and attR are generated in the genome [B] Adult PluTo transgenic males display strong tdTomato fluorescence in their MAGs (lower panels). The fluorescence is clearly and easily identifiable under a fluorescent dissecting microscope through the adult male cuticle (upper panels). Scale bar – 100 µm.

This transgenic line displayed strong tdTomato fluorescence in the MAGs, which was visible through the male cuticle beginning at the early adult stage (Figure 4.1B). No discernable tdTomato fluorescence was observed in other male or female tissues. In agreement with our microscopic analysis, qRT-PCR demonstrated that the Plugin promoter was capable of robust Plugin-tdTomato expression in the MAGs comparable to endogenous Plugin expression, with negligible transcripts observed in the male rest-of- body and female whole-body tissues (Figure 4.2A). Moreover, both endogenous Plugin and Plugin- tdTomato fusion proteins were observed exclusively in the MAGs at ~80kDA and ~150kDA respectively, using antibodies targeting Plugin and tdTomato (Figure 4.2B, bands 2 and 4 respectively)21. Several tdTomato-specific bands were observed not recognized by α-Plugin, suggesting of cleavage of the fusion protein or use of an alternative start codon after the antibody binding site (Figure 4.2B, band 3).

132

Figure 4.2 | Endogenous Plugin and PluTo transgene expression. [A] Quantitative qRT-PCR of endogenous Plugin and Plugin-tdTomato expression levels in male and female PluTo and G3 pupae and adult tissues over time. An age of zero days refers to uneclosed pupae (not dissected) and ages 1 and 2 are minimal ages in days post-eclosion (MAGs were dissected out of males). In MAGs, transgenic Plugin-tdTomato was induced similarly to endogenous Plugin levels, and age significantly affected expression. Figure 4.2 (Continued) In male rest of body tissues, transgenic Plugin-tdTomato was expressed at negligible, although higher, levels at ages 1 and 2 days. In female whole body tissues, transgenic Plugin-tdTomato was also expressed at negligible levels. Endogenous Plugin levels could not be reliably detected in wild type females aged 1 or 2 days. Analysis of variance followed by post-hoc testing within ages showed the following significant differences:- MAGs (left panel): ANOVA (Age F(2,15) = 16.1, p = 0.0002, Genotype-Primer combination F(2,15) = 2.7, p = 0.099, Interaction F(4,15) = 1.92, p = 0.160). Tukey’s post-hoc test: Age 1 – PluTo Plugin-tdTomato vs G3 Plugin (mean difference ± SEM = 43.3 ± 16.63), adj. p value = 0.0359. Male Rest of body (center panel) ANOVA (Genotype-Primer combination F(2,11) = 12.3, p = 0.0015, Age F(1,11) = 1.99, p = 0.186, Interaction F(2,11) = 0.53, p = 0.601). Tukey’s post-hoc tests: Age 1 – PluTo Plugin-tdTomato vs PluTo Plugin (mean difference ± SEM = 0.021 ± 0.005), adj. p value = 0.0067; Age 2 – PluTo Plugin-tdTomato vs PluTo Plugin (mean difference ± SEM = 0.015 ± 0.005), adj. p value = 0.0406; Age 2 – PluTo Plugin-tdTomato vs G3 Plugin (mean difference ± SEM = 0.015 ± 0.005), adj. p value = 0.037. The downward error bar for Age 1 G3 Plugin cannot be plotted on a logarithmic axis as it extends below 0. Female Rest of body (right panel): ANOVA (Genotype- Primer combination F(2,16) = 4.88, p = 0.022, Age F(2,16) = 0.638, p = 0.541, Interaction F(4,16) = 0.6, p = 0.668). Tukey’s post-hoc tests: no pairwise comparisons were detected with adj. p value < 0.05. [B] Western blot of endogenous Plugin and Plugin-tdTomato in adult male and female tissues with α-Plugin in green, and α-tdTomato in red. Endogenous Plugin is observed at similar levels in the MAGs of both wild type G3 and PluTo transgenics at ~80 kDA (band 2). The α-Plugin antibody also recognizes an additional band at 27 kDA (band 1), which may be non- specific or cleaved protein. Full length transgenic Plugin-tdTomato is observed at ~150 kDA recognized by both α- Plugin antibody and α-tdTomato (band 4). An additional Plugin-tdTomato species at 120 kDA is recognized exclusively by the α-tdTomato antibody (band 3), which may represent cleaved protein, or result from an alternative start codon excluding the α-Plugin antibody binding site. Multiple high molecular weight bands (bands 6) are observed as dimers and multimers of Plugin and Plugin-tdTomato species in wild type G3 and PluTo males, respectively. Detectable Plugin and Plugin-tdTomato protein are not observed in male rest-of-body nor female whole-body tissues. The α-tdTomato antibody recognizes unspecific low molecular weight moieties in these tissues (bands 5).

133

4.2 Plugin-tdTomato localizes to vesicles within the anterior compartment of the MAGs

To determine the expression of Plugin-tdTomato in more detail and compare it to endogenous Plugin, we first characterized the localization of Plugin in the MAGs of wild type mosquitoes by cryo-immune electron microscopy (IEM) using an α-Plugin antibody 21. In Anopheles, the MAGs are composed of a thin epithelial sheath encasing two secretory compartments, a large anterior compartment and a smaller posterior compartment, each containing a single layer of holocrine secretory cells that produce MAG substances of different electron densities 37. This global organization was observed in our IEM analysis (Figure 4.3A), in which we detected multiple Plugin-positive electron-dense vesicles in cells of the anterior compartment only (Figure 4.3B–E).

Figure 4.3 | Endogenous Plugin localization in the MAGs. [A] Transmission electron micrograph of the boundary between anterior (ant) and posterior (post) MAG compartments in wild type males. Red box indicates area magnified in panel B, white box indicates area magnified in panel D. Inset image provided for orientation within MAGs and approximate area magnified. Scale bar 5 µm. [B] The anterior compartment has peripherally located nuclei (nuc) and contains electron-dense vesicles that fill the lumen (lum). The central aedeagus (aed) is marked for orientation. White box indicates area magnified in C. Scale bar 2 µm. [C] Plugin is localized within the electron- dense secretory vesicles by gold-labelled α-Plugin (white arrowheads). Scale bar 0.5 µm. [D] Vesicles within the anterior (ant) and posterior (post) compartments are different in size and electron density, suggesting different contents. White box indicates area magnified in E. Scale bar 2 µm. [E] Plugin is localized only within the more electron-dense secretory vesicles of the anterior compartment by gold-labelled α-Plugin (white arrowheads). Scale bar 0.5 µm.

134

We then performed live fluorescence microscopy and immunofluorescence analysis (IFA) using the α-

Plugin antibody on PluTo MAGs. Plugin-tdTomato was localized to vesicles within the anterior MAG compartment (Figure 4.4A, B), similar to Plugin in the cryo-EM. Vesicles remained fluorescent as they moved into the interior of the anterior compartment, away from the epithelium. In the IFA, Plugin was localized to channels formed by a muscle network on the outside of the MAGs (Figure 4.4A, B), as described previously21. It was not detected in vesicles in the interior of the glands, except when MAG epithelium integrity was disrupted (Figure 4.4B), suggesting incomplete penetration of the Plugin antibody, and/or the absence of regulatory elements affecting the precise localization pattern of Plugin- tdTomato within the anterior compartment. No Plugin-tdTomato or Plugin staining was detected in the posterior compartment or in the aedeagus.

Figure 4.4 | Plugin-tdTomato localization in the MAGs and mating plug. In all panels, Plugin-tdTomato fluorescence is shown in red, Plugin in green, and DNA stained with DAPI in blue. [A] Plugin-tdTomato-filled vesicles occurred throughout the secretory cells of the anterior MAG compartment (ant) but was not detected in the posterior compartment (post) or aedeagus (aed). Stacked Z-projection of the top 3.5 µm. The box in the merge panel represents the region shown in B. Scale bar – 100 µm. [B] Close-up of Plugin and Plugin-tdTomato where MAG epithelial integrity has been disrupted manually. Some Plugin-tdTomato vesicles are co-stained with Plugin (arrowheads). Stacked Z-projection of the top 3.5 µm. Scale bar – 20 µm. [C] The female reproductive tract, consisting of the ovaries (Ov), atrium (At), and the sperm storage organ spermatheca (Sp) immediately after mating. The red fluorescent mating plug is visible within the atrium. Scale bar – 100 µm. [D] A single Z-slice of the mating plug within the female atrium immediately post-mating. Plugin-tdTomato is visible throughout the mating plug, except the tip. Scale bar – 50 µm.

135

4.3 Plugin-tdTomato is sexually transferred to the female atrium in the mating plug

To determine if Plugin-tdTomato was incorporated into the plug and transferred to females, we performed forced-mating assays22,38 with PluTo males and wild type females, immediately followed by fluorescence imaging. We observed significant tdTomato fluorescence highlighting the characteristic structure of the mating plug within the female atrium (Figure 4.4C). Dissection of the female reproductive tract immediately after mating identified a strong fluorescent signal specifically along the length of the mating plug, except the proximal tip, demonstrating that the fusion protein had been incorporated in the plug and transferred during copulation (Figure 4.4D).

4.4 Testing PluTo males for in vivo assessment of successful mating

We next performed a time course analysis to determine whether sexual transfer of a mating plug containing the Plugin-tdTomato fusion could be reliably detected without sacrificing the female. Females mated to PluTo males were analyzed for mating plug fluorescence through the female cuticle at different time points after mating. As previously determined39, the plug shows naturally-occurring autofluorescence under a filter for green fluorescent protein, but this signal, likely derived from plug lipids, is only reliably detected for 2 hours post mating (Figure 4.5). However, in females mated to PluTo males, strong, sharp fluorescence was clearly visible under a red filter for at least 8 hours after mating, demonstrating that this fusion protein can be used to reliably determine the occurrence of successful insemination in a non-invasive manner for a significantly longer span of time (Figure 4.5). We did not systematically test later time points as the plug becomes digested and difficult to detect reliably by fluorescence.

136

Figure 4.5 | PluTo transgenic males allow in vivo assessment of male mating success within the female. Following copulation, the tdTomato-labeled mating plug is distinctly visible through the female cuticle. Autofluorescence is also visible through a wide pass GFP filter set. Mating plug autofluorescence decreases over 4 hours following copulation whereas Plugin-tdTomato fluorescence remains visible for up to 8 hours. Exposure times are indicated below. Scale bar – 100 µm.

4.5 Expression of the plugin-tdTomato fusion protein does not affect the mating competitiveness of

PluTo males and does not impair female reproductive fitness

Given our ability to detect a fluorescent mating plug in mated female for at least 8 hours after copulation, we reasoned that the Plugin-tdTomato fusion protein could be used to assay male mating competitiveness, an important aspect determining the potential success of genetic control programs based on male releases. We therefore assessed whether expression of the transgene affected the mating fitness of PluTo males by performing mating capture assays where 100 PluTo and 100 wild type G3 males were released in cages with 100 wild type virgin females, and mating couples were captured while in copula. Fluorescence analysis of 73 couples collected across 4 replicate cages showed no difference in the mating competitiveness of the two groups of males (p = 0.3492, two-tailed binomial test), with 44%

137 females inseminated by PluTo individuals. Analysis of the remaining females that were not caught in copula but were left in the cage with males confirmed this result, with 33 of 69 mated females (48%) showing the PluTo transgene in sperm stored in their spermathecae (p = 0.810, two-tailed binomial test).

Moreover, in separate experiments, females mated to PluTo males had comparable reproductive fitness as females mated to wild type G3 males, measured as the number of eggs developed after blood feeding, and the rates of oviposition and infertility of the broods (Figure 4.6A–C). Taken together, these results show that PluTo males induce normal post-mating behavior in females with which they mate, suggesting that Plugin-tdTomato may be a suitable construct for evaluating fitness in mosquito lines for semi-field and field releases.

Figure 4.6 | Reproductive fitness of females mated to PluTo males. [A-C] Wild type females mated to either wild type or PluTo males were scored for eggs developed [A], egg laying behavior [B], and the rate of infertility within broods [C]. In A and C, each dot indicates the brood of an individual female. No significant differences were observed between females in the two groups (A: Student’s t-test on square root-transformed data, p = 0.38; B: Fisher’s exact test, p = 0.189; C: Mann Whitney U test p = 0.187).

138

DISCUSSION

The control of the Anopheles mosquito via the use of LLINs or IRS is key to malaria control strategies and has contributed considerably to the reduction in number of cases and deaths witnessed since the beginning of the century 40. However, the effectiveness of our best vector control tools is jeopardized by the rapid insurgence and spread of insecticide resistance in Anopheles populations, incentivizing the generation of novel methods of genetic control. These methods rely on the release of males that have been modified to express genetic traits that either induce sterility in their female mates, or confer resistance to Plasmodium parasites that cause malaria. Some past attempts to release chemical- and radiation- sterilized anophelines reportedly reduced male mating competitiveness 12, which limits the ability of males to sterilize females or to transmit desired genetic traits. While release of sterile males works effectively in other insects 8-10, the intricate mating biology of many anophelines including the major

Afrotropical vectors A. gambiae, A. coluzzii, A. arabiensis and A. funestus will make developing vector control programs relying on male release difficult in these species. 11,12,41,42. These species indeed are characterized by copulations in mating swarms, which are formed every night at dusk by large groups of males. Given the highly skewed sex ratio within these swarms, competition between males for the many fewer females is fierce (reviewed in 43) and a number of factors ranging from the desired genetic manipulations induced in the laboratory to the colonization process and mass rearing, can negatively impact male mating success. Determining the mating competitiveness of males prior to a release is therefore essential for the success of future genetic control efforts.

Currently there are no straightforward methods to determine the mating fitness of released males without sacrificing the female. Available methods necessitate dissections of females shortly after copulation to assess the presence of sperm in the spermatheca and the mating plug in the atrium. While the mating plug is slightly autofluorescent, we show here that this signal can only be reliably detected for

139 a short period of time after copulation. To achieve non-invasive determination of male mating success in competition assays, the only option at present is to blood feed mated females, allow them to oviposit, and screen their progeny for specific genetic traits of the relevant male to determine progeny paternity, while keeping the males alive and in isolation if needed. Our study provides an easy, non-invasive and effective method to assess male mating success in a short period of time and without the need to kill the female. As an example, mating plug-labeled males could be released in a swarm (whether natural or in semi-field cages), and after capturing mating couples, females could be fluorescently analyzed for the occurrence of insemination. Although this method relies on microscopes equipped with fluorescence, which are not likely to be available in remote field settings, the possibility to detect plugs for 8 hours after copulation would allow identification of transgenic matings back in the laboratory. Moreover, although in this study we only tested live females after mating, it is possible that fluorescence will also persist in dead females, such as those captured in light traps, as the fluorescent signal of transgenic lines can generally be observed for a few hours after death.

The Plugin upstream regulatory region used in this study was sufficient to achieve MAG-specific expression of our transgene. Such a promoter will facilitate the study of seminal secretions produced by the male glands, for instance via the generation of transgenic RNA interference (RNAi) lines stably expressing double-stranded RNAs (dsRNAs) against targets of interest. Expressing dsRNA transgenes will help to circumvent the limited silencing efficiency achieved by transient injections of dsRNA molecules targeting genes expressed in the MAGs 21,22, allowing detailed functional analyses.

Finally, by fusing a fluorescent protein to the coding region of Plugin, we ensured incorporation of the marker into the mating plug and its transfer to the female. This system therefore allows for efficient transfer of desired factors from the male to the female, and their slow release once in the female atrium.

Interestingly, this property could be utilized to deliver female-specific sterilizing compounds or other factors that may reduce the reproductive fitness of mated females. Although such factors are not readily

140 available, future optimization of the system may afford specific activation of sterilants or toxins only after delivery to the female reproductive tract.

CONCLUSIONS

In conclusion, this study generates a valuable tool for assessing the mating competitiveness of males in semi-field and field studies as well as for laboratory experiments aimed at determining the function of

MAG seminal secretions. This tool may prove invaluable when testing the feasibility of releasing genetically modified mosquitoes for the control of malaria-transmitting Anopheles populations.

Importantly, given the observation that males from the most important Afrotropical anopheline species transfer mating plugs during copulation 24,44, this system will be readily transferable to other species like

A. arabiensis and A. funestus, key vectors for Plasmodium falciparum transmission in sub-Saharan Africa.

METHODS

Plasmid construction

To clone the PluTo transgene, we separately cloned the promoter, Plugin coding sequence (AGAP009368), and tdTomato with 6S linker into the pDSAY transgenesis vector backbone 36 by standard Golden Gate cloning methods 45. We amplified 2,688 bp of the region upstream of the Plugin start codon from genomic

DNA isolated from wild type A. gambiae (G3 strain) by PCR using GGpP FWD [5’

CAGGTCTCAATCCTTGTAGGGCTTGTTGACGGG 3’] and GGpP REV [5’

CAGGTCTCATCATGTCTACGGTTGAATCAGTGATACAAGCAAA 3’] primers to provide a Golden Gate compatible overhang for the pDSAY destination vector. We amplified the Plugin coding sequence from wild type G3 male abdominal cDNA extracts by PCR using GGcP FWD [5’

141

CCTCTCAATGAAGGCTTTGGTAGCTCTGCTCTG 3’] and GGcP REV [5’

CAGGTCTCACCTTCGACGACCAGCACA3’] to remove the stop codon, and fuse seamlessly to the Plugin promoter fragment. tdTomato was PCR amplified from a plasmid provided by the Tsien Lab 34 using GGtdT

FWD [5’ CAGGTCTCATCCAGCTCCTCCTCCTCCATGGTGAGCAAGGGCGAGG 3’] and GGtdT REV [5’

CAGGTCTCAAAGCTTACTTGTACAGCTCGTCCATGCC3’] primers to add a 6-serine linker and provide a

Golden Gate compatible overhang to facilitate cloning into pDSAY. These fragments were cloned into the pDSAY φC31 transgenesis vector backbone 36, which provides an SV40 terminator at the 3’ end of Plugin- tdTomato as well as a selectable marker cassette (3xP3-EYFP) to detect transgenic individuals (Figure 1A).

Transgenesis and PCR confirmation

Embryo microinjections were performed essentially as described 46,47 with additional co-injection of a

Vasa-φC31 integrase helper plasmid 36 (350 ng/µl and 80 ng/µl of transgenesis and helper plasmids, respectively). A total of 996 X1 docking line embryos were injected, with 172 embryos surviving to hatching. Following outcrossing of the injected P0 individuals to wild type G3, a single F1 transgenic PluTo male was isolated and outcrossed to G3 females. F2 transgenic heterozygotes were intercrossed to yield

F3 homozygotes. Homozygotes were identified by strong YFP fluorescence intensity and isolated.

Transgene insertion into X1 was confirmed by PCR on three distinct PluTo males and three unintegrated

X1 docking line males using the NEB Q5® High-Fidelity 2 X Master Mix kit. Whole body mosquitoes were collected and DNA prepared using the Qiagen DNeasy Blood & Tissue Kit. Transgene insertion was confirmed by PCR with EYFP REV [5’GTCGTCCTTGAAGAAGATGGTG 3’] and X1 FWD [5’

AGGGAAGATTGGAATCCATC 3’]. Control PCRs for the empty and unintegrated docking site and S7 controls were performed with X1 FWD [5’ AGGGAAGATTGGAATCCATC 3’], X1 REFV [5’ ACTGCAACCCATTCACACTG

3’], and S7 FWD [5’ GGCGATCATCATCTACGTGC 3’], S7 REV [5’ GTAGCTGCTGCAAACTTCGG 3’], respectively.

142

Mosquito maintenance

A. gambiae mosquitoes (wild type G3 and PluTo transgenic strains) were reared in cages at 28 °C and 70

% humidity on a 12 h/12 h cycle. Adult mosquitoes were fed a 10 % (wt/vol) glucose solution ad libitum and were given a human blood meal once a week for the purposes of line maintenance. For colony cages, adults were kept in mixed sex cages for up to two weeks post eclosion and one-week post egg-lay. For experiments, males and females were separated as pupae to ensure the virgin status of females and kept in separate cages as experimental conditions demanded. Mating couples were collected either in copula, or though forced mating assays (protocol available on https://www.beiresources.org/Publications/MethodsinAnophelesResearch.aspx) and mating was confirmed via mating plug autofluorescence or quantitative PCR (below).

Tissue sample collection for qRT-PCR

Male and female G3 and PluTo mosquitoes were sexed and separated into four cages. From each of these cages, sample groups were sacrificed at 24 h intervals from pupation (d0) until two days post eclosion

(d2). To test whether Plugin or Plugin-tdTomato expression is constrained to the male accessory glands

(MAGs), MAG samples (n = 10) were collected from eclosed males and compared to the rest-of-body (ROB)

(n = 5) while female eclosed mosquitoes were collected as whole body samples (n = 5). ROB samples were collected in 200 µL of RNAlater® (Sigma-Aldrich), while MAG samples were transferred to 15 µL of

RNAlater using a needle. All samples were stored at –20 ˚C until specimen collection was completed. Three replicates were performed.

RNA extraction and cDNA synthesis

ROB samples were defrosted on ice, RNAlater was removed and replaced with 200 μL TRI Reagent®

(Applied Biosystems), while 185 µL TRI reagent was added to MAG samples to dilute out the RNA-later and avoid tissue loss. RNA was extracted according to the manufacturer’s instructions. RNA was quantified

143 by using a NanoDrop Spectrophotometer (Thermo Scientific) and DNase treated using Turbo DNAse

(Thermo-Fisher). cDNA synthesis was performed as in 48. Approximately 4 µg of RNA was used in 100 μL cDNA reactions. Reactions were diluted to 200 μL with nuclease-free water for storage at –20 °C.

Quantitative RT-PCR

Samples for quantitative RT-PCR were diluted tenfold and quantified in triplicate using standard curves.

PCRs were run in Fast SYBR Green Master Mix (Thermo-Fisher) on a Step One Plus thermocycler (Applied

Biosystems). Endogenous Plugin cDNA was amplified using 9368_FWD [5’ TGATTCAACCGTAGACATGAAGG

3’] and 9368_REV [5’ CCACCATACAACGGAACGAC 3’] primers. Transgenic Plugin-tdTomato was amplified using qPluginTOM_F [5’ ATCTCAACAGGAGCCCAATG3’] and qPluginTOM_R [5’ CCCTTGCTCACCATGGAG

3’]. Quantities were normalized against the ribosomal protein RpL19 using previously described primers

48.

Western blotting

Male accessory glands (n = 10) and ROB or whole body (n = 5) samples from 7-day-old PluTo and G3 age- matched males and females were collected by dissection on ice. Samples were homogenized on ice in protein extraction buffer (25 mM Tris HCl pH 7.4, 150 mM NaCl, 0.1 % SDS, 1 % Triton X-100, 10 mM EDTA pH 8, 1 X protease inhibitor, 1 % phosphatase inhibitor) and quantified using a Bradford assay. Protein samples were run at 150 V and 250 mA for 1.5 h on an XCell SureLock vertical system in a 4–12 % Bis-Tris

NuPAGE gel (Thermo-Fisher). Protein was then transferred to PVDF using iBLOT2 transfer system (Thermo-

Fisher). After transfer, membrane was washed twice in 1 X PBS-T (0.1 % TWEEN in 1 X PBS) and blocked in an automated shaker for 1 h at room temperature in Odyssey® Blocking Buffer (PBS) (Li-cor). Membrane was then stained with goat -tdTomato (Sicgen) and rabbit -Plugin (GenScript Corp; 21) (1:1000 concentrations for both antibodies) on an automated shaker at 4 ˚C overnight. Membrane was washed with PBS-T, and stained with secondary antibodies (donkey -rabbit 800 and donkey -goat 680 (Li-cor),

144 both at 1:10,000) in blocking buffer with 0.01 % SDS for 1 h at room temperature. Membrane was washed thoroughly before imaging with the Odyssey® CLx imager. Images were analyzed in StudioImageLite (Li- cor).

Cryo-immune Electron Microscopy

Virgin 4-day-old male reproductive tracts (including the last three abdominal segments) were fixed overnight at 4 ˚C (4 % paraformaldehyde, 0.1 % glutaraldehyde in 0.1 M sodium phosphate buffer at pH

7.4). Samples were washed with 1 X PBS, and infiltrated with 2.3 M sucrose in 1 X PBS, 0.2 M glycine for

15 min. Tissues were mounted on a pin and frozen in liquid nitrogen prior to sectioning. Samples were sectioned by cryo-Ultra Microtome (Reichert-Jung, Ultracut)(-120 ˚C) to obtain thick sections (0.5 μm) for visual screening with toluidine blue, or ultrathin sections (70–80 nm) for target tissue analysis. Ultrathin sections were transferred to carbon-coated copper grids, blocked in 1 X PBS, 1 % BSA for 10 min and stained 30 min with primary antibody at RT (1:30 dilutions of rabbit -Plugin). The grids were washed in

1 X PBS for 15 min and then labelled with Protein-A gold (Sigma-Aldrich, 15 nm diameter) diluted in blocking solution for 20 min, following a final wash in distilled water for 15 min. Contrasting of the labelled grids was carried out on ice in 0.3 % uranyl acetate in 2 % methyl cellulose for 10 min. All micrographs were captured using a JEOL 1200EX 80 kV electron microscope and recorded with an AMT 2k CCD camera.

Three biological replicates were performed, and a representative selection are shown in the text. Samples were prepared by the Harvard Medical School EM core facility.

Fluorescence imaging

MAGs were collected from 5–7 day-old virgin PluTo males and fixed in 4 % paraformaldehyde solution for

1 h, then rinsed 1 X PBS. Tissues were soaked in 80 % ethanol at 4 ˚C for 3 min to remove lipids, rinsed 3 times in 1 X PBS, bleached in 3 % hydrogen peroxide for 5 min to quench endogenous autofluorescence, and rinsed in 1 X PBS 3 times. Samples were blocked and permeabilized overnight in blocking solution (1

145

% BSA, 0.3 % Tween-20 in 1 X PBS), then stained with rabbit -Plugin (1:200) for approximately 12 h.

Tissues were washed 5 times in 1 X PBS followed by a 30 min incubation in blocking solution. Tissues were then stained for 1 h at room temperature with secondary antibody (Alexafluor 647 Donkey anti-Rabbit

(1:1000) in blocking solution, rinsed in blocking solution, then stained with DAPI (5 µg/ml) in blocking solution for 10 min. Tissues were washed at room temperature 5 times with blocking solution, once with

PBS-T, and twice with 1 X PBS. Tissues were then mounted in Vectashield® Mounting Medium (Vector

Laboratories) and imaged using an upright SPE confocal microscope (Harvard NeuroDiscovery Center Leica

SPE). Live unfixed samples were dissected on ice in 1 X PBS and imaged immediately on a Leica dissecting microscope with fluorescence.

Mating plug digestion and reproductive fitness assays

Four-day-old virgin G3 females were forced mated to four-day-old PluTo or G3 males and screened for mating plugs. Females without a visible mating plug were excluded from the study. Females force mated to PluTo males were then further divided into 4 groups of 10 individuals and each group imaged under fluorescent lighting at t = 0.5, 1, 1.5, 2, 4, 6, and 8 h post mating. After imaging, females were blood fed, placed in isolated oviposition cups and given 4 d to lay eggs, after which total numbers of eggs were tallied and infertility was scored.

Mating Competitiveness Assay

Reproductive behavioral effects on male mating efficacy resulting from our transgene were evaluated using a mating competitiveness assay where 100 G3 and 100 PluTo males (4–5-day-old) were allowed to compete and mate freely for 100 G3 females (4-day-old) in an open cage environment. Couples were caught in copula for 1 h as previously described 48, and male genotype determined by fluorescence of the

PluTo transgene. Uncaptured females were left in the cage for 24 h, after which females were collected and individually processed for DNA by brief homogenization in extraction buffer (0.12 % Tris-Cl, 0.037 %

146

EDTA, 0.29 % NaCl). Mating status and male genotype were determined using a modified qPCR protocol using both Y-specific primers YQPCR_FWD [5’ GGATCTGGCCAAGAGGAGTA 3’], YQPCR_REV [5’

CCCAACCAAGGTACTCTAACG3’], and tdTomato primers Q485 [5’ TGGAGTTCAAGACCATCTAC 3’], Q486 [5’

GTGTCCACGTAGTAGTAGCC3’].

Statistical analysis

Data were analyzed using GraphPad Prism 6.0, with statistical tests used indicated in the figure legends.

Quantitative RT-PCR: since expression of Plugin-tdTomato was not detectable in the G3 background, precluding a full multi-factorial analysis (genotype x gene x time), we analyzed Plugin and Plugin-tdTomato expression levels using two-way analysis of variance incorporating age (0, 1, and 2 d) and genotype-primer combination (G3-Plugin, PluTo-Plugin, PluTo-Plugin-tdTomato) as factors, followed by Tukey’s multiple comparison test, with adjusted p-values for multiple testing.

147

REFERENCES

1 Smidler, A. L., Scott, S. N., Mameli, E., Shaw, W. R. & Catteruccia, F. A transgenic tool to assess Anopheles mating competitiveness in the field. Parasit Vectors 11, 651, doi:10.1186/s13071-018- 3218-5 (2018).

2 WHO. World Malaria Report. 196 (World Health Organization, Geneva, 2017).

3 Ranson, H. & Lissenden, N. Insecticide Resistance in African Anopheles Mosquitoes: A Worsening Situation that Needs Urgent Action to Maintain Malaria Control. Trends Parasitol 32, 187-196, doi:10.1016/j.pt.2015.11.010 (2016).

4 Knox, T. B. et al. An online tool for mapping insecticide resistance in major Anopheles vectors of human malaria parasites and review of resistance status for the Afrotropical region. Parasites & Vectors 7, 76, doi:10.1186/1756-3305-7-76 (2014).

5 Edi, C. V., Koudou, B. G., Jones, C. M., Weetman, D. & Ranson, H. Multiple-insecticide resistance in Anopheles gambiae mosquitoes, Southern Cote d'Ivoire. Emerging infectious diseases 18, 1508- 1511, doi:10.3201/eid1809.120262 (2012).

6 Cisse, M. B. et al. Characterizing the insecticide resistance of Anopheles gambiae in Mali. Malaria journal 14, 327, doi:10.1186/s12936-015-0847-4 (2015).

7 Knipling, E. F. Sterile-male method of population control. Science 130, 902-904 (1959).

8 Krafsur, E. S., Whitten, C. J. & Novy, J. E. Screwworm eradication in North and Central America. Parasitol Today 3, 131-137 (1987).

9 Lindquist, D. A., Abusowa, M. & Hall, M. J. The New World screwworm fly in Libya: a review of its introduction and eradication. Medical and Veterinary Entomology 6, 2-8 (1992).

10 Hendrichs, J., Franz, G. & Rendon, P. Increased effectiveness and applicability of the sterile insect technique through male‐only releases for control of Mediterranean fruit flies during fruiting seasons. Journal of Applied Entomology 119, 371-377 (1995).

11 Abdel-Malek, A., Tantawy, A. & Wakid, A. Studies on the Eradication of Anopheles pharoensis by the Sterile-Male Technique Using Cobalt-60. III. Determination of the Sterile Dose and Its Biological Effects on Different Characters Related to‟ Fitness” Components. Journal of economic entomology 60, 20-23 (1967).

12 Sharma, V., Razdan, R. & Ansari, M. Anopheles stephensi: effect of gamma-radiation and chemosterilants on the fertility and fitness of males for sterile male releases. Journal of economic entomology 71, 449-452 (1978).

13 Reisen, W. K. Lessons from the past: an overview of studies by the University of Maryland and the University of California, Berkeley. Ecological aspects for application of genetically modified mosquitoes, 25-32 (2003).

148

14 Ageep, T. B. et al. Participation of irradiated Anopheles arabiensis males in swarms following field release in Sudan. Malaria journal 13, 484, doi:10.1186/1475-2875-13-484 (2014).

15 Munhenga, G. et al. Mating competitiveness of sterile genetic sexing strain males (GAMA) under laboratory and semi-field conditions: Steps towards the use of the Sterile Insect Technique to control the major malaria vector Anopheles arabiensis in South Africa. Parasites & Vectors 9, 122, doi:10.1186/s13071-016-1385-9 (2016).

16 Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nature biotechnology 34, 78-83, doi:10.1038/nbt.3439 (2015).

17 Hammond, A. M. et al. The creation and selection of mutations resistant to a gene drive over multiple generations in the malaria mosquito. PLoS genetics 13, e1007039, doi:10.1371/journal.pgen.1007039 (2017).

18 Giglioli, M. E. C. & Mason, G. F. The mating plug in anopheline mosquitoes. Proceedings of the Royal Entomological Society of London. Series A, General Entomology 41, 123-129, doi:10.1111/j.1365-3032.1966.tb00355.x (1966).

19 Baldini, F., Gabrieli, P., Rogers, D. W. & Catteruccia, F. Function and composition of male accessory gland secretions in Anopheles gambiae: a comparison with other insect vectors of infectious diseases. Pathogens and global health 106, 82-93, doi:10.1179/2047773212Y.0000000016 (2012).

20 Pondeville, E., Maria, A., Jacques, J.-C., Bourgouin, C. & Dauphin-Villemant, C. Anopheles gambiae males produce and transfer the vitellogenic steroid hormone 20-hydroxyecdysone to females during mating. Proceedings of the National Academy of Sciences 105, 19631-19636, doi:10.1073/pnas.0809264105 (2008).

21 Rogers, D. W. et al. Transglutaminase-mediated semen coagulation controls sperm storage in the malaria mosquito. PLoS biology 7, e1000272, doi:10.1371/journal.pbio.1000272 (2009).

22 Gabrieli, P. et al. Sexual transfer of the steroid hormone 20E induces the postmating switch in Anopheles gambiae. Proceedings of the National Academy of Sciences of the United States of America 111, 16353-16358, doi:10.1073/pnas.1410488111 (2014).

23 Baldini, F. et al. The Interaction between a Sexually Transferred Steroid Hormone and a Female Protein Regulates Oogenesis in the Malaria Mosquito Anopheles gambiae. PLoS biology 11, e1001695, doi:10.1371/journal.pbio.1001695 (2013).

24 Mitchell, S. N. et al. Mosquito biology. Evolution of sexual traits influencing vectorial capacity in anopheline mosquitoes. Science (New York, N.Y.) 347, 985-988, doi:10.1126/science.1259435 (2015).

25 Shaw, W. R. et al. Mating activates the peroxidase HPX15 in the sperm storage organ to ensure fertility in Anopheles gambiae. Proc Natl Acad Sci U S A 111, 5854-5859, doi:10.1073/pnas.1401715111 (2014).

149

26 Le, B. V. et al. Characterization of Anopheles gambiae Transglutaminase 3 (AgTG3) and Its Native Substrate Plugin. Journal of Biological Chemistry 288, 4844-4853, doi:10.1074/jbc.m112.435347 (2013).

27 Ravi Ram, K., Sirot, L. K. & Wolfner, M. F. Predicted seminal astacin-like protease is required for processing of reproductive proteins in Drosophila melanogaster. Proceedings of the National Academy of Sciences of the United States of America 103, 18674-18679, doi:10.1073/pnas.0606228103 (2006).

28 Park, M. & Wolfner, M. F. Male and female cooperate in the prohormone-like processing of a Drosophila melanogaster . Developmental biology 171, 694-702, doi:10.1006/dbio.1995.1315 (1995).

29 LaFlamme, B. A., Ram, K. R. & Wolfner, M. F. The Drosophila melanogaster seminal fluid protease "seminase" regulates proteolytic and post-mating reproductive processes. PLoS genetics 8, e1002435, doi:10.1371/journal.pgen.1002435 (2012).

30 Murer, V. et al. Male fertility defects in mice lacking the serine protease inhibitor protease nexin- 1. Proceedings of the National Academy of Sciences of the United States of America 98, 3029-3033, doi:10.1073/pnas.051630698 (2001).

31 Yamaguchi, A. et al. Identification and localization of the sperm CRISP family protein CiUrabin involved in gamete interaction in the ascidian Ciona intestinalis. Molecular reproduction and development 78, 488-497, doi:10.1002/mrd.21329 (2011).

32 Thailayil, J., Magnusson, K., Godfray, H., Crisanti, A. & Catteruccia, F. Spermless males elicit large- scale female responses to mating in the malaria mosquito Anopheles gambiae. Proceedings of the National Academy of Sciences 108, doi:10.1073/pnas.1104738108 (2011).

33 Giraldo-Calderon, G. I. et al. VectorBase: an updated bioinformatics resource for invertebrate vectors and other organisms related with human diseases. Nucleic acids research 43, D707-713, doi:10.1093/nar/gku1117 (2015).

34 Shaner, N. C. et al. Improved monomeric red, orange and yellow fluorescent proteins derived from Discosoma sp. red fluorescent protein. Nature biotechnology 22, 1567-1572, doi:10.1038/nbt1037 (2004).

35 Chen, X., Zaro, J. L. & Shen, W. C. Fusion protein linkers: property, design and functionality. Advanced drug delivery reviews 65, 1357-1369, doi:10.1016/j.addr.2012.09.039 (2013).

36 Volohonsky, G. et al. Tools for Anopheles gambiae Transgenesis. G3 (Bethesda, Md.), doi:10.1534/g3.115.016808 (2015).

37 Mahmood, F. Age-related changes in development of the accessory glands of male Anopheles albimanus. J Am Mosq Control Assoc 13, 35-39 (1997).

38 Shaw, W. R. et al. Wolbachia infections in natural Anopheles populations affect egg laying and negatively correlate with Plasmodium development. Nat Commun 7, 11772, doi:10.1038/ncomms11772 (2016).

150

39 Gabrieli, P. et al. Sexual transfer of the steroid hormone 20E induces the postmating switch in Anopheles gambiae. Proc Natl Acad Sci U S A 111, 16353-16358, doi:10.1073/pnas.1410488111 (2014).

40 Bhatt, S. et al. The effect of malaria control on Plasmodium falciparum in Africa between 2000 and 2015. Nature 526, 207-211, doi:10.1038/nature15535 (2015).

41 Damiens, D., Vreysen, M. & Gilles, J. Anopheles arabiensis sperm production after genetic manipulation, dieldrin treatment, and irradiation. Journal of medical entomology 50, 314-316 (2013).

42 Andreasen, M. H. & Curtis, C. F. Optimal life stage for radiation sterilization of Anopheles males and their fitness for release. Medical and veterinary entomology 19, 238-244, doi:10.1111/j.1365- 2915.2005.00565.x (2005).

43 South, A. & Catteruccia, F. Chapter Three - Sexual Selection and the Evolution of Mating Systems in Mosquitoes. In: Advances in Insect Physiology. 51. ed.: Academic Press; 2016. p. 67-92.

44 Neafsey, D. E. et al. Mosquito genomics. Highly evolvable malaria vectors: the genomes of 16 Anopheles mosquitoes. Science 347, 1258522, doi:10.1126/science.1258522 (2015).

45 Engler, C. & Marillonnet, S. Combinatorial DNA assembly using Golden Gate cloning. Methods Mol Biol 1073, 141-156, doi:10.1007/978-1-62703-625-2_12 (2013).

46 Pondeville, E. et al. Efficient PhiC31 integrase-mediated site-specific germline transformation of Anopheles gambiae. Nat Protoc 9, 1698-1712, doi:10.1038/nprot.2014.117 (2014).

47 Fuchs, S., Nolan, T. & Crisanti, A. Mosquito transgenic technologies to reduce Plasmodium transmission. Methods Mol Biol 923, 601-622, doi:10.1007/978-1-62703-026-7_41 (2013).

48 Rogers, D. W. et al. Molecular and cellular components of the mating machinery in Anopheles gambiae females. Proceedings of the National Academy of Sciences of the United States of America 105, 19390-19395, doi:10.1073/pnas.0809723105 (2008).

151

CHAPTER 5. Discussion

Respective Contributions: This chapter was written by Andrea Smidler, with editorial input from Flaminia Catteruccia.

152

5.1 Overview

Malaria still kills nearly half a million people annually1, and as the rate of reduction of malaria cases has plateaued in recent years, novel vector control technologies targeting the anopheline vector are urgently needed to push towards eradication. The control of the mosquito vector has proven the best way to reduce transmission of this disease2. However, resistance to all classes of insecticides used in malaria control programs has emerged in Anopheles populations across Africa and Asia3. With the advent of

CRISPR/Cas9 technology, genetic-based vector control strategies capable of controlling or eliminating the

Anopheles mosquito vector have become a realistic possibility. Throughout this work, we have aimed to develop a variety of novel CRISPR-based tools for the control of wild populations of Anopheles gambiae, as well as genetic tools to aid in monitoring GM mosquito field releases.

5.2 CRISPR-mediated genetically sterile males for Sterile Insect Technique

In Chapter 2 we outline the development of a novel transgenic system for ablation of male reproductive tissues, which induces complete sterility in 95% of males. These males sterilize the females with which they mate, making them candidates for use in Sterile Insect Technique-like programs. The crossing system we develop demonstrates the power of expressing germline-constitutive Cas9 and ubiquitous gRNAs to efficiently mutate target genes and achieve gene knockouts4. Mosaic mutagenesis occurring throughout development is sufficiently strong to ablate the germ cells of both males and females. The power of the system is partially explained through its bias for the mutated outcome; every cleavage event which repairs correctly becomes a target for cleavage again, until mutagenesis occurs. The incomplete penetrance of the phenotype (in 5% of males) can be partially explained by the presence of polymorphisms inhibiting correct gRNA targeting, and the occurrence of fertility-maintaining mutations, in addition to variable gRNA cleavage efficacy. Although understanding this variability is beyond the scope of this work, this

153 finding underscores the need for further investigation of the factors influencing gRNA efficacy in anophelines, a line of research not only critical for the sterilization system outlined here, but also for the development of gene drives5.

Although highly significant, 95% sterilization is not sufficiently stringent for SIT strategies. To increase sterilization frequency, disruptive mutagenesis must be increased, and fertility-maintaining mutations minimized. To accomplish this, additional gRNAs could be incorporated to minimize the chance of fertility- maintaining mutations, perhaps by targeting or spanning splice junctions, or by targeting sites prone to disruptive MMEJ repair outcomes. To minimize gRNA-binding inhibition due to polymorphisms, elimination of SNPs during establishment of stock transgenics could be undertaken, however this becomes prohibitively complex if transgenic lines are consistently refreshed with wild individuals to maintain them as field-like. A way to circumvent this issue could be targeting multiple fertility genes simultaneously, thereby decreasing the probability of fertile escapees with each additional gene cleaved.

It is important to note that even once sterilization is optimized, sex-separation of spermless males prior to release remains an important hurdle. While fluorescence-based sex separation methods relying on testis-specific6, Y-linked7, or male splicing-specific fluorescence8,9 provide an important tool (Section 1.10), they require sorting at larval stages which, while not prohibitive, is decidedly less effective than direct female killing. CRISPR-based female-killing transgenic systems have been developed by causing shredding of the X-chromosome during spermatogenesis10, but incorporating this system into the ZPG mutants would be extremely challenging. Another option could be turning females into males by targeting sex- determining genes such as doublesex (DSX)11. However sex determination is a complex biological cascade, and despite its crucial function, it varies to a surprising degree between species. In mammals ‘maleness’ is carried on the Y chromosome12, in Drosophila it is determined by the number of X chromosomes13, and in Aedes it is independent of sex chromosomes entirely14. In An. stephensi and An. gambiae the Y-linked male determining factors Guy1 and Yob were recently discovered15,16, each playing a role in male-specific

154

DSX splicing and sex-chromosome dosage compensation. Although not fully penetrant, transgenic expression of these factors in their respective species causes significant female killing15,17, making similar constructs promising candidates for developing male-only transgenic systems.

Besides preventing sperm development, the ZPG knock-out also ablated egg development in females, a finding that has facilitated the study of the function of these tissues in the development of Plasmodium parasites18. Egg-less females helped demonstrate that vector control strategies affecting ovarian development can expedite parasite development and increase disease transmission19, providing a warning against use of some population suppression drive strategies20. The egg-less phenotype of ZPG mutant females however prevented the isolation of true-breeding knockout lines, forcing us to intercross

Cas9 and gRNA transgenics each time ZPG mutant were required. A consequence of this limitation is that released sterile males would be, by necessity, transgenic. The fact that our system uses CRISPR to induce sterility may damage the chances of public acceptance of this technology21 unlike SIT systems based on male irradiation. In contrast to gene drives, the controllability of SIT systems makes them more easily accepted in conservative settings; however shared use of CRISPR may present a point of public confusion.

This issue may be exacerbated by the fact that ZPG mutant escapees would sire CRISPR-expressing offspring capable of persisting in wild populations for a few generations, presenting a public perception hurdle not encountered by other SIT systems. Despite the limited ecological relevance of such escapees, the issue of residual fertility may thwart practical implementation of CRISPR-based SIT systems, in addition to the fact they are transgenic, stressing the importance of optimization before releases can be undertaken.

In Chapter 2 (Future Directions) we describe a next generation SIT system which could perfect the ZPG mutants for release. Termed the Inducible Eunuch system, this transgenic crossing scheme uses no

CRISPR, enables mass rearing, and causes complete male sterilization and female killing, all while reducing escape rates to nearly zero due to its reliance on RIDL22 rather than on mosaic mutagenesis. Furthermore,

155

Inducible Eunuchs would be completely non-transgenic, harboring only a small sterility-causing indel, which may provide a tremendous advantage in terms of public perception. This advantage was demonstrated recently following community outreach campaigns to gain approval for release of anti-

Lyme disease GM mice on the island of Nantucket, where the local community largely voted in favor of the exclusive use of mouse-derived DNA23. In light of a recent Florida referendum failing to approve release of GM moquitoes24, the importance of public opinion should not be discounted. Designing non- transgenic systems may be key to gaining public acceptance and accomplishing successful release of disease-combatting mosquitoes in the future, making development of systems like the Inducible Eunuchs alluring for field applications.

Following the creation of multiple An. gambiae population suppression gene drives11,20, development of genetic SIT technologies in the species may seem lackluster, however their distinct properties make them an important tool still warranting development. Although both SIT and sterilizing gene drives will select for pre-and peri-copulatory resistance mechanisms in wild populations to circumvent eradication, only drives, by nature of their genetic , will select for genetic resistance mechanisms. Much how rotation of insecticides is recommended to reduce the emergence of resistance mechanisms25, so too may rotating between SIT and localized drive suppression systems to limit the impact of genetic drive resistance prove best practice.

5.3 Developing a transgenic tool to assess Anopheles mating competitiveness in the field.

Importantly, as field releases of GM Aedes are underway26,27, and releases of population suppressing gene drive mosquitoes are on the distant horizon11, we urgently need better tools to monitor the mating competitiveness of released GM males. Our PluTo system developed in Chapter 4 achieves this goal in a simple yet effective way, allowing visualization of successful copulation long after mating via the sexual

156 transfer of a fluorescent mating plug. In contrast to strategies relying on visualization of fluorescent sperm transfer to the female6, this method can be implemented without the need of dissection and can be used for mating competitiveness of spermless male systems. While a relatively small body of work and a small contribution to science, this system may prove a valuable asset for effective future GM field trials.

5.4 Developing evolutionarily stable gene drives for Anopheles gambiae population replacement.

In Chapter 3 our work presents important findings that inform the development of evolutionarily stable gene drives for population replacement in An. gambiae, and suggests that developing homing-based drives in haplolethal ribosome genes may be biologically impossible without first overcoming significant hurdles. To build gene drives that maintain the strongest possible selection against DRAs, we designed them to home into – but not disrupt – a haplolethal and haplosufficient gene pair, using the selection pressure on these genes to guarantee homing and prevent DRAs. However, we found that manipulating these haplolethal genes leaves very small room for error, and consistently obstructed drive construction.

To develop the system, we first inserted a docking site into the RpL11-Rpt1 target, followed by subsequent drive insertion using standard transgenesis techniques (Section 3.3). Both of these steps were fraught with difficulties preventing successful drive development. During attempts to establish gene drive docking lines by HDR, we observed considerable embryonic lethality preventing line establishment (Section 3.4), a hurdle that motivated development of CrIGCkid to facilitate knockin from transgenic donors by interlocus gene conversion. While CrIGCkid was initially successful at generating two distinct docking lines, it failed in later experiments, providing the first evidence of possible gRNA silencing, and preventing in depth characterization of this phenomenon (Section 3.5). Following their establishment, both docking lines displayed unexpected Minute phenotypes (Figure 3.8) most likely caused by suboptimal recoding or

157 misregulation by a heterologous 3’UTR, urging further research into anopheline codon usage and UTR regulation to overcome infertility phenotypes, and enable future drive. Inserting gene drives into the docking sites proved further problematic. While multiple drive constructs could be inserted into the docking lines, those capable of Cas9 expression died prematurely (Section 3.7), suggesting either these sites have minimal tolerance for Cas9 expression, or inserted Cas9-expressing transgenes are aberrantly regulated compared to other loci, and in the presence of ubiquitous gRNAs causes lethal mutagenesis in intolerant tissues. This is not unexpected given the role position effect plays in influencing transgenic variability in mosquitoes28, and the docking site’s intimate association with the highly transcriptionally- active RpL11 and Rpt1 genes. While achieving tighter CRISPR regulation within these sites may be feasible through use of Cas9 nickases29,30, gypsy insulators31, or tissue-specific gRNAs32, these experiments would have considerably expanded the scope and timeline of the current work, making further optimization impossible at this time.

In an attempt to characterize putative drive properties of the system in the absence of a functional self- autonomous drive, we tested the ability of the ‘empty’ docking site to be homed by trans-expressed

CRISPR from four distal drives, and attempted rescue of the nondriving drive (mNosGDdRPLT) with ectopic

VasaCas9. While many distal drives failed to express Cas9 and failed to drive, in experiments with Cas9 expressed by known drive-enabling promoters11,20,33,34, all females either died following blood feeding

(Figure 3.13 ~G20), or were infertile (Figure 3.13 ~G40, Figure 3.16). These phenotypes are consistent with oogenic defects associated with ribosome depletion, and were lessened over time, suggesting drive designs targeting ribosome genes may struggle to maintain female fertility, and providing further evidence of gRNA silencing. Similar experiments in driving males revealed nearly complete sterilization in early broods, ‘anti-drive’ transgene ratios in intermediate broods, and low fertility with normal transgenic ratios in late broods. Taken together with the findings that all surviving larvae were the product of sperm that had escaped mutagenesis, this suggests that targeting the ribosome by CRISPR in the primordial germline

158 may be fundamentally incompatible with fertility, and that transgenic expression of CRISPR may dampen over time as a ‘driving’ male ages, consistent with transgene silencing by PIWI35-38. Moreover, he absence of homing in these males indicates that multi-gRNA drive designs may promote End-joining repair over

HDR as previously observed39, or mutagenesis may precede or supersede HDR in the PGCs, implying that targeting the ribosome for drive may be fundamentally incompatible with maintaining male fertility, underscoring the critical need for Cas9 fine-tuning in these designs. Unless all of these hurdles can be overcome, these findings suggest that our gene drive design – and perhaps all ribosome-targeting replacement drive systems – may be biologically inviable, suggesting the development of homing-based evolutionarily stable gene drives for population replacement may take years of further improvements.

One troubling hurdle facing the future of gene drive design is the lessening over time of phenotypes reliant on robust gRNA expression, suggesting inactivation, we believe through mechanisms that may have evolved to silence transposon selfish-genetic elements35,36,40-44. CRISPR gene drives may prove to be equally burdensome genetic parasites as transposons and retroviruses45-47, and may succumb to deactivation accordingly. In fact while near extinction of suppression drives was thought to be caused by

DRAs48, failure to home due to gRNA silencing cannot be ruled out. If this is the case, even once DRA- inhibition is achieved, overcoming gRNA silencing may be the next big impediment facing drive scientists.

However, it is noteworthy that we have never observed complete loss of gRNA-dependent phenotypes.

The ZPG crosses still make germ cell mutants 42 months later, and mNosGDdRPLT females supplemented with VasCas9 are still wholly infertile after 40 generations (when they died 24h post-blood feeding at G20), suggesting gRNA activity is not lost, only dampened – at least over the time scales observed within this work. While significant silencing could happen in a generation36, complete silencing may never occur, allowing homing to continue for years – after all transposons still jump despite eons of counter- evolution49,50. More research is required on the subject, not just for mosquito gene drives, but also for any

159 eukaryotes expressing transgenic gRNAs51, and may prove a significant hiccup for numerous CRISPR-based transgenic applications.

If building multi-gRNA replacement drives in ribosome genes proves impossible, then the scientific community may have to relax its expectations for the ideal drive. If a suitable non-ribosomal haplolethal gene target can be identified, and multi-gRNA designs can be constructed capable of stimulating robust

HDR, then such replacement drives may still be possible. The alternative is construction of a drive within haplosufficient targets, but inhibition by heterozygous r2DRAs will greatly slow drive spread, likely preventing population replacement. Integral drive designs have been proposed for evolutionarily stable population replacement, but their reliance on single gRNAs makes the evolution of r1DRAs almost certainly inevitable52. On the other hand, population suppression gene drives impose such immense pressure on the mosquito host to evolve mechanisms of resistance that they too will likely encounter setbacks, even when targeting the most conserved sites11. In one form or another, all drives face a battle against evolution, and every drive design conceived by (wo)man will eventually be deactivated.

There is however still hope for development of our original drive designs – using evolution as our guide.

Similarly to naturally occurring selfish genetic elements which have persisted for billions of years50, synthetic self-autonomous drive systems should also be self-optimizing once they overcome the hurdle of DRAs (Section 1.8B). So even though current designs do not drive, we may be able to harness the power of evolution to guide us towards building one. By beginning with a line with little to no baseline drive, and nearly undetectable Cas9, but with all other necessary drive characteristics, maintaining these proto- drives in the lab for many generations could select for mutants with perfected drive properties. Depending on how close the proto-drive is to functioning, and the feracity* of the species in which it’s tested, such experiments may be possible within practical timescales, and in fact have already begun. By maintaining

* The state of being feracious. Producing in abundance; fertile, fruitful. (Archaic) Last defined in Webster’s Revised Unabridged Dictionary, G. & C. Merriam, 1913

160 the nearly self-autonomous mNosGDdRPLT for over 40 generations, we have inadvertently begun to select for evolution of de novo drive. While no explicit biased inheritance of the transgene has been observed

(Section 3.7B), artifacts visualized by Western hint that subtle or modified Cas9 expression may be occurring in some individuals at ~G40 (Figure 3.10) when it was absent at ~G20. And while early generations of this transgenic line were plagued by constant near-extinction events due to insertion within dRPLT, recent generations have not experienced this, no longer requiring heterozygote enrichment for maintenance. While recent generations of this line have not yet been re-sequenced, maintenance of gRNA function was confirmed recently by experimental phenotypes (Figure 3.13), suggesting that the transgene has either lost the properties which made it once unfit, or gained a de novo ability to prevent extinction (drive?) by yet unknown mechanisms. This line merits further investigation as it may reveal the secrets necessary for development of evolutionarily stable population replacement gene drives.

5.5 CONCLUDING REMARKS

From gene drives to sterile male systems, and from knockouts to knockins, and developing monitoring tools in between, this dissertation provides an extensive body of work, considerably expanding the genetic toolkit in Anopheles gambiae mosquitoes. Our work demonstrates novel principles for genetic sterilization, and presents significant advancements towards the development of evolutionarily stable gene drives for population replacement, revealing insights critical for all CRISPR-based gene drives in the process, and it is our hope that it may one day contribute to the eradication of one of humanity’s most devastating diseases.

161

5.6 REFERENCES

1 WHO. World malaria report 2017. Geneva, Switzerland: World Health Organization

2 Dyck, V. A., Hendrichs, J. & Robinson, A. S. Sterile insect technique : principles and practice in area- wide integrated pest management. (Springer, 2005).

3 Ranson, H. Current and Future Prospects for Preventing Malaria Transmission via the Use of Insecticides. Cold Spring Harbor perspectives in medicine 7, doi:10.1101/cshperspect.a026823 (2017).

4 Smidler, A. L., Terenzi, O., Soichot, J., Levashina, E. A. & Marois, E. Targeted mutagenesis in the malaria mosquito using TALE nucleases. PloS one 8, e74511, doi:10.1371/journal.pone.0074511 (2013).

5 Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife, e03401, doi:10.7554/eLife.03401 (2014).

6 Catteruccia, F., Benton, J. P. & Crisanti, A. An Anopheles transgenic sexing strain for vector control. Nature biotechnology 23, 1414-1417, doi:10.1038/nbt1152 (2005).

7 Bernardini, F. et al. Site-specific genetic engineering of the Anopheles gambiae Y chromosome. Proceedings of the National Academy of Sciences of the United States of America 111, 7600-7605, doi:10.1073/pnas.1404996111 (2014).

8 Marois, E. et al. High-throughput sorting of mosquito larvae for laboratory studies and for future vector control interventions. Malaria journal 11, 302, doi:10.1186/1475-2875-11-302 (2012).

9 Magnusson, K. et al. Transcription regulation of sex-biased genes during ontogeny in the malaria vector Anopheles gambiae. PloS one 6, e21572, doi:10.1371/journal.pone.0021572 (2011).

10 Galizi, R. et al. A CRISPR-Cas9 sex-ratio distortion system for genetic control. Scientific reports 6, 31139 (2016).

11 Kyrou, K. et al. A CRISPR–Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature biotechnology 36, 1062, doi:10.1038/nbt.4245 (2018).

12 Wilhelm, D., Palmer, S. & Koopman, P. Sex determination and gonadal development in mammals. Physiological reviews 87, 1-28 (2007).

13 Erickson, J. W. & Quintero, J. J. Indirect effects of ploidy suggest X chromosome dose, not the X: A ratio, signals sex in Drosophila. PLoS biology 5, e332 (2007).

14 Hall, A. B. et al. A male-determining factor in the mosquito Aedes aegypti. Science 348, 1268-1270 (2015).

162

15 Criscione, F., Qi, Y. & Tu, Z. GUY1 confers complete female lethality and is a strong candidate for a male-determining factor in Anopheles stephensi. eLife 5, e19281 (2016).

16 Krzywinska, E., Dennison, N. J., Lycett, G. J. & Krzywinski, J. A maleness gene in the malaria mosquito Anopheles gambiae. Science 353, 67-69 (2016).

17 Krzywinska, E. & Krzywinski, J. Effects of stable ectopic expression of the primary sex determination gene Yob in the mosquito Anopheles gambiae. Parasites & vectors 11, 648 (2018).

18 Werling, K. et al. Steroid Hormone Function Controls Non-competitive Plasmodium Development in Anopheles. Cell, doi:10.1016/j.cell.2019.02.036.

19 Macdonald, G. Epidemiological basis of malaria control. Bulletin of the World Health Organization 15, 613 (1956).

20 Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nature biotechnology 34, 78-83, doi:10.1038/nbt.3439 (2015).

21 Bloss, C. S., Stoler, J., Brouwer, K. C., Bietz, M. & Cheung, C. Public Response to a Proposed Field Trial of Genetically Engineered Mosquitoes in the United States. JAMA 318, 662-664, doi:10.1001/jama.2017.9285 (2017).

22 Thomas, D. D., Donnelly, C. A., Wood, R. J. & Alphey, L. S. Insect population control using a dominant, repressible, lethal genetic system. Science 287, 2474-2476, doi:DOI 10.1126/science.287.5462.2474 (2000).

23 Najjar, D. A., Normandin, A. M., Strait, E. A. & Esvelt, K. M. Driving towards ecotechnologies. Pathogens and global health 111, 448-458, doi:10.1080/20477724.2018.1452844 (2017).

24 Servick, K. Update: Florida Voters Split on Releasing GM . Science 10 (2016).

25 Mnzava, A. P. et al. Implementation of the global plan for insecticide resistance management in malaria vectors: progress, challenges and the way forward. Malaria journal 14, 173 (2015).

26 Carvalho, D. O. et al. Suppression of a Field Population of Aedes aegypti in Brazil by Sustained Release of Transgenic Male Mosquitoes. PLoS neglected tropical diseases 9, e0003864, doi:10.1371/journal.pntd.0003864 (2015).

27 Malavasi, A. Project Aedes transgenic population control in Juazeiro and Jacobina Bahia, Brazil. BMC Proceedings 8, O11-O11, doi:10.1186/1753-6561-8-S4-O11 (2014).

28 Benedict, M. Q., McNitt, L. M., Cornel, A. J. & Collins, F. H. Mosaic: a position-effect variegation eye-color mutant in the mosquito Anopheles gambiae. The Journal of heredity 91, 128-133 (2000).

29 Ran, F. A. et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154, 1380-1389, doi:10.1016/j.cell.2013.08.021 (2013).

163

30 Mali, P. et al. CAS9 transcriptional activators for target specificity screening and paired nickases for cooperative genome engineering. Nature biotechnology 31, 833-838, doi:10.1038/nbt.2675 (2013).

31 Carballar-Lejarazú, R., Jasinskiene, N. & James, A. A. Exogenous gypsy insulator sequences modulate transgene expression in the malaria vector mosquito, Anopheles stephensi. Proceedings of the National Academy of Sciences 110, 7176-7181 (2013).

32 Xu, L., Zhao, L., Gao, Y., Xu, J. & Han, R. Empower multiplex cell and tissue-specific CRISPR- mediated gene manipulation with self-cleaving ribozymes and tRNA. Nucleic acids research 45, e28-e28, doi:10.1093/nar/gkw1048 (2017).

33 Hammond, A. M. et al. Improved CRISPR-based suppression gene drives mitigate resistance and impose a large reproductive load on laboratory-contained mosquito populations. bioRxiv, doi:10.1101/360339 (2018).

34 Gantz, V. M. & Jasinskiene, N. Highly efficient Cas9-mediated gene drive for population modification of the malaria vector mosquito Anopheles stephensi. Proceedings of the …, doi:10.1073/pnas.1521077112 (2015).

35 Kalmykova, A. I., Klenov, M. S. & Gvozdev, V. A. Argonaute protein PIWI controls mobilization of retrotransposons in the Drosophila male germline. Nucleic acids research 33, 2052-2059, doi:10.1093/nar/gki323 (2005).

36 Khurana, J. S. et al. Adaptation to P element transposon invasion in Drosophila melanogaster. Cell 147, 1551-1563, doi:10.1016/j.cell.2011.11.042 (2011).

37 Ozata, D. M., Gainetdinov, I., Zoch, A., O'Carroll, D. & Zamore, P. D. PIWI-interacting RNAs: small RNAs with big functions. Nature reviews. Genetics 20, 89-108, doi:10.1038/s41576-018-0073-3 (2019).

38 Vagin, V. V. et al. A distinct small RNA pathway silences selfish genetic elements in the germline. Science 313, 320-324, doi:10.1126/science.1129333 (2006).

39 Oberhofer, G., Ivy, T. & Hay, B. A. Behavior of homing endonuclease gene drives targeting genes required for viability or female fertility with multiplexed guide RNAs. Proceedings of the National Academy of Sciences 115, E9343-E9352, doi:10.1073/pnas.1805278115 (2018).

40 Al-Kaff, N. S. et al. Transcriptional and posttranscriptional plant gene silencing in response to a pathogen. Science 279, 2113-2115 (1998).

41 Baulcombe, D. RNA silencing in plants. Nature 431, 356 (2004).

42 Tabara, H. et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123-132 (1999).

43 Yoder, J. A., Walsh, C. P. & Bestor, T. H. Cytosine methylation and the ecology of intragenomic parasites. Trends in genetics 13, 335-340 (1997).

164

44 Bestor, T. H. in Found. Symp. 187-195.

45 Iranzo, J., Puigbò, P., Lobkovsky, A. E., Wolf, Y. I. & Koonin, E. V. Inevitability of Genetic Parasites. Genome biology and evolution 8, 2856-2869, doi:10.1093/gbe/evw193 (2016).

46 Zeh, D. W., Zeh, J. A. & Ishida, Y. Transposable elements and an epigenetic basis for punctuated equilibria. BioEssays : news and reviews in molecular, cellular and developmental biology 31, 715- 726 (2009).

47 Kidwell, M. G. & Lisch, D. R. Hybrid genetics. Transposons unbound. Nature 393, 22-23, doi:10.1038/29889 (1998).

48 Hammond, A. M. et al. The creation and selection of mutations resistant to a gene drive over multiple generations in the malaria mosquito. PLoS genetics 13, e1007039, doi:10.1371/journal.pgen.1007039 (2017).

49 Hickman, A. B., Chandler, M. & Dyda, F. Integrating prokaryotes and eukaryotes: DNA transposases in light of structure. Critical reviews in biochemistry and molecular biology 45, 50-69 (2010).

50 Fedoroff, N. V. Transposable elements, epigenetics, and genome evolution. Science 338, 758-767 (2012).

51 Grunwald, H. A. et al. Super-Mendelian inheritance mediated by CRISPR–Cas9 in the female mouse germline. Nature 566, 105 (2019).

52 Nash, A. et al. Integral gene drives for population replacement. Biol Open 8, doi:10.1242/bio.037762 (2019).

165

APPENDIX A

Plasmids and transgene Maps

166

167

APPENDIX B

Targeted Mutagenesis in the Malaria Mosquito Using TALE Nucleases

168

Targeted Mutagenesis in the Malaria Mosquito Using TALE Nucleases

Andrea L. Smidler1,2¤a, Olivier Terenzi1,2, Julien Soichot1,2, Elena A. Levashina1,2¤b, Eric Marois1,2*

1 Institut National de la Santé et de la Recherche Médicale U963, Strasbourg, France, 2 Centre National de la Recherche Scientifique UPR9022, Strasbourg, France

Abstract

Anopheles gambiae, the main mosquito vector of human malaria, is a challenging organism to manipulate genetically. As a consequence, reverse genetics studies in this disease vector have been largely limited to RNA interference experiments. Here, we report the targeted disruption of the immunity gene TEP1 using transgenic expression of Transcription-Activator Like Effector Nucleases (TALENs), and the isolation of several TEP1 mutant A. gambiae lines. These mutations inhibited protein production and rendered TEP1 mutants hypersusceptible to . The TALEN technology opens up new avenues for genetic analysis in this disease vector and may offer novel biotechnology-based approaches for malaria control.

Citation: Smidler AL, Terenzi O, Soichot J, Levashina EA, Marois E (2013) Targeted Mutagenesis in the Malaria Mosquito Using TALE Nucleases. PLoS ONE 8(8): e74511. doi:10.1371/journal.pone.0074511 Editor: George Dimopoulos, Johns Hopkins University, Bloomberg School of Public Health, United States of America Received June 24, 2013; Accepted August 7, 2013; Published August 15, 2013 Copyright: © 2013 Smidler et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Funding: This study was funded by an Agence Nationale de la Recherche (ANR) young researcher grant to E.M. (project GEMM), by the European Commission FP7 projects INFRAVEC (grant agreement no. 228421) and EVIMalar (grant agreement no. 242095), by the Institut National de la Santé et de la Recherche Médicale (INSERM) and by the Centre National de la Recherche Scientifique (CNRS). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. Competing interests: The authors have declared that no competing interests exist. * E-mail: [email protected] ¤a Current address: Department of Immunology and Infectious Diseases, Harvard School of Public Health, Boston, Massachusetts, United States of America ¤b Current address: Department of Vector Biology, Max-Planck Institute for Infection Biology, Berlin, Germany

Introduction vulnerable points in the parasite cycle. On the other hand, the prospect of releasing engineered male mosquitoes to Malaria is caused by Plasmodium parasites transmitted to propagate malaria resistance genes through wild susceptible their human hosts by the bite of anopheline mosquitoes. vector populations has been receiving increasing attention Malaria has been charged with causing more human deaths [3,4]. Novel methods to disrupt or alter target genes of interest than any other disease in human history and continues to kill in the malaria mosquito would promote rapid progress towards about 660,000 annually [1]. A reduction in the malaria death toll these goals. Furthermore, targeted genetic modifications that has been achieved thanks to vector control using insecticides do not require the permanent introduction of transposons in the and insecticide-impregnated bednets, better health care and genome are particularly desirable, as they eliminate the progress in medical treatment, but this success is currently potential risk of subsequent unplanned transposon mobilization mitigated by the spread of resistance both of mosquitoes to by natural sources of transposition factors. insecticides and of Plasmodium to antimalarial drugs. Recently, a novel class of DNA-binding protein domain Sequencing the genome of the main malaria vector, Anopheles derived from the Xanthomonas Transcription Activator Like gambiae [2], enabled the identification of hundreds of genes Effector (TALE) proteins [5,6] has been successfully harnessed involved in the vector’s capacity to transmit Plasmodium. to custom-design sequence-specific endonucleases [7,8]. Altering the mosquito genome in a way that abates These TALE nucleases (TALENs), in which the TAL DNA- Plasmodium transmission, through transgenesis or other binding domain is fused to the FokI endonuclease domain, are sophisticated genetic engineering tools, can offer new easy to engineer (e.g., [9,10]), highly predictable in their perspectives in the fight against malaria. On one hand, sequence specificity, and highly mutagenic [11] making them experimentally altering mosquito genes of interest will advance an attractive alternative to Zinc Finger Nucleases that have our fundamental understanding of the biological interactions less predictable binding specificities and require in vitro between mosquito and parasite, and may help target optimization [12]. Mutations arise by imprecise repair of the

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 169 TALEN-induced TEP1 mutations in Anopheles gambiae

TALEN-generated double-stranded breaks by the non- and subjected the amplification product to a restriction digest homologous end joining (NHEJ) repair pathway. In a number of with NcoI (Figure 1B). Of 310 screened F2 larvae, 16 (5.16%) animal species, injection of mRNA encoding TALENs have carried a heterozygous mutation at the target locus, as readily allowed researchers to generate mutants in their genes evidenced by the appearance of a PCR product that NcoI was of interest [13–19]. Very recently, mutagenesis of an eye unable to cleave. Thus, at least 2.58% of TEP1 copies were pigmentation gene was achieved in Aedes mosquitoes using mutated after exposure to one TALEN dose (i.e., one this method [20]. In rice, disease resistant mutants have been generation). This figure is a conservative estimate of mutation produced by transgenic expression of the TALENs, whose frequency, as we subsequently observed some mutations that respective transgenes were eliminated by subsequent genetic left the NcoI recognition sequence intact. In order to obtain a crosses once the desired mutation had been fixed [21]. We mosquito population containing a higher frequency of hypothesized that a similar approach would be applicable to mutations, we self-crossed successive generations of insect vectors and set out to use TALENs to target the TEP1 mosquitoes expressing both TALENs of the pair. This was gene, a key component of the mosquito immune system. facilitated by automated COPAS selection of larvae [28] that had inherited one copy each of the left and right TALEN genes, Results which are respectively associated with a red and yellow fluorescent marker expressed in the [29]. At Since the embryo microinjection procedure is technically least 1000 double-TALEN larvae were COPAS-selected and challenging in Anopheles gambiae (as judged by poor survival cultured for each generation. In the 7th generation (i.e., and relatively low success rate of transgenesis in this species), exposure to 6 TALEN mutagenic doses), 51% (49 out of 96) of we expected that mutant recovery after direct injection of the examined individual mosquitoes carried a heterozygous TALEN-encoding plasmids or mRNA might be difficult. For this mutation in TEP1. This indicated that the frequency of reason and to enable controlled mutagenesis experiments, we mutations increased faster than predicted if mutations preferred transgenic expression of the TALENs in the mosquito accumulated linearly from one generation to the next. This germ cells. To obtain the proof-of-principle for gene targeting in observation is consistent with a model where mosquitoes A. gambiae via transgenic TALENs, we selected the well- already carrying a heterozygous mutation employ homologous characterized immune gene TEP1 as a target. TEP1, a protein recombination to repair new TALEN-induced breaks in the wild- similar to vertebrate complement factor C3, binds Plasmodium type chromosome, thereby effectively copying the existing parasites as they invade the mosquito intestine and kills them mutation onto the newly damaged chromosome. This suggests in a manner probably dependent on its thioester site located in that in addition to NHEJ, TALEN-caused breaks can also be the C-terminus [22–25]. TEP1 mutants will be instrumental in repaired by homologous recombination. Therefore, the further dissecting the antiparasitic complement-like system in observed number of NHEJ mutations is an underestimation of mosquitoes. the true rate of TALEN activity. TALENs function in pairs, each member of which binds a We wondered if a single TALEN of the pair is capable of chosen 12 to 24-nucleotide sequence. The two selected target causing mutations in TEP1. To investigate this, we sampled sequences are separated by 14-16 nucleotides, the optimal mosquito larvae from lines carrying a single TALEN maintained distance for the two FokI domains of the TALENs to properly at high population levels for 8 (right TALEN) or 10 (left TALEN) dimerize and create a double-stranded break. Cleavage of the generations, and again sampled larvae after about 16 bound DNA molecule occurs near the center of the sequence generations. We expected that such a high number of separating the two target sites. We designed a single pair of generations would have allowed rare “monotalenic” mutational TALENs to target a site within the TEP1 gene centered on an events to accumulate in the population. Out of 96 individual NcoI restriction site (5’-CCATGG-3’), offering the possibility to larvae tested by PCR for each sample, none carried a mutation easily screen individual mosquitoes for mutations that in the TEP1 NcoI site. The observed absence of mutations destroyed the NcoI site (Figure 1A). Mutations in this region are among 576 haploid genomes exposed to single TALEN activity expected to strongly affect TEP1 protein function: frame-shifts for 7, 9 or 15 generations suggests that single TALENs never resulting in premature stop codons would remove the C- or rarely induce mutations at their target site. To strengthen this terminal third of the protein including its thioester domain, while point, we purified DNA from 2600 pooled larvae whose amino-acid deletions or insertions would alter the length of the genomes had been exposed to one TALEN of the pair for 7 or alpha-helix connecting the CUB domain to the thioester domain 9 generations, and PCR-amplified the target region. PCR [26], presumably resulting in destabilization of the protein’s products appeared to be fully cleaved by NcoI, pointing to the structure. absence of TALEN-induced mutations. To increase the chance A distinct transgenic mosquito line was generated for each of detection of a minor fraction of mutated products, we purified TALEN of the pair, the expression of which was driven by the the region of the gel in which uncleaved PCR products may A. gambiae Vasa promoter, active in the mosquito germline exist, cloned them into a plasmid, and examined E. coli [27]. Therefore, F1 mosquitoes arising from a cross between transformants containing single copies of the PCR products. the two lines will express both TALENs simultaneously and are Again, all cloned fragments were cleaved by NcoI. Although we expected to produce F2 gametes carrying mutations in the note that deep sequencing of amplicons would provide a more TEP1 gene. To screen individual larvae within the F2 progeny, sensitive assay to detect rare mutations, this result further we PCR-amplified a TEP1 fragment spanning the target site suggests that single TALENs rarely or never generate mutants.

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 170 TALEN-induced TEP1 mutations in Anopheles gambiae

Figure 1. TALEN mutagenesis of the TEP1 gene. A: Fragment from the TEP1 gene showing the target site of the TALEN pair. Nucleotides bound by each TALEN are underlined, TALEN repeats are color-coded to show repeat/nucleotide specificity. The NcoI restriction site centrally located at the TALEN cleavage site is highlighted. Inset: scheme of the entire TEP1 protein showing the location of TALEN-induced mutations (SP: signal peptide; CUB: CUB domain interrupted by the TED: thioester domain; the star indicates the position of the thioester site). B: PCR assay to identify TEP1 mutant mosquitoes. A PCR product spanning the TALEN target site is generated from individual mosquitoes (small larva or a leg from a living adult) and incubated with NcoI. Full cleavage of the PCR product (w) denotes a wild-type individual. Partial cleavage (h) denotes a heterozygous TEP1 mutant. Absence of cleavage (H) corresponds to a homozygous TEP1 mutant. C: TALEN-induced mutations in the TEP1 gene. Left and right TALEN target nucleotide sequences are shown in green and blue respectively, with the 15bp spacer sequence between the TALENs in black. The NcoI restriction site is highlighted in orange. Deletions are designated by a red dash or by Δ+number of missing bases. Insertions are shown in lowercase red letters. Uppercase red letters correspond to natural polymorphisms between multiple TEP1 alleles. doi: 10.1371/journal.pone.0074511.g001

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 171 TALEN-induced TEP1 mutations in Anopheles gambiae

The apparent absence of single TALEN background activity To examine the effect of TEP1 dosage in the process of was likely facilitated by our design scheme in which we used parasite killing, we compared oocyst infection levels between obligate heterodimeric FokI domains in TALEN construction TEP1 heterozygous mutant, homozygous mutant and control [30,31]. mosquitoes (Figure 3d). The heterozygous mutant had an The TALEN-generated TEP1 mutations were very similar in intermediate susceptibility phenotype, which was closer to, but nature to mutations obtained in other organisms (Figure 1C) significantly different from, the control. This suggests that the and consisted mainly of small deletions and insertions (indels) efficiency of parasite killing depends on TEP1 protein levels. that are the hallmark of imprecise NHEJ-mediated DNA repair. The antiparasitic role of TEP1 was discovered and Some mutations deleted a number of nucleotides in multiples characterized using RNA interference assays, in which of three, resulting in the deletion of one to a few amino acids synthetically produced double-stranded RNA homologous to a from the TEP1 protein. Other mutations introduced a frame- fragment of native TEP1 is injected at a high concentration in shift in the TEP1 coding sequence, resulting in the loss of the the body of adult mosquitoes. This technique was generalized entire C-terminal half of the protein, which contains features for the functional characterization of hundreds of mosquito crucial for TEP1 function including the thioester domain. Using genes [34]. However, injection per se was reported to impinge a PCR selection procedure from single legs taken from live on Plasmodium development by the potential induction of the mosquitoes, we recovered five homozygous mutant mosquito wounding response [35]. Therefore the mutant TEP1 lines lines. These mutant lines were representative of the main developed here will be useful in studies that must exclude the classes of mutations: line M∆T and M∆MV are deletions of 1 confounding effect of the injection procedure itself. To examine (Threonine) and 2 (Methionine and Valine) amino-acids, how the classical TEP1 RNAi phenotype compares with the respectively; line M∆3+6 lacks 3 endogenous amino-acids but mutant phenotype, we compared parasite loads in M∆T to gained an insertion of 6 exogenous ones, line M∆ct1 and M∆ct2 dsTEP1-injected mosquitoes of the parental line (Figure 3e). are frame-shift mutations causing the loss of the entire protein The mutant and RNAi phenotypes were very similar, validating C-terminus (Figure 2A). Immunoblotting analysis failed to a posteriori the high efficiency of RNAi knockdown. detect TEP1 in the Mut∆ct1 and M∆ct2 lines, while it revealed strongly reduced protein levels in M∆T, M∆MV and M∆3+6 lines Discussion (Figure 2B, top). Although small amounts of TEP1 protein could be detected in whole mosquito extracts of the M∆T, M∆MV and Here, we obtained proof of principle that TALENs can be M∆3+6 lines, these proteins did not undergo cleavage and were used for targeted mutagenesis in the genome of the malaria impaired in their secretion to the hemolymph, as no TEP1 mosquito, which is notably difficult to manipulate genetically. signal was observed in immunoblotting of hemolymph samples Recently, Aryan et al. [20] reported disruption of the eye (Figure 2B, bottom). Therefore, the alpha helix connecting the pigmentation kmo gene in the dengue vector mosquito Aedes CUB and thioester domains of TEP1 [26,32] seems to be very aegypti by injection of TALEN-encoding plasmid DNA. In the sensitive to insertion or deletion of single amino acids that same species, successful mutagenesis of GFP and of the destabilize the structure and prevent proper protein synthesis odorant receptor co-receptor (orco) was achieved by injecting and secretion. mRNA encoding Zinc Finger Nucleases [36]. In both cases and We next assessed the phenotype of these mutant lines in in other animal systems for which TALEN mutagenesis has comparison to the parental lines by infection assays using been reported, mutants were obtained directly upon injection of Plasmodium berghei-infected mice (Figure 3). Similar to the DNA or RNA into embryos. In contrast, we employed RNA interference knockdown phenotype of TEP1 [22], all transgenically-expressed TALENs for this purpose. Besides the mutations in the homozygous state resulted in a dramatic assurance of expressing the pair of TALENs in all germ cells, increase in the number of developing parasites within the this offered the possibility to increase the frequency of mutant mosquito gut. These results confirm the pivotal role of TEP1 in alleles in successive TALEN-expressing generations of antiparasitic responses. While it cannot be fully excluded that mosquitoes. We disrupted the antiparasitic gene TEP1 as a RNAi-mediated TEP1 knockdown may also affect genes first target; future work will make use of the obtained sharing some nucleotide identity with the TEP1 sequence, such hypersusceptible mutant lines to further dissect the role of this as other closely-related genes in the TEP family [33], mutations important anti-malarial factor in parasite killing. Beyond the in TEP1 should leave the expression of other genes research field of mosquito immunity, this study paves the way unaffected. for numerous other applications such as obtaining mutations in Interestingly, while midguts from TEP1 mutant mosquitoes Anopheles genes considered to be essential for Plasmodium often displayed impressive oocyst numbers that could exceed parasite development [37–39]. Disrupting these genes, or 1500, each experiment also yielded mutant midguts bearing no altering specific domains on the encoded proteins, may render or only a few oocysts. Upon blood feeding on a single infected homozygous mutant mosquitoes unable to support parasite mouse, only well-gorged mosquito females were selected for development. Such parasite-refractory mutant mosquitoes subsequent dissection. It is therefore unlikely that some could be used in anti-malaria intervention schemes, including females ingested dramatically reduced numbers of P. berghei vector population replacement in endemic regions. Unless a . Thus, it is apparent that TEP1-independent gene-drive strategy is specifically designed to spread desired mechanisms are at work to limit Plasmodium infection in a mutations, this could be achieved by the repeated release of subset of mosquitoes. mass-produced male mosquitoes. Where knockout mutants in

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 172 TALEN-induced TEP1 mutations in Anopheles gambiae

Figure 2. TEP1 mutant proteins. A: Fragment of the TEP1 protein encompassing the TALEN-induced mutations is shown for the wild-type (WT) and for those mutants that we maintain as homozygous mosquito lines (M∆T, M∆MV, M∆3+6, M∆ct1 and M∆ct2). Gaps in the protein sequence denote amino acid deletions. Inserted exogenous amino acids are shown in red. Nonsense amino acids followed by a stop codon (*), resulting from frame-shift mutations, are shown in blue. B: Immunoblots to evaluate the presence of mutant TEP1 protein in whole mosquito extract (top panels) or hemolymph. Hemolymph prophenoloxidase (PPO) serves as a loading control. In the control samples, both TEP1 full-length (full) and C-terminal fragment (cleaved) are visible. Cross-reacting background bands, some running in close proximity to TEP1 fragments, are marked on the left with ‘>’ signs. doi: 10.1371/journal.pone.0074511.g002 a given target gene might compromise the fitness of the genes for use in the Sterile Insect Technique to reduce vector mosquitoes, TALEN mutagenesis offers the possibility of a populations [40]. Of note, the TALEN transgenes used to gene therapy to cure mosquitoes of malaria by selecting obtain a desired mutation can subsequently be discarded by mutations that prevent a protein’s interactions with parasite selection, thereby rendering the obtained homozygous mutant factors while preserving its other vital functions. TALEN mosquitoes transgene-free. This could facilitate mosquito mutagenesis could also be employed to knock out male fertility

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 173 TALEN-induced TEP1 mutations in Anopheles gambiae

Figure 3. TEP1 mutant mosquitoes are hypersusceptible to P. berghei. Mosquito females from five different homozygous mutant mosquito lines and from control parental lines were offered a blood meal on a P. berghei-infected mouse. Seven days after infection, the midgut was dissected and the number of oocysts developing in each midgut was evaluated. The statistical significance of differences in mean parasite numbers was measured with a Mann-Whitney test (mutant versus control) and with a Kruskall-Wallis test followed by Dunn’s post-test (to compare all groups in [d]). (a) M∆T mosquitoes are compared to the two parental, non-mutant TALEN lines. (b) 4 different mutant lines are compared to the parental mosquito line that initially served to produce TALEN transgenic lines. This control line was verified to show the same level of susceptibility to P. berghei as the two TALEN daughter lines (not shown). In this experiment, three mutant lines showed significantly elevated parasite numbers compared to the control while the M∆ct2 line did not, presumably due to a different physiological condition of this mosquito culture. In two independent experiments (c, d), we used mosquitoes of different genotypes marked with distinct fluorescence markers and cultured the larvae together in the same water to eliminate potential confounding factors due to rearing conditions. On the day of dissection, genotypes were separated on the basis of fluorescence. The same M∆ct2 line shows significantly elevated parasite numbers. (d) Heterozygous M∆ct2 mosquitoes are compared to control and homozygous M∆ct2 mosquitoes: the susceptibility phenotype of the heterozygote is intermediate. (e) Control mosquitoes of the parental line, mosquitoes of the parental line injected with TEP1 double-stranded RNA, and homozygous TEP1 mutant mosquitoes of the M∆T line are compared. TEP1 mutant and dsRNA-injected mosquitoes show comparable susceptibility to P. berghei. doi: 10.1371/journal.pone.0074511.g003

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 174 TALEN-induced TEP1 mutations in Anopheles gambiae

release interventions that adhere to local regulations regarding directly in the case of homozygous individuals, or cloned genetically modified organisms. (CloneJET PCR Cloning Kit, Fermentas) and subsequently sequenced. Materials and Methods Mutagenesis and mutant line recovery Ethics statement For mutagenesis experiments, the two TALEN lines were Our experimental protocols were approved by Comité de crossed. From the F2 generation onwards, 1000 mosquito Qualification Institutionnel (CQI), the ethics evaluation larvae that inherited a single copy of each TALEN were committee of INSERM (IRB00003888, FWA00005831). COPAS-purified to initiate the next generation. To isolate TEP1 Mosquitoes were reared and blood-fed on anesthetized mice in mutants, mosquitoes from this population were out-crossed to compliance with French and European laws on animal house the parental line. The progeny carried a single TALEN and was procedures (agreement E67-482-2 of the Direction of therefore no longer subjected to TALEN mutagenesis. Among Veterinary services of the French Ministry of Agriculture). this progeny, we screened single adult mosquitoes by PCR on one leg using the Phire direct animal tissue PCR kit (Thermo, Assembling the TALENs Fisher). The identified heterozygous mutants were individually TALENs targeting the sites shown in Figure 1 were crossed to the parental line. In the F2 progeny, we identified constructed by Golden Gate Cloning according to [9]. For homozygous mutants by leg PCR and pooled them to start TALEN assembly, we prepared two transgenesis-compatible destination vectors (annotated sequences provided in Text S1) distinct homozygous mutant families. encoding a Venus yellow fluorescent and a DsRed fluorescent transgenesis reporter gene, respectively. These vectors also RNAi and infection assays contain a phage ϕC31 attB site for genomic integration into RNAi silencing of TEP1 by double-stranded RNA injection transgenic lines harboring attP sites. The first module used in into the thorax of adult mosquitoes and mosquito infections on Golden Gate assembly provided the Vasa promoter mice carrying Plasmodium berghei GFP-con 259cl2 were characterized in [27]. The last module closed the TALEN performed as described [22]. To semi-automatically quantify assembly with either a FokI-DD or a FokI-RR domain [30], the number of oocysts in photographs of dissected mosquito designed for obligate heterodimerization of the two TALENs midguts, we used the “watershed segmentation” and “analyze and codon-optimized for A. gambiae. The annotated sequence particles” plugins of the ImageJ software after digital of the three plasmids is provided in Text S1. The TALEN C and subtraction of the image background and smoothing of the N-terminal domains flanking the repeat region were identical to signal. Oocyst counts obtained by this method are consistent those used in [7]. with those obtained by manual counting of oocysts. A. gambiae lines and mosquito transgenesis Immunoblotting The TALEN-encoding vectors were inserted by ϕC31 integrase-mediated transgenesis [41] into the genome of A. Whole-body mosquito extracts were obtained by grinding gambiae lines X1 and X13, which are derived from laboratory one adult female mosquito in 80 µl of protein sample buffer. strain G3. These two lines carry a PiggyBac transgene on The sample was denatured for 3 min at 95°C and centrifuged. chromosome II, containing an attP docking site. Individual 8 µl were loaded on 8% SDS-polyacrylamide gels. Hemolymph transgenic larvae carrying the inserted left or right TALEN samples were prepared as described [42], 10 µl of the samples genes at the X1 or X13 attP site were identified by their red or were loaded. Immunoblotting was performed using standard yellow fluorescence, respectively. Resulting adults were procedures [43] with rabbit polyclonal antibodies raised against crossed to their non-fluorescent parental line. In the F2 the prophenoloxidase PPO2 [44] or against the C-terminal half generation, fluorescent homozygous larvae were COPAS- of TEP1 [45]. selected [28] to establish stable TALEN-expressing lines. Supporting Information Identification of TEP1 Mutants In putative mutants (mosquitoes arising from parents Text S1. The nucleotide sequences of plasmids and expressing both TALENs), we PCR amplified a region of TEP1 building blocks used to construct the TALEN-expressing spanning the TALEN target site using Phusion or Phire Polymerases (Thermo, Fisher) and primer 5’- transgenesis vectors are provided. Building blocks not listed TCAACTTGGACATCAACAAGAAGGCCGA-3’ in combination here are published in references 5,7. with either 5’-GCATATCTTTGTGCCACACTTT-3’ or 5’- (DOCX) GCCACCGTAACGAATTTCCA-3’. The PCR product was digested with NcoI (Fermentas), the recognition site of which is Acknowledgements centrally located in the sequence cut by the TALENs. PCR products corresponding to mutant TEP1 alleles were not We thank the Boch and Bonas Laboratories (University of digested by NcoI. These PCR products were sequenced Halle, Germany) for sharing the TAL construction kit.

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 175 TALEN-induced TEP1 mutations in Anopheles gambiae

Author Contributions

Conceived and designed the experiments: EM AS. Performed the experiments: AS OT JS EM. Analyzed the data: AS EM EAL. Wrote the manuscript: EM AS EAL.

References

1. WHO (2012) Malaria Report. 21. Li T, Liu B, Spalding MH, Weeks DP, Yang B (2012) High-efficiency 2. Holt RA, Subramanian GM, Halpern A, Sutton GG, Charlab R et al. TALEN-based gene editing produces disease-resistant rice. Nat (2002) The genome sequence of the malaria mosquito Anopheles Biotechnol 30: 390-392. doi:10.1038/nbt.2199. PubMed: 22565958. gambiae. Science 298: 129-149. doi:10.1126/science.1076181. 22. Blandin S, Shiao SH, Moita LF, Janse CJ, Waters AP et al. (2004) PubMed: 12364791. Complement-like protein TEP1 is a determinant of vectorial capacity in 3. Catteruccia F (2007) Malaria vector control in the third millennium: the malaria vector Anopheles gambiae. Cell 116: 661-670. doi:10.1016/ progress and perspectives of molecular approaches (in press). Pest S0092-8674(04)00173-4. PubMed: 15006349. Manag Sci. 23. Blandin SA, Wang-Sattler R, Lamacchia M, Gagneur J, Lycett G et al. 4. Paton D, Underhill A, Meredith J, Eggleston P, Tripet F (2013) (2009) Dissecting the genetic basis of resistance to malaria parasites in Contrasted Fitness Costs of Docking and Antibacterial Constructs in Anopheles gambiae. Science 326: 147-150. doi:10.1126/science. the EE and EVida3 Strains Validates Two-Phase Genetic 1175241. PubMed: 19797663. Transformation System. PLOS ONE 8: e67364. doi:10.1371/ 24. Dong Y, Aguilar R, Xi Z, Warr E, Mongin E et al. (2006) Anopheles journal.pone.0067364. PubMed: 23840679. gambiae immune responses to human and rodent Plasmodium parasite 5. Boch J, Scholze H, Schornack S, Landgraf A, Hahn S et al. (2009) species. PLOS Pathog 2: e52. doi:10.1371/journal.ppat.0020052. Breaking the code of DNA binding specificity of TAL-type III effectors. PubMed: 16789837. Science 326: 1509-1512. doi:10.1126/science.1178811. PubMed: 25. Molina-Cruz A, Garver LS, Alabaster A, Bangiolo L, Haile A et al. 19933107. (2013) The human malaria parasite Pfs47 gene mediates evasion of 6. Moscou MJ, Bogdanove AJ (2009) A simple cipher governs DNA the mosquito immune system. Science 340: 984-987. doi:10.1126/ recognition by TAL effectors. Science 326: 1501. doi:10.1126/science. science.1235264. PubMed: 23661646. 1178817. PubMed: 19933106. 26. Baxter RH, Chang CI, Chelliah Y, Blandin S, Levashina EA et al. (2007) 7. Miller JC, Tan S, Qiao G, Barlow KA, Wang J et al. (2011) A TALE Structural basis for conserved complement factor-like function in the nuclease architecture for efficient genome editing. Nat Biotechnol 29: antimalarial protein TEP1. Proc Natl Acad Sci U S A 104: 11615-11620. 143-148. doi:10.1038/nbt.1755. PubMed: 21179091. doi:10.1073/pnas.0704967104. PubMed: 17606907. 8. Cermak T, Doyle EL, Christian M, Wang L, Zhang Y et al. (2011) 27. Papathanos PA, Windbichler N, Menichelli M, Burt A, Crisanti A (2009) Efficient design and assembly of custom TALEN and other TAL The vasa regulatory region mediates germline expression and maternal effector-based constructs for DNA targeting. Nucleic Acids Res, 39: transmission of proteins in the malaria mosquito Anopheles gambiae: a e82-. PubMed: 21493687. versatile tool for genetic control strategies. BMC Mol Biol 10: 65. doi: 9. Geissler R, Scholze H, Hahn S, Streubel J, Bonas U et al. (2011) 10.1186/1471-2199-10-65. PubMed: 19573226. Transcriptional activators of human genes with programmable DNA- 28. Marois E, Scali C, Soichot J, Kappler C, Levashina EA et al. (2012) specificity. PLOS ONE 6: e19509. doi:10.1371/journal.pone.0019509. High-throughput sorting of mosquito larvae for laboratory studies and PubMed: 21625585. for future vector control interventions. Malar J 11: 302. doi: 10. Reyon D, Tsai SQ, Khayter C, Foden JA, Sander JD et al. (2012) 10.1186/1475-2875-11-302. PubMed: 22929810. FLASH assembly of TALENs for high-throughput genome editing. Nat 29. Horn C, Wimmer EA (2000) A versatile vector set for animal Biotechnol 30: 460-465. doi:10.1038/nbt.2170. PubMed: 22484455. transgenesis. Dev Genes Evol 210: 630-637. doi:10.1007/ 11. Chen S, Oikonomou G, Chiu CN, Niles BJ, Liu J et al. (2013) A large- s004270000110. PubMed: 11151300. scale in vivo analysis reveals that TALENs are significantly more 30. Szczepek M, Brondani V, Büchel J, Serrano L, Segal DJ et al. (2007) mutagenic than ZFNs generated using context-dependent assembly. Structure-based redesign of the dimerization interface reduces the Nucleic Acids Res. toxicity of zinc-finger nucleases. Nat Biotechnol 25: 786-793. doi: 12. (2011) Move over ZFNs. Nat Biotechnol 29: 681-684. doi:10.1038/nbt. 10.1038/nbt1317. PubMed: 17603476. 1935. PubMed: 21822235. 31. Gupta A, Meng X, Zhu LJ, Lawson ND, Wolfe SA (2011) Zinc finger 13. Huang P, Xiao A, Zhou M, Zhu Z, Lin S et al. (2011) Heritable gene protein-dependent and -independent contributions to the in vivo off- targeting in zebrafish using customized TALENs. Nat Biotechnol 29: target activity of zinc finger nucleases. Nucleic Acids Res 39: 381-392. 699-700. doi:10.1038/nbt.1939. PubMed: 21822242. doi:10.1093/nar/gkq787. PubMed: 20843781. 14. Liu J, Li C, Yu Z, Huang P, Wu H et al. (2012) Efficient and specific 32. Le BV, Williams M, Logarajah S, Baxter RH (2012) Molecular basis for modifications of the Drosophila genome by means of an easy TALEN genetic resistance of Anopheles gambiae to Plasmodium: structural strategy. J Genet Genomics 39: 209-215. doi:10.1016/j.jgg. analysis of TEP1 susceptible and resistant alleles. PLOS Pathog 8: 2012.04.003. PubMed: 22624882. e1002958. PubMed: 23055931. 15. Carlson DF, Tan W, Lillico SG, Stverakova D, Proudfoot C et al. (2012) 33. Blandin S, Levashina EA (2004) Thioester-containing proteins and Efficient TALEN-mediated gene knockout in livestock. Proc Natl Acad insect immunity. Mol Immunol 40: 903-908. doi:10.1016/j.molimm. Sci U S A 109: 17382-17387. doi:10.1073/pnas.1211446109. PubMed: 2003.10.010. PubMed: 14698229. 23027955. 34. Catteruccia F, Levashina EA (2009) RNAi in the malaria vector, 16. Bedell VM, Wang Y, Campbell JM, Poshusta TL, Starker CG et al. Anopheles gambiae. Methods Mol Biol 555: 63-75. doi: (2012) In vivo genome editing using a high-efficiency TALEN system. 10.1007/978-1-60327-295-7_5. PubMed: 19495688. Nature 491: 114-118. doi:10.1038/nature11537. PubMed: 23000899. 35. Nsango SE, Abate L, Thoma M, Pompon J, Fraiture M et al. (2012) 17. Cade L, Reyon D, Hwang WY, Tsai SQ, Patel S et al. (2012) Highly Genetic clonality of Plasmodium falciparum affects the outcome of efficient generation of heritable zebrafish gene mutations using homo- infection in Anopheles gambiae. Int J Parasitol 42: 589-595. doi: and heterodimeric TALENs. Nucleic Acids Res 40: 8001-8010. doi: 10.1016/j.ijpara.2012.03.008. PubMed: 22554991. 10.1093/nar/gks518. PubMed: 22684503. 36. Degennaro M, McBride CS, Seeholzer L, Nakagawa T, Dennis EJ et al. 18. Watanabe T, Ochiai H, Sakuma T, Horch HW, Hamaguchi N et al. (2013) orco mutant mosquitoes lose strong preference for humans and (2012) Non-transgenic genome modifications in a hemimetabolous are not repelled by volatile DEET. Nature. insect using zinc-finger and TAL effector nucleases. Nat Commun 3: 37. Ghosh AK, Devenport M, Jethwaney D, Kalume DE, Pandey A et al. 1017. doi:10.1038/ncomms2020. PubMed: 22910363. (2009) Malaria parasite invasion of the mosquito salivary gland requires 19. Sung YH, Baek IJ, Kim DH, Jeon J, Lee J et al. (2013) Knockout mice interaction between the Plasmodium TRAP and the Anopheles saglin created by TALEN-mediated gene targeting. Nat Biotechnol 31: 23-24. proteins. PLOS Pathog 5: e1000265. PubMed: 19148273. doi:10.1038/nbt.2477. PubMed: 23302927. 38. Bhattacharyya MK, Kumar N (2001) Effect of xanthurenic acid on 20. Aryan A, Anderson MA, Myles KM, Adelman ZN (2013) TALEN-Based infectivity of Plasmodium falciparum to Anopheles stephensi. Int J Gene Disruption in the Dengue Vector Aedes aegypti. PLOS ONE 8: Parasitol 31: 1129-1133. doi:10.1016/S0020-7519(01)00222-3. e60082. doi:10.1371/journal.pone.0060082. PubMed: 23555893. PubMed: 11429178.

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 176 TALEN-induced TEP1 mutations in Anopheles gambiae

39. Volz J, Müller HM, Zdanowicz A, Kafatos FC, Osta MA (2006) A plasmodium response in the malaria mosquito Anopheles gambiae. genetic module regulates the melanization response of Anopheles to PLOS Biol 8: e1000434. PubMed: 20652016. Plasmodium. Cell Microbiol 8: 1392-1405. doi:10.1111/j. 43. Sambrook J, Fritsch EF, Maniatis T (1989) Molecular cloning A 1462-5822.2006.00718.x. PubMed: 16922859. laboratory manual. Cold Spring Harbor: CSH Laboratory Press. 40. Thailayil J, Magnusson K, Godfray HC, Crisanti A, Catteruccia F (2011) 44. Fraiture M, Baxter RH, Steinert S, Chelliah Y, Frolet C et al. (2009) Two Spermless males elicit large-scale female responses to mating in the mosquito LRR proteins function as complement control factors in the malaria mosquito Anopheles gambiae. Proc Natl Acad Sci U S A 108: TEP1-mediated killing of Plasmodium. Cell Host Microbe 5: 273-284. 13677-13681. doi:10.1073/pnas.1104738108. PubMed: 21825136. doi:10.1016/j.chom.2009.01.005. PubMed: 19286136. 41. Meredith JM, Basu S, Nimmo DD, Larget-Thiery I, Warr EL et al. (2011) 45. Levashina EA, Moita LF, Blandin S, Vriend G, Lagueux M et al. (2001) Site-specific integration and expression of an anti-malarial gene in Conserved Role of a Complement-like Protein in Phagocytosis transgenic Anopheles gambiae significantly reduces Plasmodium Revealed by dsRNA Knockout in Cultured Cells of the Mosquito, infections. PLOS ONE 6: e14587. doi:10.1371/journal.pone.0014587. Anopheles gambiae. Cell 104: 709-718. doi:10.1016/ PubMed: 21283619. S0092-8674(01)00267-7. PubMed: 11257225. 42. Rono MK, Whitten MM, Oulad-Abdelghani M, Levashina EA, Marois E (2010) The major yolk protein vitellogenin interferes with the anti-

PLOS ONE | www.plosone.org August 2013 | Volume 8 | Issue 8 | e74511 177 APPENDIX C

Engineering the control of mosquito-borne infectious diseases

178

Gabrieli et al. Genome Biology 2014, 15:535 http://genomebiology.com/2014/15/11/535

REVIEW Engineering the control of mosquito-borne infectious diseases Paolo Gabrieli1, Andrea Smidler2,3 and Flaminia Catteruccia2,4*

transgenic mosquitoes [7-10] to the creation of the first Abstract gene knock-outs [11-13], the discovery of genetic tools has Recent advances in genetic engineering are bringing revolutionized our ability to functionally study and edit the new promise for controlling mosquito populations mosquito genome. In the fight against infectious diseases, that transmit deadly pathogens. Here we discuss past vector populations can be modified using these tools in and current efforts to engineer mosquito strains that two principal ways: 1) they can be made refractory to are refractory to disease transmission or are suitable disease transmission by the introduction of genes with for suppressing wild disease-transmitting populations. anti-pathogenic properties; 2) they can be rendered sterile or modified in such ways that the population size will crash below the threshold necessary to support Introduction disease transmission (Figure 1) [14]. Both strategies have Mosquitoes transmit a variety of infectious agents that strengths and limitations that are inherent to their design are a scourge on humanity. Malaria, dengue fever, yellow and properties. fever, and other mosquito-borne infectious diseases infect Genetic engineering technologies include those that millions of people and account for hundreds of thousands allow heterologous gene expression and those that modify of deaths each year, posing a huge burden for public endogenous genes or entire portions of the mosquito health and on the economic growth of countries where genome.Herewereviewthegenetictoolsthatarecurrently these diseases are endemic [1]. Given the lack of effective in use and those that promise to become available in the vaccines against many mosquito-borne pathogens, near future, with particular focus on those techniques that national programs are heavily reliant on the use of are capable of reprogramming the genomes of field insecticides to control mosquito populations in order populations. We also discuss current field trials in which to stop disease transmission [2]. Unfortunately, the genetically modified mosquitoes are being released, and will alarming pace of emergence of insecticide resistance in mention ecological hurdles and potential environmental mosquitoes [3] is threatening chemical-based campaigns and regulatory issues stemming from the release of and is forcing scientists to develop alternative strategies genetically modified insects into the wild. to combat vector-borne diseases. Moreover, insecticide- treated bed nets and indoor residual sprays principally target mosquitoes that feed indoors at night and that rest First generation of anti-pathogenic strains inside houses, thereby neglecting those species that prefer The expression of exogenous genes - through the to bite and rest outdoors or at earlier hours of the day, and transposon-mediated integration of transgenes - was the inducing some degree of insecticide-avoidance behavior first genomic technology to be developed in mosquitoes, (behavioral resistance) in indoor-biting individuals [4-6]. and gave birth to the modern field of mosquito genome Recent major advances in the field of genetic engineering engineering [7-10]. In this initial system, different exogen- are providing an unprecedented opportunity to conceive ous ‘effector’ genetic elements are cloned between the and create designer mosquito strains in order to control transposon terminal repeats (usually using the PiggyBac natural vector populations. From the generation of the first transposon [10]) to form a synthetic element that, in the presence of the integrating enzyme transposase, inserts into * Correspondence: [email protected] the mosquito genome at quasi-random loci (Figure 2a). In 2 Department of Immunology and Infectious Diseases, Harvard School of order to identify successful transformants, synthetic Public Health, Avenue , Boston, MA 021155, USA 4Department of Microbiology, Perugia University, Perugia 06100, Italy transposons are generally designed to carry a fluorescent Full list of author information is available at the end of the article reporter construct, such as the green fluorescent protein

179 Gabrieli et al. Genome Biology 2014, 15:535 Page 2 of 9 http://genomebiology.com/2014/15/11/535

incorporate cargoes with anti-pathogenic properties to (a) Population suppression render mosquitoes refractory to disease transmission. Releasing sterile males into the population can Both Anopheles and Aedes mosquito species, the vectors cause transient or permanent population suppression of malaria and dengue, respectively, have been modified to reduce their vectorial capacity. To stop the development Large of Plasmodium parasites, the causative agents of malaria, scientists have developed transgenic Anopheles stephensi lines that express single chain variable fragment antibodies (scFvs) [21-23] or synthetic antimalarial factors [24,25] (Figure 2b). Transgenic lines that express ScFvs against the ookinete proteins Chitinase 1 and Pfs25 [38,39] Population size orthepredominantsurfaceproteinofthesporozoites, Extinction circumsporozoite protein [40,41], show reduced ookinete crossing of midgut walls or sporozoite invasion of the Time Sterile releases salivary glands, respectively. Similarly, An. stephensi (5-10x) strains have been generated that secrete the synthetic dodecapeptide SM1 (an acronym for salivary gland- and (b) Population replacement midgut-binding peptide 1) into the midgut lumen during blood feeding. SM1 binding to the epithelium - probably A small release of drive-containing individuals through a mosquito midgut receptor - prevents ookinetes will result in drive predominance from invading the midgut in the rodent malaria 1.0 0.0 Plasmodium berghei model, thereby reducing both the prevalence and the intensity of infection [24]. Additionally, the incorporation of bee venom phospholip- ase A2 into transgenic An. stephensi inhibits ookinete inva- sion of the midgut by modifying epithelial membranes [25]. Anopheles gambiae, the principal vector of malaria in

Drive frequency sub-Saharan Africa, has been engineered to ectopically Wild type frequency express the endogenous antimicrobial peptide cecropin A 0.0 1.0 [26] and the synthetic peptide Vida3 [27], a hybrid Time peptide based on natural antimicrobial peptide sequences Drive release that have strong activity against Plasmodium sporogonic Figure 1 Methods for the genetic control of vector populations. forms [28]. (a) Population suppression can be achieved by releasing large Different laboratories have also developed Anopheles numbers of males that render their wild female mates incapable of having viable progeny. This includes releasing either males that are strains modified in key endogenous cellular pathways that sterile and produce no progeny at all (as in sterile insect technique regulate parasite development, namely the insulin-growth (SIT)) [15] or males that pass on lethal transgenes to the next factor signaling (ISS) and the immune deficiency (IMD) generation, producing progeny that die before they can transmit pathways. In An. stephensi, overexpression of Akt, a disease (as in the release of insects carrying dominant lethals, RIDL) critical regulator of ISS, elicits mitochondrial dysfunction [16]. For SIT strategies, multiple releases of a large excess (5x to 10x) of sterile males relative to the target population are normally carried out that enhances parasite killing in the midgut, even if at over large areas. (b) Population replacement occurs when traits carried some cost to mosquito survival [42,43]. To overcome by a small number of engineered mosquitoes replace traits that fitness costs, an inhibitor of ISS, the phosphatase and naturally exist in field populations [17]. The desired engineered tensin homolog (PTEN), was instead overexpressed [44]. trait - for instance, an anti-pathogen gene that renders mosquitoes PTEN inhibits phosphorylation of the ISS protein FOXO, refractory to disease transmission - is driven to fixation in the field population using a genetic drive (as described in Figure 2h). and its expression blocks Plasmodium development by enhancing the integrity of the midgut barrier, although this causes an increase in the female lifespan with possible negative consequences for disease transmission [44]. In (GFP), that acts as a selectable marker [18]. The promoter another study, An. stephensi mosquitoes were engineered of choice for the expression of selectable markers is often to express the active form of the IMD-regulated NF-κB the neuronal 3xP3 promoter [19], which is expressed transcription factor Rel2-S. Rel2-S activates the expression during larval development allowing easy detection of of several antimicrobial and anti-Plasmodium peptides, fluorescence and facilitating high-throughput sorting and when overexpressed in the midgut and in the fat body, by automated live sorters [20]. Moreover, this system can it strongly inhibits parasite development [45].

180 Gabrieli et al. Genome Biology 2014, 15:535 Page 3 of 9 http://genomebiology.com/2014/15/11/535

(a) Transposons (b) Hyper-immune mosquitoes (c) RIDL Strains tTA tet

tRE tet Transcriptional activator ffector cargo GFP R E TR a tetracycline

Transposase Transposase Wild type tTA tTA tTA First generation tTA tTA tRE tet Transcriptional activator TR Effector cargo GFP TR No tetracycline Transgenic

(d) HEGs/ ZFNs/ TALENs/ CRISPR (e) Knock-out mutants (g) Sex distorters (h) Gene drives

Wild type DRIVE Wild type HEG White eye Wild type Wild type X-shredder Wild type

ZFN fokI

fokI ZFN

TALEN fokI

Second generation (f) Knock-ins at endogenous loci fokI TALEN HA epitope

Endogenous gene

Cas9 Endogenous gene HA epitope

Figure 2 Current and future genetic engineering technologies for vector control. (a) First-generation technologies make use of transposable elements to insert genetic cargo randomly into the genome. The transposable element is mobilized by a transposase enzyme produced by another plasmid, which recognizes and cleaves the terminal repeats (TR) of the transposon cassette and mediates insertion of the transposable element into the genome. Insertion is visualized using selectable markers such as the green fluorescent protein (GFP) [19]. (b) Mosquitoes can be engineered to carry anti-pathogenic effector genes that reduce the pathogen load [21-31]. In the figure, the effector gene blocks Plasmodium ookinete invasion of the midgut epithelium, preventing oocyst development. (c) Schematic of the RIDL system currently used for suppression of Aedes aegypti populations [16]. In the presence of tetracycline, expression of the tetracycline transactivator (tTA) is repressed. In the absence of tetracycline, tTA binds tothe tetracycline-responsive element (tRE) and drives its own expression in a positive feedback loop that leads to the accumulation of toxic levels of tTA. The progeny of released males carrying this transgene are not viable. Other combinations of inducible systems and toxic genes can be used in place of tTA and tRE to achieve population suppression. (d) Second generation technologies include HEGs, ZFNs, TALENs and CRISPR/Cas9 [11-13,32,33]. These technologies facilitate double-stranded DNA breaks in the genome at desired loci. (e) HEGs, TALENs and ZFNs have been used in Ae. aegypti and Anopheles gambiae to generate null mutants [11-13], including eye color mutants [11]. (f) ZFNs have been used to generate site-specific knock-ins of exogenous sequences in Ae. aegypti [34]. The figure illustrates a possible application for knock-in technology, which would enable scientists to fuse protein domains to the end of endogenous genes. These domains include those encoding fluorescent proteins or epitope tags, such as an HA tag (shown). (g) Sex distorter strains make use of an HEG, I-PpoI, to destroy sperm carrying an X chromosome (X-shredder), producing male-only populations. When mated to wild-type females, transgenic males sire only sons, potentially leading to population suppression [35]. (h) Gene drives are genetic elements that are inherited in a non-Mendelian fashion and can spread through populations. Gene drives using HEGs have been successfully developed to drive through laboratory mosquito populations [36], whereas evolutionarily stable drives enabled by CRISPR/Cas9 have been proposed [37].

Engineering pathogen resistance has not been limited aimed at spreading pathogen-refractory genes into wild to anophelines. Dengue virus infections in Aedes aegypti populations. mosquitoes have been attenuated by exploiting the natural antiviral RNA interference pathway. An inverted- First generation of sterile strains for population repeat RNA capable of forming double-stranded RNAs suppression that target the pre-membrane protein coding region of the Early transposon-based technology has been also used to DENV-2 serotype was expressed in the midgut [29] or in generate mosquito strains aimed at the suppression or the salivary glands [30]. This modification reduced viral elimination of vector populations through the release of titers by more than five-fold compared to those in control sterile males (the sterile insect technique (SIT)) [15]. mosquitoes. It should be noted, however, that multiple The alternative sister strategy is the release of insects dengue serotypes (as well as multiple human malaria para- carrying a dominant lethal (RIDL) modification [16]. SIT sites) exist, complicating population replacement efforts is based on the release of large numbers of sterile males,

181 Gabrieli et al. Genome Biology 2014, 15:535 Page 4 of 9 http://genomebiology.com/2014/15/11/535

usually sterilized with high doses of irradiation or chemical (Figure 2d). The flexibility of these tools resides in the use sterilants, that upon mating with field females produce no of protein precursors that can be designed to bind fertile progeny causing suppression or elimination of local sequences of interest within the mosquito genome [11-13]. populations (Figure 1a) [15]. The sterilization process Repetitive zinc finger (ZF) and transcription activator-like usually induces severe fitness costs in the male, such that effector (TALE) modules have been successfully fused to larger numbers of males than those initially predicted by the endonucleolytic domains of a type II endonuclease, simple models need to be released to achieve the desired normally FokI, to generate knock-out and knock-in level of suppression [46]. Genetic engineering can not only mutants [11-13,34] (Figure 2e,f). These modified nucleases enable high-throughput sorting of male-only populations cause site-specific double-stranded DNA breaks that can based on sex-specific fluorescent markers [47,48], but be repaired by the non-homologous end-joining (NHEJ) can also enable the design of strains in which specific pathway, an error-prone repair pathway that often sterility-inducing transgenes or genetic mutations have been results in small indels. As a basic proof-of-principle, introduced without causing the fitness costs associated with this technology has been used to generate eye-color irradiation [49,50]. The most successful RIDL example is mutants (Figure 2e) [11], but it can also help elucidate provided by the Ae. aegypti strain OX513A [16], which pathways that are important for vector competence. carries an inducible dominant genetic system that kills late For example, TALE nucleases (TALENs) have been larval stages. This system is composed of a gene encoding used in An. gambiae to generate null mutants of the the tetracycline transactivator (tTA) protein under the thioester-containing protein 1 (TEP1) gene, a complement- control of the tetracycline-responsive element (tRE). like factor that opsonizes Plasmodium parasites in the Binding of tetracycline to tTA prevents tTA from activating midgut and mediates their killing. Mutant strains are, transcription; when tetracycline is removed, tTA instead therefore, hyper-susceptible to Plasmodium infection [13], binds to tRE, thereby inducing its own expression via a and although not directly employable for malaria control, positive feedback loop. The accumulation of tTA is toxic to they allow detailed genetic analyses of anti-Plasmodium cells and ultimately leads to organismal death (Figure 2c). immune pathways. Similarly, the zinc-finger nuclease This repressible system allows the generation of males that (ZFN)-mediated knock-out of the odorant receptor co- are fertile in the laboratory but that, once released, sire receptor (ORCO) in Ae. aegypti has enabled the analysis unviable progeny upon mating with field females. These of pathways involved in host-seeking behavior for blood RIDL strains are already being released in different feeding [12], opening up new avenues for the development geographical locations as part of field trials. of mosquito repellents and attractants. In another study, A different approach, initially developed in Ae. aegypti the CO2 response of Ae. aegypti mosquitoes was analyzed and now transferred to Aedes albopictus and An. stephensi, in mutants that have a defect in the AaegGr3 gene, which is based on a bimodular system that severely impairs the encodes a subunit of the heteromeric CO2 receptor, functionality of the female flight muscles, disrupting the contributing to our understanding of mosquito attraction female’s ability to fly (fsRIDL) [51-53]. The first module to humans [34]. This mutant, the first knock-in to be consists of tTA under the control of the female-specific reported in mosquitoes, was generated by the disruptive Actin-4 transcriptional regulatory elements, which drive insertion of a fluorescent reporter gene into the AeagGr3 gene expression in the indirect flight muscles of locus. Such knock-in technology could also be used to female pupae. The second module comprises a lethal facilitate in-frame insertions of protein tags into genes of gene (Nipp1Dm or michelob_x in Ae. aegypti, VP16 in interest, further enabling the study of complex pathways Ae. albopictus and Nipp1Dm in An. stephensi) under the in mosquitoes (Figure 2f). control of tRE. In the absence of tetracycline, expression Homing endonucleases (HEGs) have also been success- ofthelethalgenespecificallyinthefemaleflight fully used to manipulate the mosquito genome [32,54,55]. muscles causes cell death and inability to fly. As HEGs are double-stranded DNases targeting large (12 to males are unaffected by the transgene, their release 40 bp) asymmetric recognition sites that occur extremely will generate flightless female progeny that are unable rarely in genomes [56]. An. gambiae strains have been to mate, bite, and transmit disease, eventually leading generated that express I-PpoI, a HEG that recognizes and to population suppression [51]. cuts a site in a multi-copy rDNA gene, which in this species is located exclusively on the X chromosome Second generation transgenesis provides [35,57]. When I-PpoI is expressed specifically during increased flexibility spermatogenesis, it cleaves these multiple target sequences New genome-editing tools now allow scientists to modify causing shredding of the paternal X chromosomes in sperm endogenous genes with increasing flexibility and ease, cells [35,57]. This feature was originally meant to generate and are being utilized in the laboratory with promising male-only populations by preventing fathers from transmit- results to reduce the vectorial capacity of mosquito vectors ting the X chromosome to embryos; but I-PpoIexpression

182 Gabrieli et al. Genome Biology 2014, 15:535 Page 5 of 9 http://genomebiology.com/2014/15/11/535

in sperm cells induces complete embryonic lethality, indicated that I-SceI could rapidly invade the receptive probably as a consequence of the shredding of the target strain, providing the first evidence of the gene-drive maternal X chromosome upon unintended transfer of capabilities of HEGs in mosquitoes [36]. the enzyme to the embryo [57]. These strains induce The range of applications enabled by HEGs and other a high level of infertility in large cage trials, as discussed nuclease-based technologies (ZFNs and TALENs) has some below [58]. An improved version of these strains, which limitations, especially in terms of specificity, flexibility and carries a less thermostable version of I-PpoI with reduced stability. For example, ZFNs do not always have the desired in vivo half life, has been generated that is instead active sequence specificity when assembled into arrays, which only in the testes, causing the specific shredding of limits the number of loci that can be targeted [62]. HEGs the paternal X chromosome in sperm without directly have been shown to cleave non-target sites (for a review affecting the embryo [35] (Figure 2g). The resulting see [63]), and laborious in vitro studies are necessary to sex-distorter strains produce >95% male offspring and generate new enzymes that have the required sequence are able to suppress wild-type mosquito populations specificity [64]. Furthermore, as these systems cut a single in laboratory cages [35]. genomic sequence at a time, new transgenic strains must be created for each target sequence. A new genome- Gene drives for population replacement engineering tool, CRISPR/Cas9 (for clustered regularly For the implementation of population replacement interspaced short palindromic repeats/CRISPR-associated strategies aimed at curbing mosquito-borne diseases, protein 9), has the potential to overcome these limitations the anti-pathogen constructs described above need to and stimulate the generation of effective gene drives for be driven genetically through natural populations so that vector control. Discovered as the molecular machinery of a the disease refractory traits will spread (Figure 2h). A bacterial acquired immune defense system [65], CRISPR/ number of artificial gene-drive systems capable of forcing Cas9 was soon co-opted to engineer the genomes of a wide their own spread in a non-Mendelian manner are being variety of organisms with high flexibility and efficiency [33]. developed that could be used for this purpose. In the Cas9 is an endonucleolytic protein that can recognize and model organism Drosophila melanogaster,thefirst cleave specific genomic sequences with the help of a gene-drive mechanism was developed on the basis of small artificial guide RNA (gRNA). When the gRNA a toxin-antidote system [59]. This synthetic system, and Cas9 form a complex, they catalyze DNA cleavage named Medea after the mythological figure of the upon recognition of the target site by the gRNA. The woman who killed her own children to take revenge reliance on easily designed gRNAs for the recognition of on her husband’s betrayal, is based on expression in target sequences results in a significant increase in the the zygote of a toxic gene, such as a microRNA number of genomic loci that can be cleaved when against a maternal mRNA essential for embryonic compared to other systems, as RNA-guided engineering development [59,60]. Transgenic females carry an does not require modification of the Cas9 protein ‘antidote’, that is, an allele of the gene that is insensitive to itself. Moreover, a number of loci can be targeted the toxin, allowing transgenic progeny to survive and simultaneously by providing multiple gRNAs, thereby spread the transgene. Although Medea has yet to be reducing the possible emergence of resistance to cleavage adapted to disease vectors, HEG-based technologies have [37]. Although research demonstrating the use of CRISPR/ been suggested and tested as gene drives in mosquitoes Cas9 in mosquitoes has yet to be published, it is likely that [36,61]. In this system, the drive encodes DNA-cutting this technology will soon enable the development of machinery that cleaves a wild-type target locus from a innovative and evolutionarily stable gene drives for transgene located at the homologous locus. Repair of the the control of vector-borne diseases. Nevertheless, further DNA break by homologous recombination causes the research is needed to demonstrate the improved perform- transgene to copy into the cleaved locus, causing a ance of this system over already existing technology, hemizygous cell to become homozygous for the transgene including minimizing off-target cleavage events and the (Figure 2h). If this mechanism occurs in the germline, possibility to revert the effects of the introduced gene the transgene can spread through the population, architectures [37]. potentially carrying an anti-pathogenic construct with it. Proof-of-principle use of HEGs to facilitate gene-drive Current field trials utilizing genetically modified mechanisms in An. gambiae was based on the I-SceI mosquitoes to fight disease enzyme, which targeted its own recognition sequence that Intensive research is ongoing to generate improved had been artificially introduced into a GFP reporter engineered strains that are suitable for vector-control gene [36]. Homing of the HEG into its target sequence, pre- programs, but the first generation of genetically modified viously integrated into the mosquito genome, would there- mosquitoes is already being released in the field. Since fore generate GFP null mutants. Small cage experiments 2009, the UK-based biotech company Oxitec has been

183 Gabrieli et al. Genome Biology 2014, 15:535 Page 6 of 9 http://genomebiology.com/2014/15/11/535

pushing the boundaries of genetic control by operating and must produce appropriate swarming and mating the first releases of transgenic Ae. aegypti RIDL strains to behavior [71]. suppress wild populations [66-69]. Their aim is to test the efficacy of these strains as a tool against dengue, a viral Ecological hurdles and environmental and disease for which no vaccine or effective drugs are available. regulatory considerations Repeated releases of the RIDL strain OX513A achieved a The implementation of genetically modified mosquitoes sizable reduction of wild populations, bringing new promise in vector control programs is challenged by a number for disease control. The first program was operated on of ecological, environmental and regulatory issues Grand Cayman Island, a British Territory in the Caribbean (summarized in Figure 3). Two crucial behavioral [66]. An average of 465 males/hectare (ha)/week were components of the released males are dispersal ability, released across 10 hectares over a 4-week period, which affects the possibility of targeting populations representing about 16% of the male population in the in impenetrable regions [68], and mating competitiveness, field. A total of 9.6% of fluorescent larvae were detected especially for species with complex sexual behaviors [72]. from eggs collected in ovitraps three weeks after the Indeed, the mating fitness of released males has proven to release, demonstrating that RIDL males could mate with be an important limiting factor in previous campaigns wild females and sire progeny, despite their reduced field aimed at reducing the size of Anopheles populations (for a competitiveness. A subsequent program, using 3,500 comprehensive discussion of these issues see [73] and males/ha/week, was carried out over a 23-week period references therein). Generally, anopheline species mate in and achieved 80% suppression of the wild population in a 16-ha area [67]. To accomplish this task, 3.3 million engineered males were reared and released, stressing the s need to optimize mass-rearing protocols [69]. OX513A al tri Effe d s cts el es non-ta on fi ven rget C was also released in a forested area in Pahang, Malaysia, i- titi sp o m e ec n e p ie s s m s ta and transgenic males were shown to live as long as their o H n d c o t n g r m a in E iz t col o o wild-type brothers from the same laboratory strain, even if e a og n n g ic t M a a it a e l l o c c h t r n u r i their dispersal ability was reduced [68]. Releases of y a r a n r d m n g o r le t s o s f a f e OX513A are currently being performed in Brazil [69], r r r E a e o c r o i o n p v b a GM lo d a y e V l t l h g s where additional trials are planned and the mosquito i l e y i e i a d e mosquito

c

b B e g M

t

r

a o

releases -

e l

production factory is being expanded. Large outdoor field a r

f

a

L

f b

e

s

i

r

o c

e

t

cages have also been employed to test the potential use of d

s

p i

v

s

i e

r D

the flightless Ae. aegypti fsRIDL strain [51,70]. This strain s

R i s t

S egulation y a m did not, however, achieve complete suppression of target f a et r y og an pr populations, suggesting that it may not be suitable for d s the R pecificity of is ce large-scale releases [70]. Reduced mating competitiveness k n m ta an ep ag cc of transgenic males probably contributed to test failure em ic a ent ubl but other explanations, including the different genetic P O backgrounds of released individuals and wild populations, ve ies rsigh genc have also been proposed [70]. t of regulatory a In the case of malaria vectors, large caged laboratory trials Figure 3 Challenges for the field release of transgenic have been established to test the mating competiveness of mosquitoes. This scheme summarizes the ecological, behavioral and sterile An. gambiae males carrying the HEG I-PpoI. When regulatory issues faced by disease control programs based on the released at 5- to 10-fold coverage in large cages, I-PpoI release of genetically modified mosquitoes. Ecological requirements males induced high levels of infertility, leading to the are shown in green, behavioral requirements in orange, while regulatory issues are presented in blue. Light-grey sections highlight suppression of caged populations in 4 to 5 weeks, operational tools that may be used to comply with the requirements. despite showing reduced mating competitiveness [58]. Behavioral requirements include key fitness parameters such as the Males carrying a less thermostable version of I-PpoI, which dispersal ability and mating competitiveness of released males, and causes sex distortion rather than male infertility, also can be tested in large laboratory cage trials and then in semi-field achieved elimination of caged populations within six settings to select the mosquito strains with the greatest probability of success. Ecological hurdles comprise heterogeneity in the genetics, generations when released at a 3x ratio [35]. Before behavior and natural habitats of vector species (biodiversity), and the field release of these strains is contemplated, their possible unintended side-effects on non-target species or on the competitive performance and sterilizing activity will ecosystem. Monitoring of these effects must be constantly in progress need to be tested in semi-field settings, such as those in the release phase. The risks, safety and specificity of the engineered provided by large outdoor enclosures, where mosqui- strains need to be evaluated by appropriate regulatory agencies, and early public engagement is a priority. toes are exposed to normal environmental conditions

184 Gabrieli et al. Genome Biology 2014, 15:535 Page 7 of 9 http://genomebiology.com/2014/15/11/535

elaborate swarms that are highly demanding energetically, engineering for mosquito control is to provide future gener- and in which males are subject to strong competition to ations with the undisputable benefits of a world free of find a mate [74]. Reduction of competitiveness can be vector-borne pathogens, while ensuring that possible caused by a number of factors including but not limited to unanticipated ecological and environmental consequences mass rearing, inbreeding, transposon expression and are eliminated. insertion sites in the genome [75-77]. The latter issue can now be partially overcome by utilizing ‘docking’ strains Abbreviations that are selected on the basis of limited fitness costs, using Cas9: CRISPR-associated protein 9; CRISPR: Clustered regularly interspaced short palindromic repeats; fsRIDL: Female-specific RIDL; GFP: Green the PhiC31 integration system [78]. fluorescent protein; gRNA: Guide RNA; HEG: Homing endonuclease; Other ecological features, including the biodiversity of IMD: Immune deficiency pathway; ISS: Insulin-growth factor signaling; native vector species, will also determine the success of NHEJ: Non-homologous end-joining; ORCO: Odorant receptor co-receptor; PTEN: Phosphatase and tensin homolog; RIDL: Release of insects carrying a a release campaign (Figure 3). Malaria transmission is dominant lethal; scFv: Single chain variable fragment antibody; SIT: Sterile supported by over 30 major primary vectors [79], many insect technique; SM1: Salivary gland- and midgut-binding peptide 1; of which are morphologically indistinguishable [80]. TALE: Transcription activator-like effector; TALEN: Transcription activator-like effector nuclease; TEP1: Thioester-containing protein 1; tRE: These often sympatric species exhibit distinct behaviors in Tetracycline-responsive element; tTA: Tetracycline transactivator; terms of mating, blood feeding and resting, and inhabit ZF: Zinc finger; ZFN: Zinc finger nucleases. diverse ecological niches, making their control extremely arduous [81]. Such complexity represents a significant Competing interests hurdle to the implementation of genetic engineering for The authors declare that they have no competing interests. malaria control; elimination of this disease solely by Acknowledgements transgenic means would require the simultaneous Work in our laboratories is supported by the European Research Council FP7 release of all malaria-transmitting species in any given ERC Starting Grant (project ‘Anorep’ (grant ID: 260897)), the area, a highly arduous task. By contrast, dengue virus (grant ID: 093553), and an NIH grant (grant ID: NIH 1R01AI104956-01A1) awarded to FC. infections are transmitted worldwide principally by Ae. aegypti and few other Aedes species. Although genetic Author details variations between different Ae. aegypti populations have 1University of Pavia, Pavia 27100, Italy. 2Department of Immunology and Infectious Diseases, Harvard School of Public Health, Avenue Louis Pasteur, been detected [82], pilot RIDL anti-dengue campaigns Boston, MA 021155, USA. 3Department of Genetics, Harvard Medical School, suggest that a single transgenic strain can adapt to differ- Avenue Louis Pasteur, Boston, MA 02115, USA. 4Department of Microbiology, ent ecological contexts [67-69]. The same strain could Perugia University, Perugia 06100, Italy. potentially be deployed to reduce the spread of the other viral diseases transmitted by these mosquitoes, such as yellow fever and Chikungunya, the latter being an emerging threat in the Americas [83]. References Finally, although the scope of this review is to describe 1. World Health Organization: The World Malaria Report 2013; 2013. 2. Hemingway J, Ranson H: Insecticide resistance in insect vectors of human the state of the art in transgenic technologies for disease disease. Annu Rev Entomol 2000, 45:371–391. control, we should mention that the release of genetically 3. Ranson H, N’Guessan R, Lines J, Moiroux N, Nkuni Z, Corbel V: Pyrethroid modified mosquitoes generates environmental and safety resistance in African anopheline mosquitoes: what are the implications for malaria control? Trends Parasitol 2011, 27:91–98. challenges that deserve to be meticulously addressed in 4. Taylor B: Changes in the feeding behaviour of a malaria vector, each individual case (outlined in Figure 3). Unintended Anopheles farauti Lav., following use of DDT as a residual spray in ecological side effects, accidental spread to non-target houses in the British Solomon Islands Protectorate. Trans R Entomol Soc Lond 1975, 127:277–292. species, and horizontal transfer of the transgenes are 5. Russell TL, Govella NJ, Azizi S, Drakeley CJ, Kachur SP, Killeen GF: Increased all unlikely but possible negative scenarios that can proportions of outdoor feeding among residual malaria vector and must be safely minimized [84]. Test trials under populations following increased use of insecticide-treated nets in rural Tanzania. Malar J 2011, 10:80. high containment levels and in confined laboratory 6. Sougoufara S, Diedhiou SM, Doucoure S, Diagne N, Sembene PM, Harry M, and semi-field settings should be used to determine Trape JF, Sokhna C, Ndiath MO: Biting by Anopheles funestus in broad specificity and safety of modified vectors, and constant daylight after use of long-lasting insecticidal nets: a new challenge to malaria elimination. Malar J 2014, 13:125. monitoring should occur during the release phase. This is 7. Coates CJ, Jasinskiene N, Miyashiro L, James AA: Mariner transposition and especially important when releasing gene-drive architec- transformation of the yellow fever mosquito, Aedes aegypti. Proc Natl tures that are capable of spreading through entire popula- Acad Sci U S A 1998, 95:3748–3751. 8. Jasinskiene N, Coates CJ, Benedict MQ, Cornel AJ, Rafferty CS, James AA, tions, such as those afforded by meiotic drives, HEGs and Collins FH: Stable transformation of the yellow fever mosquito, Aedes CRISPRs. The fast and exciting pace of progress provided aegypti, with the Hermes element from the housefly. Proc Natl Acad Sci by genetic-engineering technologiesrequiresanopenand USA1998, 95:3743–3747. 9. Catteruccia F, Nolan T, Loukeris TG, Blass C, Savakis C, Kafatos FC, Crisanti A: early discussion to engage regulatory agencies, the scientific Stable germline transformation of the malaria mosquito Anopheles community, and the public [85]. The end goal of genetic stephensi. Nature 2000, 405:959–962.

185 Gabrieli et al. Genome Biology 2014, 15:535 Page 8 of 9 http://genomebiology.com/2014/15/11/535

10. Grossman GL, Rafferty CS, Clayton JR, Stevens TK, Mukabayire O, Benedict MQ: 32. Windbichler N, Papathanos PA, Catteruccia F, Ranson H, Burt A, Crisanti A: Germline transformation of the malaria vector, Anopheles gambiae,with Homing endonuclease mediated gene targeting in Anopheles gambiae the piggyBac transposable element. Insect Mol Biol 2001, 10:597–604. cells and embryos. Nucleic Acids Res 2007, 35:5922–5933. 11. Aryan A, Anderson MA, Myles KM, Adelman ZN: TALEN-based gene 33. Hsu PD, Lander ES, Zhang F: Development and applications of CRISPR-Cas9 disruption in the dengue vector Aedes aegypti. PLoS One 2013, 8:e60082. for genome engineering. Cell 2014, 157:1262–1278. 12. DeGennaro M, McBride CS, Seeholzer L, Nakagawa T, Dennis EJ, Goldman C, 34. McMeniman CJ, Corfas RA, Matthews BJ, Ritchie SA, Vosshall LB: Multimodal Jasinskiene N, James AA, Vosshall LB: Orco mutant mosquitoes lose strong integration of carbon dioxide and other sensory cues drives mosquito preference for humans and are not repelled by volatile DEET. Nature 2013, attraction to humans. Cell 2014, 156:1060–1071. 498:487–491. 35. Galizi R, Doyle LA, Menichelli M, Bernardini F, Deredec A, Burt A, Stoddard BL, 13. Smidler AL, Terenzi O, Soichot J, Levashina EA, Marois E: Targeted Windbichler N, Crisanti A: A synthetic sex ratio distortion system for the mutagenesis in the malaria mosquito using TALE nucleases. PLoS One control of the human malaria mosquito. Nat Commun 2014, 5:3977. 2013, 8:e74511. 36. Windbichler N, Menichelli M, Papathanos PA, Thyme SB, Li H, Ulge UY, 14. Catteruccia F: Malaria vector control in the third millennium: progress and Hovde BT, Baker D, Monnat RJ Jr, Burt A, Crisanti A: A synthetic homing perspectives of molecular approaches. Pest Manag Sci 2007, 63:634–640. endonuclease-based gene drive system in the human malaria mosquito. 15. Knipling EF: Possibilities of insect control or eradication through the use Nature 2011, 473:212–215. of sexually sterile males. J Econ Entomol 1955, 48:902–904. 37. Esvelt KM, Smidler AL, Catteruccia F, Church GM: Concerning RNA-guided 16. Phuc HK, Andreasen MH, Burton RS, Vass C, Epton MJ, Pape G, Fu G, Condon KC, gene drives for the alteration of wild populations. Elife 2014, e03401. Scaife S, Donnelly CA, Coleman PG, White-Cooper H, Alphey L: Late-acting 38. Barr PJ, Green KM, Gibson HL, Bathurst IC, Quakyi IA, Kaslow DC: dominant lethal genetic systems and mosquito control. BMC Biol 2007, 5:11. Recombinant Pfs25 protein of Plasmodium falciparum elicits malaria 17. Wilke AB, Marrelli MT: Genetic control of mosquitoes: population transmission-blocking immunity in experimental animals. J Exp Med 1991, suppression strategies. Rev Inst Med Trop Sao Paulo 2012, 54:287–292. 174:1203–1208. 18. Chalfie M, Tu Y, Euskirchen G, Ward WW, Prasher DC: Green fluorescent 39. Li F, Templeton TJ, Popov V, Comer JE, Tsuboi T, Torii M, Vinetz JM: protein as a marker for gene expression. Science 1994, 263:802–805. Plasmodium ookinete-secreted proteins secreted through a common 19. Kokoza V, Ahmed A, Wimmer EA, Raikhel AS: Efficient transformation of micronemal pathway are targets of blocking malaria transmission. J Biol the yellow fever mosquito Aedes aegypti using the piggyBac Chem 2004, 279:26635–26644. transposable element vector pBac[3xP3-EGFP afm]. Insect Biochem Mol 40. Burkot TR, Da ZW, Geysen HM, Wirtz RA, Saul A: Fine specificities of Biol 2001, 31:1137–1143. monoclonal antibodies against the Plasmodium falciparum 20. Marois E, Scali C, Soichot J, Kappler C, Levashina EA, Catteruccia F: High- circumsporozoite protein: recognition of both repetitive and throughput sorting of mosquito larvae for laboratory studies and for non-repetitive regions. Parasite Immunol 1991, 13:161–170. future vector control interventions. Malar J 2012, 11:302. 41. de Lara CM, Coleman J, Beerntsen BT, Myles KM, Olson KE, Rocha E, Krettli AU, 21. Isaacs AT, Li F, Jasinskiene N, Chen X, Nirmala X, Marinotti O, Vinetz JM, James AA: Virus-expressed, recombinant single-chain antibody blocks James AA: Engineered resistance to Plasmodium falciparum development sporozoite infection of salivary glands in Plasmodium gallinaceum-infected in transgenic Anopheles stephensi. PLoS Pathog 2011, 7:e1002017. Aedes aegypti. Am J Trop Med Hyg 2000, 62:427–433. 22. Isaacs AT, Jasinskiene N, Tretiakov M, Thiery I, Zettor A, Bourgouin C, James AA: 42. Corby-Harris V, Drexler A, Watkins de Jong L, Antonova Y, Pakpour N, Transgenic Anopheles stephensi coexpressing single-chain antibodies resist Ziegler R, Ramberg F, Lewis EE, Brown JM, Luckhart S, Riehle MA: Plasmodium falciparum development. Proc Natl Acad Sci U S A 2012, Correction: activation of Akt signaling reduces the prevalence and 109:E1922–E1930. intensity of malaria parasite infection and lifespan in Anopheles stephensi 23. Sumitani M, Kasashima K, Yamamoto DS, Yagi K, Yuda M, Matsuoka H, mosquitoes. PLoS Pathog 2010, 6. doi:10.1371/annotation/738ac91f-8c41- Yoshida S: Reduction of malaria transmission by transgenic mosquitoes 4bf5-9a39-bddf0b777a89. expressing an antisporozoite antibody in their salivary glands. Insect Mol 43. Luckhart S, Giulivi C, Drexler AL, Antonova-Koch Y, Sakaguchi D, Napoli E, Biol 2013, 22:41–51. Wong S, Price MS, Eigenheer R, Phinney BS, Pakpour N, Pietri JE, Cheung K, 24. Ito J, Ghosh A, Moreira LA, Wimmer EA, Jacobs-Lorena M: Transgenic Georgis M, Riehle M: Sustained activation of Akt elicits mitochondrial anopheline mosquitoes impaired in transmission of a malaria parasite. dysfunction to block Plasmodium falciparum infection in the mosquito Nature 2002, 417:452–455. host. PLoS Pathog 2013, 9:e1003180. 25. Moreira LA, Ito J, Ghosh A, Devenport M, Zieler H, Abraham EG, Crisanti A, 44. Hauck ES, Antonova-Koch Y, Drexler A, Pietri J, Pakpour N, Liu D, Blacutt J, Nolan T, Catteruccia F, Jacobs-Lorena M: Bee venom phospholipase Riehle MA, Luckhart S: Overexpression of phosphatase and tensin homolog inhibits malaria parasite development in transgenic mosquitoes. J Biol improves fitness and decreases Plasmodium falciparum development in Chem 2002, 277:40839–40843. Anopheles stephensi. Microbes Infect 2013, 15:775–787. 26. Kim W, Koo H, Richman AM, Seeley D, Vizioli J, Klocko AD, O'Brochta DA: 45. Dong Y, Das S, Cirimotich C, Souza-Neto JA, McLean KJ, Dimopoulos G: Ectopic expression of a cecropin transgene in the human malaria vector Engineered anopheles immunity to Plasmodium infection. PLoS Pathog mosquito Anopheles gambiae (Diptera: Culicidae): effects on 2011, 7:e1002458. susceptibility to Plasmodium. J Med Entomol 2004, 41:447–455. 46. Barclay HJ: Mathematical models for the use of sterile insects. In Sterile 27. Meredith JM, Basu S, Nimmo DD, Larget-Thiery I, Warr EL, Underhill A, Insect Technique - Principles and Practice in Area-Wide Integrated Pest McArthur CC, Carter V, Hurd H, Bourgouin C, Eggleston P: Site-specific Management. Edited by Dyck VA, Hendrichs J, Robinson AS. Dordrecht, integration and expression of an anti-malarial gene in transgenic The Netherlands: Springer; 2005:147–174. Anopheles gambiae significantly reduces Plasmodium infections. 47. Catteruccia F, Benton JP, Crisanti A: An Anopheles transgenic sexing strain PLoS One 2011, 6:e14587. for vector control. Nat Biotechnol 2005, 23:1414–1417. 28. Arrighi RB, Nakamura C, Miyake J, Hurd H, Burgess JG: Design and activity 48. Smith RC, Walter MF, Hice RH, O'Brochta DA, Atkinson PW: Testis-specific of antimicrobial peptides against sporogonic-stage parasites causing expression of the beta2 tubulin promoter of Aedes aegypti and its murine . Antimicrob Agents Chemother 2002, 46:2104–2110. application as a genetic sex-separation marker. Insect Mol Biol 2007, 29. Franz AW, Sanchez-Vargas I, Adelman ZN, Blair CD, Beaty BJ, James AA, 16:61–71. Olson KE: Engineering RNA interference-based resistance to dengue virus 49. Andreasen MH, Curtis CF: Optimal life stage for radiation sterilization of type 2 in genetically modified Aedes aegypti. Proc Natl Acad Sci U S A Anopheles males and their fitness for release. Med Vet Entomol 2005, 2006, 103:4198–4203. 19:238–244. 30. Mathur G, Sanchez-Vargas I, Alvarez D, Olson KE, Marinotti O, James AA: 50. Massonnet-Bruneel B, Corre-Catelin N, Lacroix R, Lees RS, Hoang KP, Nimmo D, Transgene-mediated suppression of dengue viruses in the salivary Alphey L, Reiter P: Fitness of transgenic mosquito Aedes aegypti males glands of the yellow fever mosquito, Aedes aegypti. Insect Mol Biol 2010, carrying a dominant lethal genetic system. PLoS One 2013, 8:e62711. 19:753–763. 51. Fu G, Lees RS, Nimmo D, Aw D, Jin L, Gray P, Berendonk TU, White-Cooper H, 31. Li C, Marrelli MT, Yan G, Jacobs-Lorena M: Fitness of transgenic Anopheles Scaife S, Kim Phuc H, Marinotti O, Jasinskiene N, James AA, Alphey L: stephensi mosquitoes expressing the SM1 peptide under the control of a Female-specific flightless phenotype for mosquito control. Proc Natl vitellogenin promoter. J Hered 2008, 99:275–282. Acad Sci U S A 2010, 107:4550–4554.

186 Gabrieli et al. Genome Biology 2014, 15:535 Page 9 of 9 http://genomebiology.com/2014/15/11/535

52. Labbe GM, Scaife S, Morgan SA, Curtis ZH, Alphey L: Female-specific 74. Maiga H, Niang A, Sawadogo SP, Dabire RK, Lees RS, Gilles JR, Tripet F, flightless (fsRIDL) phenotype for control of Aedes albopictus. PLoS Negl Diabate A: Role of nutritional reserves and body size in Anopheles Trop Dis 2012, 6:e1724. gambiae males mating success. Acta Trop 2014, 132:S102–S107. 53. Marinotti O, Jasinskiene N, Fazekas A, Scaife S, Fu G, Mattingly ST, Chow K, 75. Catteruccia F, Godfray HC, Crisanti A: Impact of genetic manipulation on the Brown DM, Alphey L, James AA: Development of a population fitness of Anopheles stephensi mosquitoes. Science 2003, 299:1225–1227. suppression strain of the human malaria vector mosquito, Anopheles 76. Reed DH, Lowe EH, Briscoe DA, Frankham R: Fitness and adaptation in a stephensi. Malar J 2013, 12:142. novel environment: effect of inbreeding, prior environment, and lineage. 54. Traver BE, Anderson MA, Adelman ZN: Homing endonucleases catalyze Evolution 2003, 57:1822–1828. double-stranded DNA breaks and somatic transgene excision in Aedes 77. Baeshen R, Ekechukwu NE, Toure M, Paton D, Coulibaly M, Traore SF, Tripet F: aegypti. Insect Mol Biol 2009, 18:623–633. Differential effects of inbreeding and selection on male reproductive 55. Aryan A, Anderson MA, Myles KM, Adelman ZN: Germline excision of phenotype associated with the colonization and laboratory maintenance transgenes in Aedes aegypti by homing endonucleases. Sci Rep 2013, of Anopheles gambiae. Malar J 2014, 13:19. 3:1603. 78. Amenya DA, Bonizzoni M, Isaacs AT, Jasinskiene N, Chen H, Marinotti O, 56. Hafez M, Hausner G: Homing endonucleases: DNA scissors on a mission. Yan G, James AA: Comparative fitness assessment of Anopheles stephensi Genome 2012, 55:553–569. transgenic lines receptive to site-specific integration. Insect Mol Biol 2010, 57. Windbichler N, Papathanos PA, Crisanti A: Targeting the X chromosome 19:263–269. during spermatogenesis induces Y chromosome transmission ratio 79. Kiszewski A, Mellinger A, Spielman A, Malaney P, Sachs SE, Sachs J: A global distortion and early dominant embryo lethality in Anopheles gambiae. index representing the stability of malaria transmission. Am J Trop Med PLoS Genet 2008, 4:e1000291. Hyg 2004, 70:486–498. 58. Klein TA, Windbichler N, Deredec A, Burt A, Benedict MQ: Infertility 80. Coluzzi M, Sabatini A, Petrarca V, Di Deco MA: Chromosomal resulting from transgenic I-PpoI male Anopheles gambiae in large cage differentiation and adaptation to human environments in the Anopheles trials. Pathog Glob Health 2012, 106:20–31. gambiae complex. Trans R Soc Trop Med Hyg 1979, 73:483–497. 59. Chen CH, Huang H, Ward CM, Su JT, Schaeffer LV, Guo M, Hay BA: A 81. Ferguson HM, Dornhaus A, Beeche A, Borgemeister C, Gottlieb M, Mulla MS, synthetic maternal-effect selfish genetic element drives population Gimnig JE, Fish D, Killeen GF: Ecology: a prerequisite for malaria replacement in Drosophila. Science 2007, 316:597–600. elimination and eradication. PLoS Med 2010, 7:e1000303. 60. Akbari OS, Matzen KD, Marshall JM, Huang H, Ward CM, Hay BA: A synthetic 82. Powell JR, Tabachnick WJ: History of domestication and spread of Aedes – gene drive system for local, reversible modification and suppression of aegypti - a review. Mem Inst Oswaldo Cruz 2013, 108:11 17. insect populations. Curr Biol 2013, 23:671–677. 83. Staples JE, Fischer M: Chikungunya virus in the Americas - what a – 61. Burt A: Site-specific selfish genes as tools for the control and genetic vectorborne pathogen can do. N Engl J Med 2014, 371:887 889. engineering of natural populations. Proc Biol Sci 2003, 270:921–928. 84. David AS, Kaser JM, Morey AC, Roth AM, Andow DA: Release of genetically 62. Lam KN, Van Bakel H, Cote AG, van der Ven A, Hughes TR: Sequence engineered insects: a framework to identify potential ecological effects. – specificity is obtained from the majority of modular C2H2 zinc-finger Ecol Evol 2013, 3:4000 4015. arrays. Nucleic Acids Res 2011, 39:4680–4690. 85. Oye KA, Esvelt K, Appleton E, Catteruccia F, Church G, Kuiken T, Lightfoot SB, 63. Stoddard BL: Homing endonucleases from mobile group I introns: McNamara J, Smidler A, Collins JP: Regulating gene drives. Science 2014, – discovery to genome engineering. Mob DNA 2014, 5:7. 345:626 628. 64. Thyme SB, Boissel SJ, Arshiya Quadri S, Nolan T, Baker DA, Park RU, Kusak L, Ashworth J, Baker D: Reprogramming homing endonuclease specificity doi:10.1186/s13059-014-0535-7 through computational design and directed evolution. Nucleic Acids Res Cite this article as: Gabrieli et al.: Engineering the control of 2014, 42:2564–2576. mosquito-borne infectious diseases. Genome Biology 2014 15:535. 65. Barrangou R, Fremaux C, Deveau H, Richards M, Boyaval P, Moineau S, Romero DA, Horvath P: CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315:1709–1712. 66. Harris AF, Nimmo D, McKemey AR, Kelly N, Scaife S, Donnelly CA, Beech C, Petrie WD, Alphey L: Field performance of engineered male mosquitoes. Nat Biotechnol 2011, 29:1034–1037. 67. Harris AF, McKemey AR, Nimmo D, Curtis Z, Black I, Morgan SA, Oviedo MN, Lacroix R, Naish N, Morrison NI, Collado A, Stevenson J, Scaife S, Dafa’alla T, Fu G, Phillips C, Miles A, Raduan N, Kelly N, Beech C, Donnelly CA, Petrie WD, Alphey L: Successful suppression of a field mosquito population by sustained release of engineered male mosquitoes. Nat Biotechnol 2012, 30:828–830. 68. Lacroix R, McKemey AR, Raduan N, Kwee Wee L, Hong Ming W, Guat Ney T, Rahidah AAS, Salman S, Subramaniam S, Nordin O, Hanum ATN, Angamuthu C, Marlina Mansor S, Lees RS, Naish N, Scaife S, Gray P, Labbé G, Beech C, Nimmo D, Alphey L, Vasan SS, Han Lim L, Wasi AN, Murad S: Open field release of genetically engineered sterile male Aedes aegypti in Malaysia. PLoS One 2012, 7:e42771. 69. Carvalho DO, Nimmo D, Naish N, McKemey AR, Gray P, Wilke AB, Marrelli MT, Virginio JF, Alphey L, Capurro ML: Mass production of genetically modified Aedes aegypti for field releases in Brazil. J Vis Exp 2014, e3579. 70. Facchinelli L, Valerio L, Ramsey JM, Gould F, Walsh RK, Bond G, Robert MA, Lloyd AL, James AA, Alphey L, Scott TW: Field cage studies and progressive evaluation of genetically-engineered mosquitoes. PLoS Negl Trop Dis 2013, 7:e2001. 71. Knols BG, Njiru BN, Mathenge EM, Mukabana WR, Beier JC, Killeen GF: MalariaSphere: a greenhouse-enclosed simulation of a natural Anopheles gambiae (Diptera: Culicidae) ecosystem in western . Malar J 2002, 1:19. 72. Ferguson HM, John B, Ng'habi K, Knols BG: Redressing the sex imbalance in knowledge of vector biology. Trends Ecol Evol 2005, 20:202–209. 73. Benedict MQ, Robinson AS: The first releases of transgenic mosquitoes: an argument for the sterile insect technique. Trends Parasitol 2003, 19:349–355.

187 APPENDIX D

Concerning RNA-guide gene drives for the alteration of wild populations

188

FEATURE ARTICLE

elifesciences.org

EMERGING TECHNOLOGY Concerning RNA-guided gene drives for the alteration of wild populations

Abstract Gene drives may be capable of addressing ecological problems by altering entire populations of wild organisms, but their use has remained largely theoretical due to technical constraints. Here we consider the potential for RNA-guided gene drives based on the CRISPR nuclease Cas9 to serve as a general method for spreading altered traits through wild populations over many generations. We detail likely capabilities, discuss limitations, and provide novel precautionary strategies to control the spread of gene drives and reverse genomic changes. The ability to edit populations of sexual species would offer substantial benefits to humanity and the environment. For example, RNA-guided gene drives could potentially prevent the spread of disease, support agriculture by reversing pesticide and herbicide resistance in insects and weeds, and control damaging invasive species. However, the possibility of unwanted ecological effects and near-certainty of spread across political borders demand careful assessment of each potential application. We call for thoughtful, inclusive, and well-informed public discussions to explore the responsible use of this currently theoretical technology. DOI: 10.7554/eLife.03401.001

KEVIN M ESVELT*, ANDREA L SMIDLER, FLAMINIA CATTERUCCIA* AND GEORGE M CHURCH*

Introduction agricultural and environmental damage wrought by invasive species. Despite numerous advances, the field of molec- Scientists have long known of naturally occur- ular biology has often struggled to address key ring selfish genetic elements that can increase biological problems affecting public health and the odds that they will be inherited. This advan- the environment. Until recently, editing the tage allows them to spread through popula- genomes of even model organisms has been dif- tions even if they reduce the fitness of individual *For correspondence: ficult. Moreover, altered traits typically reduce ev- organisms. Many researchers have suggested [email protected] olutionary fitness and are consequently eliminated that these elements might serve as the basis for (KME); [email protected] by . This restriction has profoundly (FC); [email protected] (GMC) limited our ability to alter ecosystems through ‘gene drives’ capable of spreading engineered molecular biology. traits through wild populations (Craig et al., 1960; Reviewing editor: Diethard If we could develop a general method of Wood et al., 1977; Sinkins and Gould, 2006; Tautz, Max Planck Institute for ensuring that engineered traits would instead be Burt and Trivers, 2009; Alphey, 2014). Austin Evolutionary Biology, Germany favored by natural selection, then those traits Burt was the first to propose gene drives based Copyright Esvelt et al. This article could spread to most members of wild populations on site-specific ‘homing’ endonuclease genes is distributed under the terms of the over many generations. This capability would over a decade ago (Burt, 2003). These genes Creative Commons Attribution License, allow us to address several major world problems, bias inheritance by cutting the homologous chro- which permits unrestricted use and redistribution provided that the original including the spread of insect-borne diseases, the mosome, inducing the cell to copy them when it author and source are credited. rise of pesticide and herbicide resistance, and the repairs the break. Several efforts have focused

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 1 of 21 189 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

on the possibility of using gene drives targeting his proposal. Indeed, the well-recognized poten- mosquitoes to block malaria transmission (Scott tial for gene drives to combat vector-borne dis- et al., 2002; Windbichler et al., 2007, 2008, eases such as malaria and dengue virtually ensures 2011; Li et al., 2013a; Galizi et al., 2014). that this strategy will eventually be attempted in However, development has been hindered by the mosquitoes. difficulty of engineering homing endonucleases While considerable scholarship has been to cut new target sequences (Chan et al., 2013a; devoted to the question of how gene drives Thyme et al., 2013; Takeuchi et al., 2014). might be safely utilized in mosquitoes (Scott Attempts to build gene drives with more easily et al., 2002; Touré et al., 2004; Benedict et al., retargeted zinc-finger nucleases and TALENs suf- 2008; Marshall, 2009; UNEP, 2010; Reeves fered from instability due to the repetitive nature et al., 2012; David et al., 2013; Alphey, 2014), of the genes encoding them (Simoni et al., 2014). few if any studies have examined the potential The recent discovery and development of the ecological effects of gene drives in other species. RNA-guided Cas9 nuclease has dramatically After all, constructing a drive to spread a partic- enhanced our ability to engineer the genomes of ular genomic alteration in a given species was diverse species. Originally isolated from ‘CRISPR’ simply not feasible with earlier genome editing acquired immune systems in bacteria, Cas9 is a methods. Disconcertingly, several published gene non-repetitive enzyme that can be directed to cut drive architectures could lead to extinction or almost any DNA sequence by simply expressing a other hazardous consequences if applied to ‘guide RNA’ containing that same sequence. In sensitive species, demonstrating an urgent need little more than a year following the first dem- for improved methods of controlling these ele- onstrations in human cells, it has enabled gene ments. After consulting with experts in many fields insertion, deletion, and replacement in many dif- as well as concerned environmental organizations, ferent species (Bassett et al., 2013; Cho et al., we are confident that the responsible develop- 2013; Cong et al., 2013; DiCarlo et al., 2013; ment of RNA-guided gene drive technology is Friedland et al., 2013; Gratz et al., 2013; Hu best served by full transparency and early engage- et al., 2013; Hwang et al., 2013; Jiang et al., ment with the public. 2013a, 2013b; Jinek et al., 2013; Li et al., Here we provide brief overviews of gene 2013b; Mali et al., 2013c; Tan et al., 2013; drives and Cas9-mediated genome engineering, Upadhyay et al., 2013; Wang et al., 2013b). detail the mechanistic reasons that RNA-guided Building RNA-guided gene drives based on gene drives are likely to be effective in many the Cas9 nuclease is a logical way to overcome species, and outline probable capabilities and the targeting and stability problems hindering limitations. We further propose novel gene drive gene drive development. Less obvious is the architectures that may substantially improve our extent to which the unique properties of Cas9 control over gene drives and their effects, discuss are well-suited to overcoming other molecular possible applications, and suggest guidelines and evolutionary challenges inherent to the con- for the safe development and evaluation of this struction of safe and functional gene drives. promising but as yet unrealized technology. A dis- We submit that Cas9 is highly likely to enable cussion of risk governance and regulation intended scientists to construct efficient RNA-guided gene specifically for policymakers is published separately drives not only in mosquitoes, but in many other (Oye et al., 2014). species. In addition to altering populations of insects to prevent them from spreading disease Natural gene drives (Curtis, 1968), this advance would represent an entirely new approach to ecological engineering In nature, certain genes ‘drive’ themselves through with many potential applications relevant to human populations by increasing the odds that they will health, agriculture, biodiversity, and ecological be inherited (Burt and Trivers, 2009). Examples science. include endonuclease genes that copy them- The first technical descriptions of endonuclease selves into chromosomes lacking them (Burt gene drives were provided by Austin Burt in his and Koufopanou, 2004), segregation distorters landmark proposal to engineer wild populations that destroy competing chromosomes during more than a decade ago (Burt, 2003). Any of the meiosis (Lyttle, 1991), transposons that insert rapidly expanding number of laboratories with copies of themselves elsewhere in the genome expertise in Cas9-mediated genome engineering (Charlesworth and Langley, 1989), Medea ele- could attempt to build a gene drive by substituting ments that eliminate competing siblings who do Cas9 for the homing endonucleases described in not inherit them (Beeman et al., 1992; Chen

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 2 of 21 190 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

et al., 2007), and heritable microbes such as Wood and Newton, 1991). In this model, the Y Wolbachia (Werren, 1997). chromosome (or its equivalent in other sex- determination systems) would encode an endo- Endonuclease gene drives nuclease that cuts and destroys the X chromosome during male meiosis, thereby ensuring that Natural homing endonuclease genes exhibit most viable sperm contain a Y chromosome drive by cutting the corresponding locus of chro- (Newton et al., 1976; Wood and Newton, mosomes lacking them. This induces the cell to 1991; Windbichler et al., 2007, 2008; Galizi repair the break by copying the nuclease gene et al., 2014). The progressively dwindling number onto the damaged chromosome via homologous of females will culminate in a population crash recombination (Figure 1A) (Burt and Koufopanou, or extinction (Craig et al., 1960; Lyttle, 1977; 2004). The copying process is termed ‘homing’, Burt, 2003; Schliekelman et al., 2005; Deredec while the endonuclease-containing cassette that et al., 2008, 2011; North et al., 2013; Burt, is copied is referred to as a ‘gene drive’ or simply 2014). a ‘drive’. Because copying causes the fraction of Whether a standard gene drive will spread offspring that inherit the cassette to be greater than through a target population depends on molec- 1/2 (Figure 1B), these genes can drive through ular factors such as homing efficiency, fitness a population even if they reduce the reproductive cost, and evolutionary stability (Marshall and fitness of the individual organisms that carry them. Hay, 2012); only the rate of spread is determined Over many generations, this self-sustaining pro- 187 by the mating dynamics, generation time, and cess can theoretically allow a gene drive to spread other characteristics of the target population. from a small number of individuals until it is pre- In contrast, models suggest that the deleterious sent in all members of a population. and complex effects of genetic load and sex- biasing suppression drives render them more Engineered gene drives sensitive to population-specific ecological vari- To build an endonuclease gene drive, an endonu- ables such as density-dependent selection (Burt, clease transgene must be inserted in place of a 2003; Schliekelman et al., 2005; Huang et al., natural sequence that it can cut. If it can efficiently 2007; Deredec et al., 2008; Marshall, 2009; cut this sequence in organisms with one transgene Yahara et al., 2009; Deredec et al., 2011; Alphey and one natural locus, reliably induce the cell to and Bonsall, 2014). copy the transgene, and avoid being too costly to No engineered endonuclease gene drive the organism, it will spread through susceptible capable of spreading through a wild population wild populations. has yet been published. However, the Crisanti and Standard drives spread genomic changes and Russell laboratories have constructed gene drives associated traits through populations. Burt‘s orig- that can only spread through laboratory mosquito inal study proposed using them to drive the (Windbichler et al., 2011) and fruit fly (Chan spread of other transgenes or to disrupt existing et al., 2011; Simoni et al., 2014) populations that genes (Figure 2A,B) (Burt, 2003). The gene drive have been engineered to contain the endonucle- copying step can take place immediately upon ase cut site. The Burt and Crisanti laboratories are fertilization (Figure 2C) or occur only in germline attempting to build a male-biasing suppression cells that are immediate precursors to sperm or drive using an endonuclease that serendipitously eggs, leaving most of the organism‘s somatic cuts a conserved sequence repeated hundreds cells with only one copy of the drive (Figure 2D). of times in the X chromosome of the mosquito Suppression drives reduce the size of the targeted Anopheles gambiae (Windbichler et al., 2007, population. Austin Burt outlined an elegant strategy 2008; Galizi et al., 2014). If successful, their work involving the use of gene drives to disrupt genes promises to substantially reduce the population of that cause infertility or lethality only when both this important malaria vector. copies are lost (Burt, 2003). These ‘genetic load’ All engineered gene drives based on homing drives would spread rapidly through minimally endonucleases cut the natural recognition site of impaired heterozygotes when rare, and eventu- the relevant enzyme. Despite early hopes, it has ally cause the population to crash or even become proven difficult to engineer homing endonucle- extinct due to the accumulated load of recessive ases to cleave new target sequences. Numerous mutations. A second approach would mimic natu- laboratories have sought to accomplish this goal rally occurring ‘meiotic’ or ‘gametic’ drives that bias for well over a decade with only a few recent suc- the sex ratio (Craig et al., 1960; Hickey and Craig, cesses (Chan et al., 2013a; Thyme et al., 2013; 1966; Hamilton, 1967; Newton et al., 1976; Takeuchi et al., 2014). More recently, a team

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 3 of 21 191 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

cutting and homing, both declined in effectiveness over time due to the evolutionary instability of the modular repeats inherent to those proteins. These early attempts demonstrate that it is possible to build synthetic gene drives, but also emphasize the importance of cutting any desired gene and remaining stable during copying. The recent discovery of the RNA-guided Cas9 nuclease represents a possible solution.

RNA-guided genome editing via the Cas9 nuclease One straightforward method of genome editing relies on the same mechanism employed by endonuclease gene drives: cut the target gene and supply an edited version for the cell to use as a template when it fixes the damage. Most eukar- yotic genome engineering over the past decade was accomplished using zinc-finger nucleases (Urnov et al., 2005) or TALENs (Christian et al., 2010), both of which are modular proteins that can be redesigned or evolved to target new sequences, albeit only by specialist laboratories (Esvelt and Wang, 2013). Genome editing was democratized by the discovery and adaptation of Cas9, an enzyme that can be programmed to cut target DNA sequences specified by a guiding RNA molecule (Deltcheva et al., 2011; Jinek et al., 2012; Cho et al., 2013; Cong et al., 2013; Jinek et al., 2013; Mali et al., 2013c). Cas9 is a component of Type II CRISPR acquired immune systems in bacteria, which Figure 1. The spread of endonuclease gene drives. allow cells to ‘remember’ the sequences of pre- (A) When an organism carrying an endonuclease gene viously encountered viral genomes and protect drive (blue) mates with a wild-type organism (grey), the themselves by recognizing and cutting those gene drive is preferentially inherited by all offspring. sequences if encountered again. They accom- This can enable the drive to spread until it is present plish this by incorporating DNA fragments into in all members of the population–even if it is mildly a memory element, transcribing it to produce deleterious to the organism. (B) Endonuclease gene RNAs with the same sequence, and directing Cas9 drives are preferentially inherited because the endonu- to cut any matching DNA sequences (Deltcheva clease cuts the homologous wild-type chromosome. et al., 2011). The only restriction is that Cas9 When the cell repairs the break using homologous will only cut target ‘protospacer’ sequences that recombination, it must use the gene drive chromosome as a repair template, thereby copying the drive onto the are flanked by a protospacer-adjacent motif wild-type chromosome. If the endonuclease fails to cut (PAM) at the 3′ end. The most commonly used or the cell uses the competing non-homologous Cas9 ortholog has a PAM with only two required end-joining repair pathway, the drive is not copied, bases (NGG) and therefore can cut protospac- so efficient gene drives must reliably cut when ers found approximately every 8 base pairs homology-directed repair is most likely. (Jinek et al., 2012). DOI: 10.7554/eLife.03401.002 Remarkably, it is possible to direct Cas9 to cut a specific protospacer in the genome using constructed new versions of the fruit fly gene drive only a single guide RNA (sgRNA) less than 100 using modular zinc-finger nucleases or TALENs in base pairs in length (Jinek et al., 2012). This place of the homing endonuclease (Simoni et al., guide RNA must begin with a 17-20 base pair 2014), both of which can be engineered to cut ‘spacer’ sequence identical to the targeted proto­ new target sequences. While initially successful at spacer sequence in the genome (Fu et al., 2014).

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 4 of 21 192 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations Because RNA-guided genome editing relies on exactly the same copying mechanism as endonuclease gene drives, it is reasonable to ask whether it might be possible to build gene drives based on Cas9. In principle, RNA-guided gene drives might be capable of spreading almost any genomic alteration that can be gener- ated using Cas9 through sexually reproducing populations.

Will RNA-guided gene drives enable us to edit the genomes of wild populations? Although we cannot be certain until we try, cur- rent evidence suggests that RNA-guided gene drives will function in some and possibly most sexually reproducing species. Learning how to Figure 2. Consequences and timing of gene drive insert a drive into the germline and optimize its replication. (A) Gene drives can carry other genes function in each new species will likely require with them as cargo. For example, a transgene that months to years depending on generation blocks malaria transmission could be driven through length, with subsequent drives in the same spe- wild mosquito populations. There is no selection to cies taking less time. Because inserting the maintain the function of a cargo gene. (B) Gene drive into the germline with Cas9 involves the drives can disrupt or replace other genes. For example, a drive might replace a mosquito gene same molecular copying process as the drive important for malaria transmission. Because it itself will utilize, successful insertion may produce cannot spread without disrupting the target gene, a working if not particularly efficient RNA- this strategy is evolutionarily stable. (C) If homing guided gene drive. But if population-level engi- occurs in the zygote or early embryo, all organisms neering is to become a reality, all molecular that carry the drive will be homozygous in all of their factors relevant to homing – cutting, specificity, tissues. (D) If homing occurs in the late germline cells copying, and evolutionary robustness – must be that contribute to sperm or eggs, the offspring will considered. Below, we provide a detailed technical remain heterozygous in most tissues and avoid the analysis of the extent to which Cas9 can address consequences of drive-induced disruptions. each of these challenges. Capabilities, limitations, DOI: 10.7554/eLife.03401.003 control strategies, and possible applications are discussed in subsequent sections.

The process of editing a target gene involves Cutting choosing protospacers within the gene, building one or more guide RNAs with matching spac- The first requirement for every endonuclease gene ers, and delivering Cas9, guide RNAs, and an drive is to cut the target sequence. Incomplete edited repair template lacking those proto- cutting was a problem for the homing endonuclease spacers into the cell (Figure 3). drive constructed in transgenic mosquitoes (72% Cas9 is efficient enough to cut and edit multiple cutting) and also for the homing endonuclease, genes in a single experiment (Li et al., 2013c; zinc-finger nuclease, and TALEN drives in fruit Wang et al., 2013a). The enzyme is active in flies (37%, 86%, and 70% cutting) (Windbichler a wide variety of organisms and is also quite spe- et al., 2011; Chan et al., 2013b; Simoni et al., cific, cutting only protospacers that are nearly 2014). The simplest way to increase cutting is to identical to the spacer sequence of the guide RNA target multiple adjacent sequences. However, this (Hsu et al., 2013; Mali et al., 2013a; Pattanayak is impractical for homing endonucleases and quite et al., 2013). Moreover, methods that allow difficult for zinc-finger nucleases and TALENs, Cas9 to bind but not cut enable the expression as each additional sequence requires a new of target genes to be regulated by selectively nuclease protein to be engineered or evolved recruiting regulatory proteins attached to Cas9 and then co-expressed. or the guide RNA (Gilbert et al., 2013; Mali et al., In contrast, the RNA-guided Cas9 nuclease 2013a). All of these applications were devel- can be readily directed to cleave additional oped within the last two years. sequences by expressing additional guide RNAs

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 5 of 21 193 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

out of six tested guide RNAs (Kondo and Ueda, 2013). The two least effective guide RNAs indi- vidually cut at rates exceeding 12% and 56%, but exhibited cutting rates above 91% when com- bined. Using more than two guide RNAs should further enhance cutting. The notable success of Cas9-based genome engineering in many dif- ferent species, including studies that targeted every gene in the genome (Shalem et al., 2014; Wang et al., 2014), demonstrates that most sequences can be efficiently targeted independent of species and cell type. Thus, RNA-guided gene drives should be capable of efficiently cutting any given gene.

Specificity Because cutting other sites in the genome may seriously compromise the fitness of the organism, the second requirement is to avoid cutting non- targeted sequences. Figure 3. RNA-guided genome editing via Cas9. The While several studies have reported that Cas9 Cas9 nuclease protein and guide RNA must first be is prone to cutting off-target sequences that are delivered into the target cell. This is often accom- closely related to the target (Fu et al., 2013; Hsu plished by transfecting DNA expression plasmids, but et al., 2013; Mali et al., 2013a; Pattanayak et al., delivering RNA is also effective. The guide RNA directs 2013), more recent developments and strategies Cas9 to bind target DNA ‘protospacer’ sequences that designed to improve specificity (Mali et al., 2013a; match the ‘spacer’ sequence within the guide RNA. Fu et al., 2014; Guilinger et al., 2014; Tsai et al., Protospacers must be flanked by an appropriate 2014) have demonstrated that the off-target rate protospacer-adjacent motif (PAM), which is NGG for the can be reduced to nearly undetectable levels most commonly used Cas9 protein (Jinek et al., 2012). If the spacer and protospacer are identical or have only (Figure 4). Notably, Cas9 does not appear to rep- a few mismatches at the 5′ end of the spacer, Cas9 resent a noticeable fitness burden when expressed will cut both strands of DNA, creating a blunt-ended at a moderate level in fruit flies with or without double-strand break. If supplied with a repair template guide RNAs (Kondo and Ueda, 2013). Organisms containing the desired changes and homology to the with larger genomes may require more careful tar- sequences on either side of the break, the cell may get site selection due to the increased number of use homologous recombination to repair the break potential off-target sequences present. by incorporating the repair template into the chromo- some. Otherwise, the break will be repaired by non- Copying homologous end-joining, resulting in gene disruption. Cas9 cutting is efficient enough to alter both chromo- The third and most challenging requirement somes at the same time and/or to edit multiple genes involves ensuring that the cut sequence is repaired at once (Li et al., 2013; Wang et al., 2013a). If the using homologous recombination (HR) to copy the cell being edited is a germline cell that gives rise to drive rather than the competing non-homologous eggs or sperm, the changes can be inherited by future end-joining (NHEJ) pathway (Figure 4). HR rates generations. are known to vary across cell types (Mali et al., DOI: 10.7554/eLife.03401.004 2013c), developmental stages (Fiorenza et al., 2001; Preston et al., 2006), species (Chan et al., (Figure 4). The sequences of these additional 2013b), and the phase of the cell cycle (Saleh- guide RNAs can be altered so as to avoid creating Gohari and Helleday, 2004). For example, the unstable repeats within the drive cassette endonuclease gene drive in mosquitoes was cor- (Nishimasu et al., 2014; Simoni et al., 2014). rectly copied following ∼97% of cuts (Windbichler Including more guides has been demonstrated to et al., 2011), while a similar drive in fruit flies was improve upon already high rates of cutting. For initially copied only 2% of the time (Chan et al., example, fruit flies expressing both Cas9 and 2011) and never rose above 78% even with exten- guide RNAs in their germline exhibited target sive combinatorial optimization of promoter and 3′ cutting rates exceeding 85–99% in males for four untranslated region. This difference is presumably

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 6 of 21 194 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Figure 4. Technical advantages of RNA-guided gene drives. Clockwise from lower left: The targeting flexibility of Cas9 permits the exclusive selection of target sequences with few potential off-targets in the genome. Targeting multiple sites increases the cutting frequency and hinders the evolution of drive resistant alleles, which must accumulate mutations at all of the sites. The Cas9 nuclease is can be quite specific in the sequences that it targets; fruit flies do not exhibit notable fertility or fitness defects resulting from off-target cutting when both Cas9 nuclease and guide RNAs are expressed in the germline (Kondo and Ueda, 2013). Choosing target sites with few or no close relatives in the genome, using truncated guide RNAs (Fu et al., 2014), employing paired Cas9 nickases (Mali et al., 2013a) instead of nucleases, or utilizing Cas9-FokI fusion proteins (Guilinger et al., 2014; Tsai et al., 2014) can further increase specificity. Several of these strategies can reduce the off-target mutation rate to borderline undetectable levels (Fu et al., 2014; Guilinger et al., 2014; Tsai et al., 2014). The frequency at which the drive is correctly copied might be increased by using Cas9 as a transcriptional regulator to activate HR genes and repress NHEJ genes (Gilbert et al., 2013; Mali et al., 2013a) (Figure 4—figure supplement 1). By choosing target sites within an essential gene, any non-homologous end-joining event that deletes all of the target sites will cause lethality rather than creating a drive-resistant allele, further increasing the evolutionary robustness of the RNA-guided gene drive. Other options include using distinct promoters and guide RNAs to avoid repetitiveness and increase stability (Figure 4—figure supplement )2 or employing newly characterized, engineered, or evolved Cas9 variants with improved properties (Esvelt et al., 2011; Mali et al., 2013b). These optimization strategies have also been summarized in tabular form with additional details (Figure 4—figure supplement 3). DOI: 10.7554/eLife.03401.005 The following figure supplements are available for figure 4: Figure supplement 1. Enhancing drive copying by regulating endogenous genes. DOI: 10.7554/eLife.03401.006 Figure supplement 2. Repetitiveness and evolutionary stability of multiple guide RNAs. DOI: 10.7554/eLife.03401.007 Figure supplement 3. Table of known technological advances that might be adapted to optimize gene drive efficiency. DOI: 10.7554/eLife.03401.008

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 7 of 21 195 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

due to a lower rate of HR in fruit fly spermatocytes long enough to accumulate mutations at all of relative to mosquitoes (Chan et al., 2013b). Ideally, the sites so long as cutting rates are high (Burt, drives should be activated only in germline cells at 2003). However, very large populations – such developmental stages with a high rate of HR, but as those of some insects – might require unfea- this may be challenging in many species. sibly large numbers of guide RNAs to prevent Copying efficiencies may also depend on whether resistance. In these cases it may be necessary the cut produces 5′-overhangs, 3′-overhangs, or to release several successive gene drives, each blunt ends (Kuhar et al., 2014). Because Cas9 targeting multiple sites, to overcome resistant nickases can generate either overhang type while alleles as they emerge. From an evolutionary Cas9 nucleases produce blunt ends, the enzyme perspective, the ability to preclude resistance can be adapted to the needs of the cell type and by targeting multiple sites is the single greatest organism. advantage of RNA-guided gene drives. The ability to regulate gene expression with We propose to extend this strategy by prefer- Cas9 might be used to temporarily increase the entially targeting multiple sites within the 3′ ends rate of homologous recombination while the of genes important for fitness such that any repair drive is active (Figure 4). For instance, the Cas9 event that deletes all of the target sites creates nuclease involved in cutting might simultaneously a deleterious allele that cannot compete with repress (Gilbert et al., 2013) genes involved in the spread of the drive (Figure 4). Whenever the NHEJ and therefore increase the frequency of HR drive is copied, the cut gene is replaced with (Bozas et al., 2009) if supplied with a shortened a recoded version flanked by the other compo- guide RNA that directs it to bind and block nents of the drive. Recent work has demon- transcription but not cut (Bikard et al., 2013; strated that most genes can be substantially Sternberg et al., 2014). Alternatively, an orthog- recoded with little effect on organism fitness onal nuclease-null Cas9 protein (Esvelt et al., (Lajoie et al., 2013); the 3′ untranslated region 2013) encoded within the drive cassette could might be replaced with an equivalent sequence repress NHEJ genes and activate HR genes from a related gene. Because there would be before activating the Cas9 nuclease. Lastly, Cas9 no homology between the recoded cut site and might be used directly recruit key HR-directing the drive components, the drive cassette would proteins to the cut sites, potentially biasing repair always be copied as a unit. towards that pathway. Relative to drive-resistant alleles, inhibitors of Cas9 are less likely to arise given the historical Evolutionary stability absence of RNA-guided nucleases from eukary- otes. Any inhibitors that do evolve would presum- Even a perfectly efficient endonuclease gene ably target a particular Cas9 protein or guide RNA drive is vulnerable to the evolution of drive resist- used in an earlier drive and could be evaded by ance in the population. Whenever a cut is repaired building future drives using a Cas9 ortholog with a using the NHEJ pathway, the result is typically a different guide RNA (Esvelt et al., 2013; Fonfara drive-resistant allele with insertions or dele- et al., 2013). Alternatively and least likely, organ- tions in the target sequence that prevent it isms might evolve higher RNase activity to prefer- from being cut by the endonuclease. Natural entially degrade all guide RNAs; this may be difficult sequence polymorphisms in the population could to accomplish without harming overall fitness. also prevent cutting. These alleles will typically A final evolutionary concern relates to the increase in abundance and eventually eliminate stability of the gene drive cassette itself. The zinc- the drive because most drives – like most finger nuclease and TALEN-based gene drives in transgenes – are likely to slightly reduce the fit- fruit flies suffered from recombination between ness of the organism. A second path to gene repetitive sequences: only 75% and 40% of each drive resistance would involve the target organism respective drive was sufficiently intact after one evolving a method of specifically inhibiting the copying event to catalyze a second round of drive endonuclease. copying. Because RNA-guided gene drives will The best defense against previously existing not include such highly repetitive elements, they or recently evolved drive-resistant alleles is to are likely to be more stable (Figure 4). target multiple sites. Because mutations in tar- get sites are evolutionarily favored only when Development time they survive confrontation with the gene drive, using many target sites can render it statisti- RNA-guided genome editing is advancing at a cally improbable for any one allele to survive historically unprecedented pace. Because it is

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 8 of 21 196 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

now much easier to make transgenic organisms and therefore candidate gene drives, the design- build-test cycle for gene drives will often be Box 1. Could gene limited only by the generation time of the organism drives alter human in the laboratory. Moreover, many advances from genome engineering can be directly applied to populations? RNA-guided gene drives. For example, all of the methods of increasing Cas9 specificity described Not unless we wait for many centuries. above were developed for RNA-guided genome Even in a hypothetical future in which editing in the past 2 years. Future methods of human germline editing was considered increasing the rate of HR relative to NHEJ would safe and ethical, a driven alteration would be useful for both technologies. These factors be only four times as abundant as a suggest that scientists will enjoy an increasing non-driven alteration a full century after the number of tools well-suited to rapidly building birth of an edited human. This assumes and testing gene drives in addition to those we future generations would not elect to describe above. remove the drive. Whole-genome se- None of this is to gloss over the many practical quencing - a technology that is available in difficulties that are likely to arise when construct- many modern hospitals and is widely ing a particular gene drive in a given species. expected to be ubiquitous in the near Early success is as unlikely as ever when engi- future - is quite capable of detecting the neering complex biological systems. But if half a presence of any gene drive if we decide to dozen or even a dozen design-build-test cycles look. are sufficient to produce moderately efficient gene drives, many molecular biology laboratories around the world will soon be capable of engi- neering wild populations. category includes all viruses and bacteria as well as most unicellular organisms. Highly effi- Gene drive limitations cient standard drives might be able to slowly spread through populations that employ a mix Given their potentially widespread availability, it of sexual and asexual reproduction, such as cer- will be essential to develop a comprehensive tain plants, but drives intended to suppress the understanding of the fundamental limitations of population would presumably force target genetic drive systems. organisms to reproduce asexually in order to First and most important, gene drives require avoid suppression. many generations to spread through popula- Third, drive-mediated genome alterations are tions. Once transgenic organisms bearing the not permanent on an evolutionary timescale. gene drive are constructed in the laboratory, While gene drives can spread traits through pop- they must be released into the wild to mate with ulations even if they are costly to each individual wild-type individuals in order to begin the pro- organism, harmful traits will eventually be out- cess of spreading the drive through the wild competed by more fit alleles after the drive has population. The total time required to spread gone to fixation. Highly deleterious traits may be to all members depends on the number of drive- eliminated even more quickly, with non-functional carrying individuals that are released, the gener- versions appearing in large numbers even before ation time of the organism, the efficiency of the drive and its cargo can spread to all members homing, the impact of the drive on individual of the population. Even when the trait is perfectly fitness, and the dynamics of mating and gene linked to the drive mechanism, the selection pres- flow in the population, but in general it will take sure favoring the continued function of Cas9 and several dozen generations (Burt, 2003; Huang the guide RNAs will relax once the drive reaches et al., 2007; Deredec et al., 2008; Marshall, fixation. Maintaining deleterious traits within a 2009; Yahara et al., 2009; Deredec et al., population indefinitely is likely to require sched- 2011). Thus, drives will spread very quickly in uled releases of new RNA-guided gene drives to fast-reproducing species but only slowly in long- periodically overwrite the broken versions in the lived organisms. environment. Second, gene drives cannot affect species Fourth, our current knowledge of the risk that exclusively practice asexual reproduction management (Scott et al., 2002; Touré et al., through clonal division or self-fertilization. This 2004; UNEP, 2010; McGraw and O’Neill, 2013;

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 9 of 21 197 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Alphey, 2014) and containment (Benedict et ability of RNA-guided gene drives to target any al., 2008; Marshall, 2009) issues associated gene may allow them to control the effects of with gene drives is largely due to the efforts of other gene drives or transgenes. researchers focused on mosquito-borne illnesses. Frameworks for evaluating ecological conse- Reversibility quences are similarly focused on mosquitoes RNA-guided gene drives could reverse genome (David et al., 2013) and the few other organisms alterations that have already spread through for which alternative genetic biocontrol methods populations. Suppose a given gene drive causes have been considered (Dana et al., 2014). While unexpected side-effects or is released without these examples provide an invaluable starting public consent. A ‘reversal’ drive released later point for investigations of RNA-guided gene could overwrite one or all genomic changes drives targeting other organisms, studies examin- spread by the first drive (Figure 5A). The new ing the particular drive, population, and associ- sequence spread by the reversal drive must also ated ecosystem in question will be needed. be recoded relative to the original to keep the Safeguards and control strategies first drive from cutting it, but any amino acid changes introduced by the first drive could be Given the potential for gene drives to alter undone. If necessary, a third drive could restore entire wild populations and therefore ecosys- the exact wild-type sequence, leaving only the tems, the development of this technology must guide RNAs and the gene encoding Cas9 as include robust safeguards and methods of con- signatures of past editing (Figure 5B). trol (Oye et al., 2014). Whereas existing gene The ability to update or reverse genomic altera- drive proposals focus on adding genes (Ito et al., tions at the speed of a drive, not just a drive-resist- 2002), disrupting existing genes (Burt, 2003), ant allele, represents an extremely important safety or suppressing populations, RNA-guided gene feature. Reversal drives could also remove conven- drives will also be capable of replacing existing tionally inserted transgenes that entered wild popu- sequences with altered versions that have been lations by cross-breeding or natural mutations that recoded to remove the sites targeted by the spread in response to human-induced selective drive (Figure 3). We hypothesize that the unique pressures. However, it is important to note that even if a reversal drive were to reach all members of the population, any ecological changes caused in the interim would not necessarily be reversed. Box 2. Could gene drives Immunization alter domesticated RNA-guided gene drives could be used to block animals or crops? the spread of other gene drives. For example, an ‘immunizing’ drive could prevent a specific In theory they could, though probably unwanted drive from being copied by recoding not without the permission of the sequences targeted by the unwanted drive farmers, scientists and breeders who (Figure 5A). This could be done preemptively typically monitor reproduction. For or reactively and would spread on a timescale example, genetic records and artificial comparable to that of the unwanted drive. A insemination are so common among combined ‘immunizing reversal’ drive might cattle and other domesticated animals spread through both wild-type individuals and that it would be exceedingly difficult those affected by an earlier gene drive, con- for a gene drive to spread through any verting both types to a recoded version that of these species. Seed farms play a could not be invaded by the unwanted drive similar role for crops. Long generation (Figure 5B). This may represent the fastest method times will represent a further barrier for of neutralizing an already-released drive. As many domesticated species. In general, with a standard reversal drive, any ecological our expectation is that beneficial changes caused in the interim would not neces- applications are more likely to involve sarily be reversed. the alteration of weeds and insect pest populations rather than the crops Precisely targeting subpopulations themselves. RNA-guided gene drives might be confined to a single genetically distinct target species

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 10 of 21 198 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Figure 5. Methods of reversing, preventing, and controlling the spread and effects of gene drives. (A) Reversal drives could correct or reverse genomic alterations made by an earlier drive with unexpected side effects. They might also be used to reverse conventionally engineered or evolved changes. Immunization drives could prevent other gene drives from affecting a specific population or provide a population with resistance to DNA viruses. Precision drives could exclusively spread through a subpopulation with a unique gene or sequence. (B) Together, these can quickly halt an unwanted drive and eventually restore the sequence to the original wild-type save for the residual Cas9 and guide RNAs. (C) Any population with limited gene flow can be given a unique sequence by releasing drives A and B in quick succession. So long as drive A does not escape into other populations before it is completely replaced by drive B, subsequent precision drives can target population B without risking spread into other populations. DOI: 10.7554/eLife.03401.009

or even a subpopulation by targeting unique genes the equivalent non-target sequence (Fu et al., or sequence polymorphisms. Because these ‘pre- 2013; Hsu et al., 2013; Mali et al., 2013a; cision drives’ will only cut the unique sequence, Pattanayak et al., 2013). they will not be able to spread through non-tar- Populations that are not genetically distinct get populations as long as that sequence is suffi- but experience only intermittent gene flow, such ciently distinct. We estimate that either the PAM as those on islands, might be given a unique or at least five base pairs of the spacer must differ sequence permitting them to be specifically within each target site in order to prevent the targeted by precision drives later on. For exam- guide RNAs in the drive from evolving to recognize ple, releasing Drive A into the island population

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 11 of 21 199 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

would recode a target gene, but exhibit no their effectiveness. Periodically releasing organ- other effect (Figure 5A). Releasing Drive B, a isms carrying new unstable drives that are capable precision drive which would exclusively spread of replacing earlier broken versions could extend through Drive A but not the wild-type allele, the suppression effect. would similarly replace Drive A with a unique Genetic approaches to population control sequence. So long as Drive A does not escape might initiate suppression only when two distinct the island before being replaced, the unique ‘interacting drives’ encounter one another sequence in the island population would allow it (Figure 6). For example, a cross between standard to be targeted with future precision drives that drives A and B might produce sterile females and could not spread through mainland populations fertile males that pass on the ‘sterile-daughter’ (Figure 5C). trait when crossed with females of any type. Scattering A- and B-carrying individuals through- Limiting population suppression out an existing population would produce many tiny pockets dominated by either A or B and very Population suppression may be one of the most few organisms in between due to the infertility of powerful applications of gene drives. The previ- AB females. Because each drive would spread ously described genetic load and sex-biasing from a small number of initially released individu- drives (Burt, 2003) could potentially lead to als scattered over a wide area, this strategy may extinction (Deredec et al., 2008, 2011). While be capable of large-scale population suppression, this outcome may be necessary to achieve com- but its effectiveness and resolution will depend pelling goals such as the eradication of malaria, on the average distance between released A and other situations may call for more refined meth- B individuals. Further suppressing the residual A ods. Here we outline a handful of alternative and B populations could be accomplished by architectures that would provide greater control releasing only members of the opposite drive type. over the extent of population suppression. Modeling studies will be needed to determine Chemical approaches to population control whether this possibility is feasible for different spe- might utilize ‘sensitizing drives’ to render target cies. Interestingly, the use of this drive type would organisms vulnerable to a particular molecule effectively induce speciation in the affected using one of three strategies (Figure 6). First, a population. sensitizing drive might reverse known mutations Finally, immunizing drives might protect spe- that confer resistance to existing or cific subpopulations from the effects of full-scale herbicides. Second, it might carry a prodrug- suppression drives released elsewhere (Figure 6). converting enzyme (Schellmann et al., 2010) that Assuming some degree of gene flow, the immu- would render a prodrug molecule toxic to organisms nized population will eventually replace the sup- that express it. Third, it could swap a conserved pressed population, though this might be delayed gene for a version that is strongly inhibited by if crossing the two drives generates a sterile- a particular small molecule. Because sensitizing daughter effect as described above. Due to the drives would have no effect in the absence of comparatively uncontrolled spread of both drive their associated molecule – and in some cases types through the wild-type population, this vice versa – they could grant very fine control over method would only be suited to large geographic the geography and extent of population suppres- areas or subpopulations with limited gene flow. sion with minimal ecological risk. For example, immunization might be used to pro- Temporal approaches to controlling populations tect the native population of a species while sup- would deliberately limit the lifetime of a suppres- pressing or eradicating populations on other sion drive by rendering its effects evolutionarily continents. unstable (Figure 6). For example, a male- determining or female-specific sterility gene car- Applications of RNA-guided gene ried by a standard drive on an autosome would drives suppress the target population, but the effect would be short-lived because any drive that RNA-guided gene drives have the potential to acquired a loss-of-function mutation in the cargo merge the fields of genomic and ecological engi- gene would be strongly favored by natural selec- neering. They may enable us to address numerous tion due to its ability to produce fertile female problems in global health, agriculture, sustaina- offspring. Notably, turning existing female-specific bility, ecological science, and many other areas sterility lines (Fu et al., 2010; Labbe et al., 2012; (Figure 7). Of these opportunities, perhaps the Alphey, 2014) into unstable drives may increase most compelling involve curtailing the spread

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 12 of 21 200 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Figure 6. Controlling population suppression. Previously proposed genetic load and meiotic suppression drives spread without limit and may incur a substantial risk of extinction. Alternative gene drive types might be used to grant finer control over the extent of suppression. ‘Sensitization drives’ would be harmless save for conferring vulnerability to a particular chemical, which could then be used as a population-specific pesticide. Evolutionarily ‘unstable drives’ would place a limit on the average number of drive copying events and thus the extent of population suppression. ‘Interacting drives’ would initiate suppression only upon encountering a specific genetic signature in the population, in this case a different gene drive. The combination would create a sterile-daughter effect (Figure 6—figure supplement 1) capable of continuing suppression for several generations. Finally, an immunizing drive could protect a subpopulation from a full genetic load or male- biasing suppression drive employed elsewhere. Interacting drive and immunizing drive approaches would be effective on very large populations spread across substantial geographic areas (Figure 6—figure supplement 2) while suffering from correspondingly reduced geographic resolution and greater ecological risk (Figure 6—figure supplement 3). Resolutions are approximations only and will vary with the specific drive utilized in each class. DOI: 10.7554/eLife.03401.010 The following figure supplements are available for figure 6: Figure supplement 1. Sample interacting drives that produce a sterile-daughter effect. DOI: 10.7554/eLife.03401.011

Figure 6. Continued on next page

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 13 of 21 201 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Figure 6. Continued Figure supplement 2. An extreme example of ecological management: the use of suppression and immunizing drives to control rat populations worldwide. DOI: 10.7554/eLife.03401.012 Figure supplement 3. Characteristics of population suppression drives. DOI: 10.7554/eLife.03401.013

species to block transmission. Several laborato- ries have identified candidate gene disruptions or Box 3. What types of transgenes that interfere with the transmission of genes can be edited malaria (Ito et al., 2002; Dong et al., 2011; Isaacs et al., 2012) and other well-studied dis- using gene drives? eases (Franz et al., 2006). Depending on their effectiveness, these alterations may or may not Genes can be edited reliably if they are allow the disease to be eradicated before the important to fitness. This is because NHEJ pathogen evolves resistance. Alternatively, the events that create drive-resistant alleles relevant vector species might be suppressed or by deleting all the protospacer cut sites eliminated using RNA-guided gene drives, then will only be harmful if they disrupt the potentially reintroduced from sheltered laboratory function of important genes. NHEJ events or island populations once disease eradication is in unimportant genes and sequences will complete. In the case of malaria, gene drive strat- produce drive-resistant alleles lacking the egies may represent particularly effective solutions targeted sites. These will spread and to the emerging problem of mosquito vectors interfere with propagation of the drive. As with an evolved preference to bite and rest out- a result, unimportant genes can be doors, traits that render them resistant to current reliably disrupted or deleted but not control strategies based on indoor insecticide edited. spraying and bednets. Genes that are carried as cargo will not be evolutionarily stable unless they directly Agricultural safety and sustainability contribute to the efficient function of the drive. This limits opportunities to spread The evolution of resistance to pesticides and transgenes unrelated to drive function, herbicides is a major problem for agriculture. It although periodically releasing new has been assumed that resistant populations will drives that overwrite earlier broken remain resistant unless the relevant alleles impose versions could potentially maintain a substantial fitness cost in the absence of pesti- functional cargo genes in a large fraction cide or herbicide. We propose that RNA-guided of the population. sensitizing drives might replace resistant alleles with their ancestral equivalents to restore vulnera- bility. For example, sensitizing drives could poten- tially reverse the mutations allowing the western of vector-borne infectious diseases, controlling corn rootworm to resist Bt toxins (Gassmann agricultural pests, and reducing populations of et al., 2014) or horseweed and pigweed to resist environmentally and economically destructive the herbicide glyphosate (Gaines et al., 2010; invasive species. Ge, 2010), which is currently essential to more sustainable no-till agriculture. Because these Eradicating insect-borne diseases three organisms undergo one generation per The human toll inflicted by infectious diseases year, comparatively large numbers of drive-bear- spread by insects is staggering. Malaria alone kills ing individuals must be released to quickly exert over 650,000 people each year, most of them an effect, but fewer than are already released to children, while afflicting 200 million more with control pests using the sterile-insect technique debilitating (WHO, 2013). Dengue, yellow (Gould and Schliekelman, 2004; Dyck et al., fever, chikungunya, trypanosomiasis, leishmaniasis, 2005). Releases would need to occur in local Chagas, and Lyme disease are also spread by areas not treated with pesticide or herbicide, insects. These afflictions could potentially be which would quickly become reservoirs of sensi- controlled or even eradicated by altering vector tizing drives that could spread into adjacent

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 14 of 21 202 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

fields. Periodically releasing new drives could po- tentially allow any given pesticide or herbicide to be utilized indefinitely. Modeling experiments will be needed to evaluate feasibility for different target species. A second form of sensitizing drive could potentially render pest populations vulnerable to molecules that never previously affected them. For example, a gene important to fitness might be replaced with a version from another species or laboratory isolate whose function is sensitive to a particular compound. In principle, this approach could eventually lead to the development and use of safer and more species-specific pesticides and herbicides.

Controlling invasive species One of the most environmentally damaging consequences of global economic activity is the introduction of invasive species, which often cause ecological disruption or even the extinc- tion of native species. Isolated ecosystems such as those on small islands are especially vulner- able. RNA-guided suppression drives might be used to promote biodiversity by controlling or even eradicating invasive species from islands Figure 7. Potential applications of RNA-guided gene or possibly entire continents. The economics of drives. Clockwise from left. Disease vectors such as malarial mosquitoes might be engineered to resist invasive species control are also compelling: pathogen acquisition or eliminated with a suppression the top ten invasives in the United States cause drive. Wild populations that serve as reservoirs for an estimated $42 billion in damages every year human viruses could be immunized using Cas9, RNAi (Pimentel et al., 2005). Black and brown rats machinery, or elite controller antibodies carried by a alone cause $19 billion in damages and may be gene drive. Reversal and immunization drives could responsible for more extinctions than any other help ensure that all transgenes are safe and controlled. nonhuman species. Drives might quickly spread protective genes through Deploying RNA-guided suppression drives threatened or soon-to-be-threatened species such as against invasive species will incur two primary amphibians facing the expansion of chytrid fungus risks related to undesired spread. First, rare mating (Rosenblum et al., 2010). Invasive species might be events may allow the drive to affect closely locally controlled or eradicated without directly affecting others. Sensitizing drives could improve the related species. Using precision drives to target sustainability and safety of pesticides and herbicides. sequences unique to the invasive species could Gene drives could test ecological hypotheses mitigate or eliminate this problem. Second, the concerning gene flow, sex ratios, speciation, and suppression drive might spread from the invasive evolution. Technical requirements for these population back into the native habitat, perhaps applications vary with the drive type required even through intentional human action. (Figure 7—figure supplement 1). Native populations might be protected using DOI: 10.7554/eLife.03401.014 an immunizing drive, but doing so would risk The following figure supplement is available for transferring immunity back into invasive popu- figure 7: lations. Instead, we might grant the invasive Figure supplement 1. Technical limitations of different population a unique sequence with a standard gene drive architectures with implications for various drive (Figure 5C), verify that these changes have applications. DOI: 10.7554/eLife.03401.015 not spread to the native population, and only then release a suppression drive targeting the recoded sequences while holding an immunizing drive in reserve. Another approach might utilize a sensitizing drive to render all populations newly

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 15 of 21 203 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

vulnerable to a specific compound, which could then be used as a pesticide for the local control of invasive populations. All of these possibili- ties will require modeling and experimentation to establish safety and feasibility before use. Most importantly, all decisions involving the use of suppression drives must involve extensive delib- erations including but not limited to ecologists and citizens of potentially affected communities.

Development and release precautions Because any consequences of releasing RNA- guided gene drives into the environment would be shared by the local if not global community, research involving gene drives capable of spreading through wild-type populations should occur only after a careful and fully transparent Figure 8. Containment strategies and ecological risk. review process. However, basic research into gene Ecological containment involves building and testing gene drives in geographic areas that do not harbor drives and methods of controlling their effects can native populations of the target species. For example, proceed without risking this type of spread so most gene drive studies involving tropical malarial long as appropriate ecological or molecular con- mosquitoes have been conducted in temperate regions tainment strategies are employed (Figure 8). in which the mosquitoes cannot survive or find mates. A great deal of information on probable eco- Molecular containment ensures that the basic require- logical outcomes can be obtained without testing ments for drive are not met when mated with wild-type or even building replication-competent gene organisms. True drives must cut the homologous drives. For example, early studies might examine wild-type sequence and copy both the gene encoding possible ecological effects by performing con- Cas9 and the guide RNAs. Experiments that cut tained field trials with organisms that have been transgenic sequences absent from wild populations and engineered to contain the desired change but copy either the gene encoding Cas9 or the guide RNAs - but not both - should be quite safe. Ecological or do not possess a functional drive to spread it. molecular containment should allow basic research into Because they do involve transgenic organisms, gene drive effectiveness and optimization to be these experiments are not completely without pursued with negligible risk. Figure 8—figure risk, but such transgenes are unlikely to spread in supplement 1 categorizes these and many other the absence of a drive. possible experiments according to estimated risk. We recommend that all laboratories seeking DOI: 10.7554/eLife.03401.016 to build standard gene drives capable of spreading The following figure supplement is available for figure 8: through wild populations simultaneously create Figure supplement 1. Estimated ecological risk of reversal drives able to restore the original pheno- experiments during RNA-guided gene drive type. Similarly, suppression drives should be development. constructed in tandem with a corresponding DOI: 10.7554/eLife.03401.017 immunizing drive. These precautions would allow the effects of an accidental release to be swiftly if partially counteracted. The prevalence of gene societal review and consent. As self-propagating drives in the environment could in principle be alterations of wild populations, RNA-guided gene monitored by targeted amplification or metagen- drives will be capable of influencing entire eco- omic sequencing of environmental samples. systems for good or for ill. As such, it is impera- Further investigation of possible monitoring tive that all research in this nascent field operate strategies will be needed. under conditions of full transparency, including independent scientific assessments of probable impacts and thoughtful, informed, and fully inclu- Transparency, public discussion, and evaluation sive public discussions. The decision of whether or not to utilize a Technologies with the potential to significantly gene drive for a given purpose should be based influence the lives of the general public demand entirely on the probable benefits and risks of

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 16 of 21 204 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

that specific drive. That is, each drive should be ecosystems (Drinkwater et al., 2014). We are judged solely by its potential outcomes, such as grateful to S Lightfoot, K Oye, JP Collins, and M its ability to accomplish the intended aims, its Wilkinson for critical reading of the manuscript, probable effects on other species, the risk of and to F Gould, T Wu, J Braff, and members of the spreading into closely related species by rare Church and Oye laboratories for helpful discus- mating events, and impacts on ecosystems and sions. KME was supported by a Wyss Technology human societies. As scientists developing these Development Fellowship. technologies, it will be our responsibility to make all empirical data and predictive models freely Funding available to the public in a transparent and under- Funder Grant Author standable format. Above all else, we must openly reference share our level of confidence in these assess- number ments as we determine how best to proceed. Wyss Institute Technology Kevin M Esvelt for Biologically Development Inspired Engineering Fellowship Discussion Wyss Institute Synthetic Kevin M Esvelt, The potentially widespread implications of for Biologically Biology Andrea L Smidler Inspired Engineering Platform RNA-guided gene drives demand a thoughtful funds and collected response. Numerous practical dif- The funders had no role in study design, data collection ficulties must be overcome before gene drives and interpretation, or the decision to submit the work will be in a position to address any of the sug- for publication. gested applications. Many of our proposals and predictions are likely to fall short simply because biological systems are complex and difficult to Author contributions engineer. Even so, the current rate of scientific KME, Conceived of the study, performed the analysis, advancement related to Cas9 and the many designed new drive types, drafted and revised the article; outcomes accessible using the simplest of gene ALS, Designed new drive types, drafted and revised the drives suggest that molecular biologists will article; FC, GMC, Revised the article soon be able to edit the genomes of wild popu- lations, reverse or update those changes in Kevin M Esvelt Synthetic Biology Platform, Wyss response to field observations, and perhaps even Institute for Biologically Inspired Engineering, Harvard engage in targeted population suppression. Medical School, Boston, United States What criteria might we use to evaluate an http://orcid.org/0000-0001-8797-3945 RNA-guided gene drive intended to address Andrea L Smidler Synthetic Biology Platform, Wyss a given problem? There are compelling argu- Institute for Biologically Inspired Engineering, Harvard ments in favor of eliminating insect-borne human Medical School, Boston, United States; Department of diseases, developing and supporting more sus- Immunology and Infectious Diseases, Harvard School tainable agricultural models, and controlling of Public Health, Boston, United States environmentally damaging invasive species. At http://orcid.org/0000-0002-3281-1161 the same time, there are valid concerns regarding Flaminia Catteruccia Department of Immunology and our ability to accurately predict the ecological Infectious Diseases, Harvard School of Public Health, and human consequences of these interven- Boston, United States; Dipartimento di Medicina tions. By bringing these possibilities before the Sperimentale e Scienze Biochimiche, Università degli scientific community and the public prior to their Studi di Perugia, Terni, Italy realization in the laboratory, we hope to initiate George M Church Synthetic Biology Platform, Wyss transparent, inclusive, and well-informed dis- Institute for Biologically Inspired Engineering, Harvard cussions concerning the responsible evaluation Medical School, Boston, United States and application of these nascent technologies Competing interests: KME: Filed for a patent (Oye et al., 2014). concerning RNA-guided gene drives. ALS: Filed for a patent concerning RNA-guided gene drives. GMC: Acknowledgements Filed for a patent concerning RNA-guided gene We thank the participants of the MIT/Woodrow drives. The other author declares that no competing Wilson Center workshop ‘Creating a Research interests exist. Agenda for the Ecological Implications of Synthetic Received 18 May 2014 Biology’ for sharing their thoughts concerning the Accepted 09 July 2014 potential impacts of this technology on various Published 17 July 2014

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 17 of 21 205 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

References in Drosophila. Science 316:597–600. doi: 10.1126/ Alphey N, Bonsall MB. 2014. Interplay of population science.1138595. genetics and dynamics in the genetic control Cho SW, Kim S, Kim JM, Kim J-S. 2013. Targeted of mosquitoes. Journal of the Royal Society: Interface genome engineering in human cells with the Cas9 11:20131071. doi: 10.1098/rsif.2013.1071. RNA-guided endonuclease. Nature Biotechnology Alphey L. 2014. Genetic control of mosquitoes. 31:230–232. doi: 10.1038/nbt.2507. Annual Review of Entomology 59:205–224. doi: Christian M, Cermak T, Doyle EL, Schmidt C, Zhang F, 10.1146/annurev-ento-011613-162002. Hummel A, Bogdanove AJ, Voytas DF. 2010. Targeting Bassett AR, Tibbit C, Ponting CP, Liu J-L. 2013. Highly DNA double-strand breaks with TAL effector efficient targeted mutagenesis of Drosophila with nucleases. Genetics 186:757–761. doi: 10.1534/ the CRISPR/Cas9 system. Cell Reports 4:220–228. genetics.110.120717. doi: 10.1016/j.celrep.2013.06.020. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Beeman RW, Friesen KS, Denell RE. 1992. Maternal- Hsu PD, Wu X, Jiang W, Marraffini LA, et al. 2013. effect selfish genes in flour beetles. Science 256:89–92. Multiplex genome engineering using CRISPR/Cas doi: 10.1126/science.1566060. systems. Science 339:819–823. doi: 10.1126/ Benedict M, D‘Abbs P, Dobson S, Gottlieb M, science.1231143. Harrington L, Higgs S, James A, James S, Knols B, Craig GB, Hickey WA, Vandehey RC. 1960. An inherited Lavery J, et al. 2008. Guidance for contained field trials male-producing factor in Aedes aegypti. Science of vector mosquitoes engineered to contain a gene 132:1887–1889. doi: 10.1126/science.132.3443.1887. drive system: recommendations of a scientific working Curtis CF. 1968. Possible use of translocations to fix group. Vector Borne and Zoonotic Diseases 8:127–166. desirable genes in insect pest populations. Nature doi: 10.1089/vbz.2007.0273. 218:368–369. doi: 10.1038/218368a0. Bikard D, Jiang W, Samai P, Hochschild A, Zhang F, Dana GV, Cooper AM, Pennington KM, Sharpe LS. Marraffini LA. 2013. Programmable repression and 2014. Methodologies and special considerations for activation of bacterial gene expression using an environmental risk analysis of genetically modified engineered CRISPR-Cas system. Nucleic Acids aquatic biocontrol organisms. Biological Invasions Research 41:7429–7437. doi: 10.1093/nar/gkt520. 16:1257–1272. doi: 10.1007/s10530-012-0391-x. Bozas A, Beumer KJ, Trautman JK, Carroll D. 2009. David AS, Kaser JM, Morey AC, Roth AM, Andow DA. Genetic analysis of Zinc-Finger nuclease-induced gene 2013. Release of genetically engineered insects: a targeting in Drosophila. Genetics 182:641–651. framework to identify potential ecological effects. doi: 10.1534/genetics.109.101329. Ecology and Evolution 3:4000–4015. doi: 10.1002/ Burt A, Koufopanou V. 2004. Homing endonuclease ece3.737. genes: the rise and fall and rise again of a selfish Deltcheva E, Chylinski K, Sharma CM, Gonzales K, element. Current Opinion in Genetics & Development Chao Y, Pirzada ZA, Eckert MR, Vogel J, Charpentier E. 14:609–615. doi: 10.1016/j.gde.2004.09.010. 2011. CRISPR RNA maturation by trans-encoded small Burt A, Trivers R. 2009. Genes in conflict: the biology RNA and host factor RNase III. Nature 471:602–607. of selfish genetic elements. Harvard University Press. doi: 10.1038/nature09886. Burt A. 2003. Site-specific selfish genes as tools Deredec A, Burt A, Godfray HCJ. 2008. The for the control and genetic engineering of natural population genetics of using homing endonuclease populations. Proceedings of the Biological Sciences B genes in vector and pest management. Genetics 270:921–928. doi: 10.1098/rspb.2002.2319. 179:2013–2026. doi: 10.1534/genetics.108.089037. Burt A. 2014. Heritable strategies for controlling insect Deredec A, Godfray HCJ, Burt A. 2011. Requirements vectors of disease. Philosophical Transactions of the for effective malaria control with homing endonuclease Royal Society B 369:20130432. doi: 10.1098/ genes. Proceedings of the National Academy of rstb.2013.0432. Sciences USA 108:E874–E880. doi: 10.1073/pnas. Chan Y-S, Naujoks DA, Huen DS, Russell S. 2011. 1110717108. Insect population control by homing endonuclease- DiCarlo JE, Norville JE, Mali P, Rios X, Aach J, based gene drive: an evaluation in Drosophila Church GM. 2013. Genome engineering in melanogaster. Genetics 188:33–44. doi: 10.1534/ Saccharomyces cerevisiae using CRISPR-Cas systems. genetics.111.127506. Nucleic Acids Research 41:4336–4343. doi: 10.1093/ Chan Y-S, Takeuchi R, Jarjour J, Huen DS, Stoddard BL, nar/gkt135. Russell S. 2013a. The design and in vivo evaluation of Dickinson DJ, Ward JD, Reiner DJ, Goldstein B. 2013. engineered I-OnuI-based enzymes for HEG gene Engineering the Caenorhabditis elegans genome using drive. PLOS ONE 8:e74254. doi: 10.1371/journal. Cas9-triggered homologous recombination. Nature pone.0074254. Methods 10:1028–1034. doi: 10.1038/nmeth.2641. Chan Y-S, Huen DS, Glauert R, Whiteway E, Russell S. Dong Y, Das S, Cirimotich C, Souza-Neto JA, McLean 2013b. Optimising homing endonuclease gene drive KJ, Dimopoulos G. 2011. Engineered anopheles performance in a semi-refractory species: the immunity to Plasmodium infection. PLOS Pathogens Drosophila melanogaster experience. PLOS ONE 7:e1002458. doi: 10.1371/journal.ppat.1002458. 8:e54130. doi: 10.1371/journal.pone.0054130. Drinkwater K, Kuiken T, Lightfoot S, McNamara J, Charlesworth B, Langley CH. 1989. The population Oye K. 2014. Creating a research agenda for genetics of Drosophila transposable elements. Annual the ecological implications of synthetic biology. Review of Genetics 23:251–287. doi: 10.1146/annurev. At http://www.synbioproject.org/library/publications/ ge.23.120189.001343. archive/6685/. Chen C-H, Huang H, Ward CM, Su JT, Schaeffer LV, Dyck VA, Hendrichs J, Robinson AS. 2005. Sterile Guo M, Hay BA. 2007. A synthetic maternal-effect insect technique: principles and practice in area-wide selfish genetic element drives population replacement integrated pest management. Springer.

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 18 of 21 206 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Esvelt KM, Carlson JC, Liu DR. 2011. A system for the Ge X, d’Avignon DA, Ackerman JJ, Sammons RD. continuous directed evolution of biomolecules. Nature 2010. Rapid vacuolar sequestration: the horseweed 472:499–503. doi: 10.1038/nature09929. glyphosate resistance mechanism. Pest Management Esvelt KM, Mali P, Braff JL, Moosburner M, Yaung SJ, Science 66:345–348. doi: 10.1002/ps.1911. Church GM. 2013. Orthogonal Cas9 proteins for Gilbert LA, Larson MH, Morsut L, Liu Z, Brar GA, RNA-guided gene regulation and editing. Nature Torres SE, Stern-Ginossar N, Brandman O, Whitehead Methods 10:1116–1121. doi: 10.1038/nmeth.2681. EH, Doudna JA, et al. 2013. CRISPR-mediated modular Esvelt KM, Wang HH. 2013. Genome-scale engineering RNA-guided regulation of transcription in eukaryotes. for systems and synthetic biology. Molecular Systems Cell 154:442–451. doi: 10.1016/j.cell.2013.06.044. Biology 9:641. doi: 10.1038/msb.2012.66. Gould F, Schliekelman P. 2004. Population genetics of Fiorenza MT, Bevilacqua A, Bevilacqua S, Mangia F. autocidal control and strain replacement. Annual 2001. Growing dictyate oocytes, but not early Review of Entomology 49:193–217. doi: 10.1146/ preimplantation embryos, of the mouse display high annurev.ento.49.061802.123344. levels of DNA homologous recombination by single- Gratz SJ, Cummings AM, Nguyen JN, Hamm DC, strand annealing and lack DNA nonhomologous end Donohue LK, Harrison MM, Wildonger J, O‘Connor- joining. Developmental Biology 233:214–224. Giles KM. 2013. Genome engineering of Drosophila doi: 10.1006/dbio.2001.0199. with the CRISPR RNA-guided Cas9 nuclease. Genetics Fonfara I, Le Rhun A, Chylinski K, Makarova KS, 194:1029–1035. doi: 10.1534/genetics.113.152710. Lécrivain AL, Bzdrenga J, Koonin EV, Charpentier E. Guilinger JP, Thompson DB, Liu DR. 2014. Fusion of 2013. Phylogeny of Cas9 determines functional catalytically inactive Cas9 to FokI nuclease improves exchangeability of dual-RNA and Cas9 among the specificity of genome modification. Nature orthologous type II CRISPR-Cas systems. Nucleic Biotechnology 32:577–582. doi: 10.1038/nbt.2909. Acids Research 42:2577–2590. doi: 10.1093/nar/ Hamilton WD. 1967. Extraordinary sex ratios. A gkt1074. sex-ratio theory for sex linkage and inbreeding has new Franz AWE, Sanchez-Vargas I, Adelman ZN, Blair CD, implications in cytogenetics and entomology. Science Beaty BJ, James AA, Olson KE. 2006. Engineering 156:477–488. doi: 10.1126/science.156.3774.477. RNA interference-based resistance to dengue virus Hickey WA, Craig GB. 1966. Genetic distortion of type 2 in genetically modified Aedes aegypti. sex ratio in a mosquito, Aedes aegypti. Genetics Proceedings of the National Academy of Sciences 53:1177–1196. USA 103:4198–4203. doi: 10.1073/pnas.0600479103. Hsu PD, Scott DA, Weinstein JA, Ran FA, Konermann Friedland AE, Tzur YB, Esvelt KM, Colaiácovo MP, S, Agarwala V, Li Y, Fine EJ, Wu X, Shalem O, et al. Church GM, Calarco JA. 2013. Heritable genome 2013. DNA targeting specificity of RNA-guided Cas9 editing in C. elegans via a CRISPR-Cas9 system. Nature nucleases. Nature Biotechnology 31:827–832. Methods 10:741–743. doi: 10.1038/nmeth.2532. doi: 10.1038/nbt.2647. Fu G, Lees RS, Nimmo D, Aw D, Jin L, Gray P, Hu S, Wang Z, Polejaeva I. 2013. Knockout of goat Berendonk TU, White-Cooper H, Scaife S, Kim Phuc H, nucleoporin 155 gene using CRISPR/Cas systems. et al. 2010. Female-specific flightless phenotype for Reproduction, Fertility, and Development 26:134. mosquito control. Proceedings of the National doi: 10.1071/RDv26n1Ab40. Academy of Sciences USA 107:4550–4554. doi: Huang Y, Magori K, Lloyd AL, Gould F. 2007. 10.1073/pnas.1000251107. Introducing desirable transgenes into insect populations Fu Y, Foden JA, Khayter C, Maeder ML, Reyon D, using Y-linked meiotic drive - a theoretical assessment. Joung JK, Sander JD. 2013. High-frequency off-target Evolution; International Journal of Organic Evolution mutagenesis induced by CRISPR-Cas nucleases in 61:717–726. doi: 10.1111/j.1558-5646.2007.00075.x. human cells. Nature Biotechnology 31:822–826. Hwang WY, Fu Y, Reyon D, Maeder ML, Tsai SQ, doi: 10.1038/nbt.2623. Sander JD, Peterson RT, Yeh JR, Joung JK. 2013. Fu Y, Sander JD, Reyon D, Cascio VM, Joung JK. Efficient genome editing in zebrafish using a CRISPR- 2014. Improving CRISPR-Cas nuclease specificity using Cas system. Nature Biotechnology 31:227–229. truncated guide RNAs. Nature Biotechnology doi: 10.1038/nbt.2501. 32:279–284. doi: 10.1038/nbt.2808. Isaacs AT, Jasinskiene N, Tretiakov M, Thiery I, Zettor A, Gaines TA, Zhang W, Wang D, Bukun B, Chisholm ST, Bourgouin C, James AA. 2012. Transgenic Anopheles Shaner DL, Nissen SJ, Patzoldt WL, Tranel PJ, stephensi coexpressing single-chain antibodies resist Culpepper AS, et al. 2010. Gene amplification confers Plasmodium falciparum development. Proceedings of the glyphosate resistance in Amaranthus palmeri. National Academy of Sciences USA 109:E1922–E1930. Proceedings of the National Academy of Sciences doi: 10.1073/pnas.1207738109. USA 107:1029–1034. doi: 10.1073/pnas.0906649107. Ito J, Ghosh A, Moreira LA, Wimmer EA, Jacobs- Galizi R, Doyle LA, Menichelli M, Bernardini F, Lorena M. 2002. Transgenic anopheline mosquitoes Deredec A, Burt A, Stoddard BL, Windbichler N, impaired in transmission of a malaria parasite. Nature Crisanti A. 2014. A synthetic sex ratio distortion 417:452–455. doi: 10.1038/417452a. system for the control of the human malaria mosquito. Jiang W, Bikard D, Cox D, Zhang F, Marraffini LA. Nature Communications 5:3977. doi: 10.1038/ 2013a. RNA-guided editing of bacterial genomes ncomms4977. using CRISPR-Cas systems. Nature Biotechnology Gassmann AJ, Petzold-Maxwell JL, Clifton EH, Dunbar 31:233–239. doi: 10.1038/nbt.2508. MW, Hoffmann AM, Ingber DA, Keweshan RS. 2014. Jiang W, Zhou H, Bi H, Fromm M, Yang B, Weeks DP. Field-evolved resistance by western corn rootworm to 2013b. Demonstration of CRISPR/Cas9/sgRNA- multiple Bacillus thuringiensis toxins in transgenic mediated targeted gene modification in Arabidopsis, maize. Proceedings of the National Academy of Sciences tobacco, sorghum and rice. Nucleic Acids Research USA 111:5141–5146. doi: 10.1073/pnas.1317179111. 41:e188. doi: 10.1093/nar/gkt780.

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 19 of 21 207 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Theoretical Biology 258:250–265. doi: 10.1016/j.jtbi. Charpentier E. 2012. A programmable dual-RNA- 2009.01.031. guided DNA endonuclease in adaptive bacterial McGraw EA, O’Neill SL. 2013. Beyond insecticides: immunity. Science 337:816–821. doi: 10.1126/ new thinking on an ancient problem. Nature Reviews science.1225829. Microbiology 11:181–193. doi: 10.1038/nrmicro2968. Jinek M, East A, Cheng A, Lin S, Ma E, Doudna J. Newton ME, Wood RJ, Southern DI. 1976. A 2013. RNA-programmed genome editing in human cytogenetic analysis of meiotic drive in the cells. eLife 2:e00471. doi: 10.7554/eLife.00471. mosquito, Aedes aegypti (L.). Genetica 46:297–318. Kondo S, Ueda R. 2013. Highly improved gene doi: 10.1007/BF00055473. targeting by germline-specific Cas9 expression in Nishimasu H, Ran FA, Hsu PD, Konermann S, Shehata Drosophila. Genetics 195:715–721. doi: 10.1534/ SI, Dohmae N, Ishitani R, Zhang F, Nureki O. 2014. genetics.113.156737. Crystal structure of Cas9 in complex with guide RNA Kuhar R, Gwiazda KS, Humbert O, Mandt T, and target DNA. Cell 156:935–949. doi: 10.1016/j. Pangallo J, Brault M, Khan I, Maizels N, Rawlings DJ, cell.2014.02.001. Scharenberg AM, et al. 2014. Novel fluorescent Nissim L, Perli SD, Fridkin A, Perez-Pinera P, Lu TK. genome editing reporters for monitoring DNA repair 2014. Multiplexed and programmable regulation of pathway utilization at endonuclease-induced breaks. gene networks with an integrated RNA and CRISPR/ Nucleic Acids Research 42:e4. doi: 10.1093/nar/ Cas toolkit in human cells. Molecular Cell 54:698–710. gkt872. doi: 10.1016/j.molcel.2014.04.022. Labbe GMC, Scaife S, Morgan SA, Curtis ZH, Alphey L. North A, Burt A, Godfray HCJ. 2013. Modelling the 2012. Female-specific flightless (fsRIDL) phenotype for spatial spread of a homing endonuclease gene in a control of Aedes albopictus. PLOS Neglected Tropical mosquito population. Journal of Applied Ecology Diseases 6:e1724. doi: 10.1371/journal.pntd.0001724. 50:1216–1225. Lajoie MJ, Kosuri S, Mosberg JA, Gregg CJ, Zhang D, Oye KA, Esvelt KM, Appleton E, Catteruccia F, Church GM. 2013. Probing the limits of genetic Church GM, Kuiken T, Lightfoot SB-Y, McNamara J, recoding in essential genes. Science 342:361–363. Smidler AL, Collins JP. 2014. Regulating gene drives. doi: 10.1126/science.1241460. Science. doi: 10.1126/science.1254287. Li J, Wang X, Zhang G, Githure JI, Yan G, James AA. Pattanayak V, Lin S, Guilinger JP, Ma E, Doudna JA, 2013a. Genome-block expression-assisted association Liu DR. 2013. High-throughput profiling of off-target studies discover malaria resistance genes in Anopheles DNA cleavage reveals RNA-programmed Cas9 gambiae. Proceedings of the National Academy nuclease specificity. Nature Biotechnology 31:839–843. of Sciences USA 110:20675–20680. doi: 10.1038/nbt.2673. doi: 10.1073/pnas.1321024110. Pimentel D, Zuniga R, Morrison D. 2005. Update on Li D, Qiu Z, Shao Y, Chen Y, Guan Y, Liu M, Li Y, Gao N, the environmental and economic costs associated with Wang L, Lu X, et al. 2013b. Heritable gene targeting in alien-invasive species in the United States. Ecological the mouse and rat using a CRISPR-Cas system. Nature Economics 52:273–288. doi: 10.1016/j.ecolecon. Biotechnology 31:681–683. doi: 10.1038/nbt.2661. 2004.10.002. Li W, Teng F, Li T, Zhou Q. 2013c. Simultaneous Preston CR, Flores CC, Engels WR. 2006. Differential generation and germline transmission of multiple gene usage of alternative pathways of double-strand break mutations in rat using CRISPR-Cas systems. Nature repair in Drosophila. Genetics 172:1055–1068. Biotechnology 31:684–686. doi: 10.1038/nbt.2652. doi: 10.1534/genetics.105.050138. Lyttle TW. 1977. Experimental population genetics of Reeves RG, Denton JA, Santucci F, Bryk J, Reed FA. meiotic drive systems. I. Pseudo-Y chromosomal drive 2012. Scientific standards and the regulation of as a means of eliminating cage populations of genetically modified insects. PLOS Neglected Tropical Drosophila melanogaster. Genetics 86:413–445. Diseases 6:e1502. doi: 10.1371/journal.pntd.0001502. Lyttle TW. 1991. Segregation distorters. Annual Rosenblum EB, Voyles J, Poorten TJ, Stajich JE. 2010. Review of Genetics 25:511–557. doi: 10.1146/annurev. The deadly chytrid fungus: a story of an emerging ge.25.120191.002455. pathogen. PLOS Pathogens 6:e1000550. doi: 10.1371/ Mali P, Aach J, Stranges PB, Esvelt KM, Moosburner journal.ppat.1000550. M, Kosuri S, Yang L, Church GM. 2013a. CAS9 Saleh-Gohari N, Helleday T. 2004. Conservative transcriptional activators for target specificity homologousrecombination preferentially repairs DNA screening and paired nickases for cooperative genome double-strand breaks in the S phase of the cell cycle in engineering. Nature Biotechnology 31:833–838. human cells. Nucleic Acids Research 32:3683–3688. doi: 10.1038/nbt.2675. doi: 10.1093/nar/gkh703. Mali P, Esvelt KM, Church GM. 2013b. Cas9 as a Schellmann N, Deckert PM, Bachran D, Fuchs H, versatile tool for engineering biology. Nature Methods Bachran C. 2010. Targeted enzyme prodrug therapies. 10:957–963. doi: 10.1038/nmeth.2649. Mini Reviews in Medicinal Chemistry 10:887–904. Mali P, Yang L, Esvelt KM, Aach J, Guell M, DiCarlo JE, doi: 10.2174/138955710792007196. Norville JE, Church GM. 2013c. RNA-guided human Schliekelman P, Ellner S, Gould F. 2005. Pest control by genome engineering via Cas9. Science 339:823–826. genetic manipulation of sex ratio. Journal of Economic doi: 10.1126/science.1232033. Entomology 98:18–34. doi: 10.1603/0022-0493-98.1.18. Marshall JM, Hay BA. 2012. Confinement of gene Scott TW, Takken W, Knols BGJ, Boëte C. 2002. The drive systems to local populations: a comparative ecology of genetically modified mosquitoes. Science analysis. Journal of Theoretical Biology 294:153–171. 298:117–119. doi: 10.1126/science.298.5591.117. doi: 10.1016/j.jtbi.2011.10.032. Shalem O, Sanjana NE, Hartenian E, Shi X, Scott DA, Marshall JM. 2009. The effect of gene drive on Mikkelsen TS, Heckl D, Ebert BL, Root DE, Doench JG, containment of transgenic mosquitoes. Journal of et al. 2014. Genome-scale CRISPR-Cas9 knockout

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 20 of 21 208 Feature article Emerging technology | Concerning RNA-guided gene drives for the alteration of wild populations

screening in human cells. Science 343:84–87. endogenous human gene correction using designed doi: 10.1126/science.1247005. zinc-finger nucleases. Nature 435:646–651. Simoni A, Siniscalchi C, Chan YS, Huen DS, Russell S, doi: 10.1038/nature03556. Windbichler N, Crisanti A. 2014. Development of Wang H, Yang H, Shivalila CS, Dawlaty MM, Cheng synthetic selfish elements based on modular nucleases AW, Zhang F, Jaenisch R. 2013a. One-step generation in Drosophila melanogaster. Nucleic Acids Research of mice carrying mutations in multiple genes by 42:7461–7472. doi: 10.1093/nar/gku387. CRISPR/Cas-mediated genome engineering. Cell Sinkins SP, Gould F. 2006. Gene drive systems for 153:910–918. doi: 10.1016/j.cell.2013.04.025. insect disease vectors. Nature Reviews Genetics Wang Y, Li Z, Xu J, Zeng B, Ling L, You L, Chen Y, 7:427–435. doi: 10.1038/nrg1870. Huang Y, Tan A. 2013b. The CRISPR/Cas System Sternberg SH, Redding S, Jinek M, Greene EC, mediates efficient genome engineering in Bombyx Doudna JA. 2014. DNA interrogation by the CRISPR mori. Cell Research 23:1414–1416. doi: 10.1038/ RNA-guided endonuclease Cas9. Nature 507:62–67. cr.2013.146. doi: 10.1038/nature13011. Wang T, Wei JJ, Sabatini DM, Lander ES. 2014. Takeuchi R, Choi M, Stoddard BL. 2014. Redesign of Genetic screens in human cells using the CRISPR- extensive protein–DNA interfaces of meganucleases Cas9 system. Science 343:80–84. doi: 10.1126/ using iterative cycles of in vitro compartmentalization. science.1246981. Proceedings National Academy of the Sciences of the Werren JH. 1997. Biology of wolbachia. Annual Review USA 111:4061–4066. doi: 10.1073/pnas.1321030111. of Entomology 42:587–609. doi: 10.1146/annurev. Tan W, Carlson DF, Lancto CA, Garbe JR, Webster DA, ento.42.1.587. Hackett PB, Fahrenkrug SC. 2013. Efficient nonmeiotic WHO. 2013. World malaria report. WHO. At allele introgression in livestock using custom http://www.who.int/malaria/publications/world_ endonucleases. Proceedingsof the National Academy malaria_report_2013/en/. of Sciences USA 110:16526–16531. Windbichler N, Menichelli M, Papathanos PA, doi: 10.1073/pnas.1310478110. Thyme SB, Li H, Ulge UY, Hovde BT, Baker D, Thyme SB, Boissel SJ, Arshiya Quadri S, Nolan T, Monnat RJ Jnr, Burt A, et al. 2011. A synthetic Baker DA, Park RU, Kusak L, Ashworth J, Baker D. homing endonuclease-based gene drive system in 2013. Reprogramming homing endonuclease the human malaria mosquito. Nature 473:212–215. specificity through computational design and directed doi: 10.1038/nature09937. evolution. Nucleic Acids Research 42:2564–2576. Windbichler N, Papathanos PA, Catteruccia F, doi: 10.1093/nar/gkt1212. Ranson H, Burt A, Crisanti A. 2007. Homing Touré YT, Oduola AMJ, Sommerfeld J, Morel CM. endonuclease mediated gene targeting in Anopheles 2004. Biosafety and risk assessment in the use of gambiae cells and embryos. Nucleic Acids Research genetically modified mosquitoes for disease control. 35:5922–5933. doi: 10.1093/nar/gkm632. In: Knols BGJ, Louis C, editors. Bridging Laboratory Windbichler N, Papathanos PA, Crisanti A. 2008. and Field Research for Genetic Control of Disease Targeting the X chromosome during spermatogenesis Vectors. Springer/Frontis. p. 217–222. induces Y chromosome transmission ratio distortion Tsai SQ, Wyvekens N, Khayter C, Foden JA, Thapar V, and early dominant embryo lethality in Anopheles Reyon D, Goodwin MJ, Aryee MJ, Joung JK. 2014. gambiae. PLOS Genetics 4:e1000291. doi: 10.1371/ Dimeric CRISPR RNA-guided FokI nucleases for highly journal.pgen.1000291. specific genome editing. Nature Biotechnology Wood RJ, Cook LM, Hamilton A, Whitelaw A. 1977. 32:569–576. doi: 10.1038/nbt.2908. Transporting the marker gene re (red eye) into a UNEP. 2010. Final report of the Ad Hoc technical laboratory cage population of Aedes aegypti (Diptera: expert group on risk assessment and risk management Culicidae), using meiotic drive at the MD locus. Journal under the cartagena protocol on biosafety. At of Medical Entomology 14:461–464. http://www.cbd.int/doc/meetings/bs/bsrarm-02/ Wood RJ, Newton ME. 1991. Sex-ratio distortion official/bsrarm-02-05-en.pdf. caused by meiotic drive in mosquitoes. The American Upadhyay SK, Kumar J, Alok A, Tuli R. 2013. Naturalist 137:379–391. doi: 10.1086/285171. RNA-guided genome editing for target gene Yahara K, Fukuyo M, Sasaki A, Kobayashi I. 2009. mutations in wheat. G3 3:2233–2238. doi: 10.1534/ Evolutionary maintenance of selfish homing g3.113.008847. endonuclease genes in the absence of horizontal Urnov FD, Miller JC, Lee YL, Beausejour CM, transfer. Proceedings of the National Academy Rock JM, Augustus S, Jamieson AC, Porteus MH, of Sciences USA 106:18861–18866. Gregory PD, Holmes MC. 2005. Highly efficient doi: 10.1073/pnas.0908404106.

Esvelt et al. eLife 2014;3:e03401. DOI: 10.7554/eLife.03401 21 of 21 209