bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1 Highly efficient transgenesis with miniMos in

2 briggsae

3 Qiutao Ding1, Xiaoliang Ren1, Runsheng Li1, Luyan Chan1, Vincy WS Ho1, Yu Bi1,

4 Dongying Xie1, and Zhongying Zhao1,2*

5 1Department of Biology, Hong Kong Baptist University, Hong Kong, China; 2State Key

6 Laboratory of Environmental and Biological Analysis, Hong Kong Baptist University,

7 Hong Kong, China

8

9 *Corresponding author

10 Zhongying Zhao, PhD

11 State Key Laboratory of Environmental and Biological Analysis

12 Department of Biology, Hong Kong Baptist University

13 Hong Kong

14 Email: [email protected]

15 Phone: 852-3411-7058

16 Running title: Efficient transgenesis with miniMos in C. briggsae

17 Keywords: miniMos, C. briggsae, heat shock, transgenesis

18

19

20

1 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

21 Abstract

22 C. briggsae as a companion species for C. elegans has played an increasingly important

23 role in study of evolution of development, gene regulation and . Aided by the

24 isolation of its sister spices, it has recently been established as a model for speciation study.

25 To take full advantage of the species for comparative study, an effective transgenesis

26 method especially those with single copy insertion is important for functional comparison.

27 Here we modified a transposon-based transgenesis methodology that had been originally

28 developed in C. elegans but worked marginally in C. briggsae. By incorporation of a heat

29 shock step, the transgenesis efficiency in C. briggsae with single copy insertion is

30 comparable to that in C. elegans. We used the method to generate 54 independent insertions

31 mostly consisting of a mCherry tag over the C. briggsae genome. We demonstrated the use

32 of the tags in identifying interacting loci responsible for hybrid male sterility between C.

33 briggsae and C. nigoni when combined with the GFP tags we generated previously. Finally,

34 we demonstrated that C. briggsae has developed native immunity against the C. elegans

35 toxin, PEEL-1, but not SUP-35, making the latter a potential negative selection marker

36 against extrachromosomal array.

37

38 Summary

39 C. briggsae has been used for comparative study against C. elegans over

40 decades. Importantly, a sister species has recently been identified, with which C. briggsae

41 is able to mate and produce viable hybrid progeny. This opens the possibility of using

42 nematode species as a model for speciation study for the first time. To take full advantage

43 of C. briggsae for comparative study, an effective transgenesis method to generate single

2 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

44 copy insertion is important especially for functional comparison. An attempt was made

45 previously to generate single copy insertion with transposon-based transgenesis

46 methodology, which had been originally developed in C. elegans but with limited success

47 in C. briggsae. Here we modified the transposon-based methodology by incorporation of

48 a heat shock step, which allows us to achieve a much higher transgenesis efficiency in C.

49 briggsae with single copy insertion. We used the method to generate 54 independent

50 insertions mostly consisting of a mCherry tag over the C. briggsae genome. We

51 demonstrated the use of the tags in identifying interacting loci responsible for hybrid male

52 sterility between C. briggsae and C. nigoni when combined with the GFP tags we generated

53 previously. Finally, we demonstrated that C. briggsae has developed native immunity

54 against the C. elegans toxin, PEEL-1, but not SUP-35, making the latter a potential

55 negative selection marker against extrachromosomal array. Taken together, the modified

56 transgenesis methodology and the transgenic strains generated in this study are expected

57 to further facilitate C. briggsae as a model for comparative study or speciation study.

58

59

60 Introduction

61 As a comparative model of , C. briggsae shares a similar

62 morphology, carries a genome of comparable size (Ren et al., 2018; Ross et al., 2011; Stein

63 et al., 2003; Yin et al., 2018) and adopts similar developmental pattern to that of C. elegans

64 (Zhao et al., 2008). As a , C. elegans genome has been subject to intensive

65 manipulations with various tools, including random mutagenesis with ultraviolet light

66 coupled with trimethylpsoralen (UV/TMP) (Thompson et al., 2013), low copy insertion of

3 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

67 transgenes with biolistic bombardment (Hochbaum, Ferguson, & Fisher, 2010; Praitis,

68 Casey, Collar, & Austin, 2001; Radman, Greiss, & Chin, 2013), targeted single copy

69 insertion using transcription activator-like effector nucleases (TALEN) system (Lo et al.,

70 2013; Wood et al., 2011), or using CRISPR/Cas9 system (Chiu, Schwartz, Antoshechkin,

71 & Sternberg, 2013; Cho, Lee, Carroll, Kim, & Lee, 2013; Dickinson & Goldstein, 2016;

72 Dickinson, Pani, Heppert, Higgins, & Goldstein, 2015; Waaijers et al., 2013), Mos1-

73 mediated single-copy insertion (MosSCI) (Frøkjær-Jensen et al., 2008) as well as random

74 single-copy insertion with miniMos (Frøkjær-Jensen et al., 2014). Many of these tools have

75 been adopted in C. briggsae for genome editing with a comparable or a much-reduced

76 successful rate. For example, biolistic bombardment was successfully adopted in C.

77 briggsae transgenesis with comparable efficiency, whereas single-copy insertion with

78 miniMos demonstrated a much lower efficiency in C. briggsae than in C. elegans (Bi et al.,

79 2015; Frøkjær-Jensen et al., 2014; Zhao et al., 2010). An efficient transgenesis method

80 with single copy insertion is essential for comparative functional study of biology between

81 C. elegans and C. briggsae. Recent work in Caenorhabditis species has demonstrated that

82 heat shock treatment or compromised function in heat shock pathway significantly

83 increased the frequency of transposition (Ryan, Brownlie, & Whyard, 2016), raising the

84 possibility of improving miniMos-based transgenesis in C. briggsae by incorporation of a

85 heat shock step or inhibition of activities of heat shock proteins.

86 Isolation of C. briggsae sister species, C. nigoni, with which it can mate and produce viable

87 progeny, paves the way of speciation study using nematode as a model for the first time

88 (Kozlowska, Ahmad, Jahesh, & Cutter, 2012; Woodruff, Eke, Baird, Félix, & Haag, 2010).

89 To isolate post-zygotic hybrid incompatible (HI) loci between C. briggsae and C. nigoni,

4 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

90 a dominant and visible marker is required for targeted introgression, in which the genome

91 of one species is labeled with the marker and repeatedly backcrossed into the other to

92 isolate the HI loci. The marker greatly facilitates tracing of its linked genomic fragment

93 during backcrossing and its associated HI phenotype. Given numerous such markers are

94 required over a genome, generation of such markers using targeted single-copy insertion

95 becomes an enormous burden. This is because that the efficiency for targeted single-copy

96 insertion is relatively low. As an alternative, over 100 GFP markers were inserted into the

97 C. briggsae genome with biolistic bombardment, which allowed genome-wide mapping of

98 HI loci between the two species (Bi et al., 2015; Ren et al., 2018; Yan, Bi, Yin, & Zhao,

99 2012).

100 Biolistic bombardment using cbr-unc-119 (Bi et al., 2015; Zhao et al., 2010) generates low

101 copy number of transgene into genome but requires tedious preparations. It also

102 needs to wait for several weeks before screening for transformed animal. In addition,

103 presumably due to dramatic mechanic shearing, the transgenic often suffer from

104 chromosomal rearrangements (Bi et al., 2015; Ren et al., 2018; Tyson et al., 2018).

105 Importantly, the insertion site cannot be precisely determined due to unknown copy number

106 of the transgene and its arrangement within host genome. As such, the transgene insertion

107 sites were estimated by genotyping introgression boundary with PCR using species-

108 specific primers (Bi et al., 2015; Yan et al., 2012). Unfortunately, the divergence between

109 the C. briggsae and C. nigoni is too far to allow efficient recombination during

110 backcrossing, resulting in a poor mapping resolution of the insertion site (Bi et al., 2019,

111 2015; Yan et al., 2012).

5 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

112 We have recently demonstrated that inter-chromosomal interactions are involved in hybrid

113 male sterility between C. briggsae and C. nigoni (Bi et al., 2019; Li et al., 2016). For

114 example, a C. briggsae X chromosome fragment in an otherwise C. nigoni background as

115 an introgression leads to male sterility. Hybrid F1 male sterility can be rescued by the

116 presence of the introgression and another C. briggsae genomic fragment on its

117 Chromosome II. We speculate that such interacting loci might be common in producing

118 HI. Availability of the dual-color labeling system makes it possible for systematic mapping

119 such interacting loci. This is because it facilitates tracking of co-segregation of the two loci.

120 It is worth noting that most existing markers in the C. briggsae genome were derived from

121 GFP, preventing effective screening for such loci. Therefore, C. briggsae animals bearing

122 a visible marker other than the GFP are desired.

123 Transposon-mediated single copy insertion at random genomic position with high

124 efficiency has been successfully developed in C. elegans (Frøkjær-Jensen et al., 2014). A

125 fly transposon Mos1 was truncated with minimal transposon sequences, termed as miniMos

126 cargo vector, in order to boost its capacity of carrying foreign DNA (Frøkjær-Jensen et al.,

127 2014). Single copy of transgene can be inserted into host genome at high frequency via co-

128 injecting a transposase-expressing vector and miniMos cargo vector into C. elegans gonad.

129 The method was also adopted in C. briggsae but with a much lower efficiency with

130 unknown reasons (Frøkjær-Jensen et al., 2014). Attempts have been made to improve

131 insertion frequency in C. briggsae. For example, the promoter (cel-Peft-3) driving

132 transposase expression was substituted with C. briggsae promoter cbr-Peft-3 or cbr-Ppie-

133 1 in order to boost transposase expression. However, the insertion frequency did not

134 improve as expected. A highly efficient method for inserting single-copy transgene has yet

6 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

135 to be established in C. briggsae. To facilitate screen for single-copy insertion, a negative

136 selection marker cel-Phsp-16.41::peel-1 was co-injected into animals to kill

137 extrachromosomal array-bearing worms during screening (Frøkjær-Jensen, Davis, Ailion,

138 & Jorgensen, 2012). However, the killing efficiency of the negative selection marker peel-

139 1 has not been investigated in C. briggsae.

140 This study established a highly efficient methodology for inserting single-copy transgene

141 into the C. briggsae genome based on miniMos. Negative selection marker against

142 transgene consisting of extrachromosomal array was also explored with limited success.

143

144 Materials and Methods

145 Nematode strains

146 C. elegans, C. briggsae and C. nigoni wild isolates used in this study were N2, AF16 and

147 JU1421, respectively. The C. briggsae dpy-5(zzy0580) knockout strain ZZY0580 was

148 generated using CRISPR/Cas9 with AF16 followed by backcrossing to AF16 for three

149 generations. Introgression strains used were ZZY10330 (zzyIR10330 (X [cbr-myo-2p::gfp,

150 cbr-unc-119(+)]), AF16>JU1421) (Bi et al., 2015), ZZY10377 (zzyIR10377 (X [cbr-myo-

151 2p::mCherry, neoR]) , AF16>JU1421) from C. briggsae transgenic strain ZZY0777 (Table

152 S2). Introgression strains ZZY10353 (zzyIR10353 (II [cbr-myo-2p::gfp, cbr-unc-119(+)]),

153 AF16>JU1421) (Bi et al., 2015) and ZZY10382 (zzyIR10382 (II [cbr-myo-2p::mCherry,

154 neoR]), AF16>JU1421) were generated from C. briggsae transgenic strain ZZY0782

155 (Table S2). Details for all the transgenic strains generated in this study were included in

156 Table S2. All the C. elegans and C. briggsae lines were maintained on 1.5% nematode

157 growth medium (NGM) seeded with E. coli OP50 at room temperature and in a 25 °C

7 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

158 incubator, respectively, unless specified otherwise.

159 Introgression

160 Introgression was performed as described (Bi et al., 2015). Specifically, C. briggsae

161 transgenic strains, ZZY0777 and ZZY0782 each expressing a single-copy of Pmyo-

162 2::mCherry located on the X and chromosome II, respectively, were backcrossed to C.

163 nigoni (JU1421) for 15 generations to give rise to ZZY10377 and ZZY10382. Introgression

164 boundaries were mapped as described (Yan et al., 2012).

165 Molecular Biology and transgenesis

166 pZZ203 was derived from pZZ0031 (Yan et al., 2012) by removing cbr-Pmyo-

167 2::gfp::his-72 UTR by cutting using KpnI and ApaI (NEB). The digested pZZ0031

168 backbone carrying cbr-unc-119(+) was gel purified, blunt end repaired, and self-ligated

169 to give rise to pZZ203. pZZ160 and pZZ161 was derived from pCFJ910[NeoR] and

170 pCFJ909[cbr-unc-119(+)], respectively, by inserting both cbr-Pmyo-2::gfp::his-72 UTR

171 and cbr-dpy-5(+) into minimal Mos1 transposon. pZZ184 was derived from the pZZ160

172 by replacing gfp::his-72 UTR and cbr-dpy-5(+) with mCherry::unc-54 UTR from pGH8

173 (Frøkjær-Jensen et al., 2008). The plasmids pZZ185 and pZZ196 for negative selection

174 were derived from pCFJ601[Peft-3::Mos1 transposase::tbb-2 UTR] by replacing the

175 Mos1 with sup-35 (Ben-David, Burga, & Kruglyak, 2017) and peel-1 (Seidel et al., 2011;

176 Seidel, Rockman, & Kruglyak, 2008), respectively. The plasmid kit containing pCFJ601,

177 pCFJ909, pCFJ910, pGH8, and pMA122 was acquired from Addgene (cat#

178 1000000031). Vector pZZ113 containing sgRNA expression cassette against cbr-dpy-5

179 was derived from PU6::unc-119_sgRNA (Addgene plasmid # 46169) as described

180 (Friedland et al., 2013). The cbr-dpy-5 gRNA target sequence is

8 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

181 GGAGCCCCAGGAGAGCCAGG. Primers used for construct building were listed in

182 Table S1. An overview of plasmid compositions was shown in Fig. S1.

183 Plasmids for microinjection were extracted using the PureLink HQ Mini Plasmid

184 Purification Kit or HiPure Plasmid Midiprep kit (Invitrogen). Genomic DNAs were

185 extracted from mix-staged animals with PureLink Genomic DNA Mini Kit (Invitrogen).

186 Injection mixture for transformation consisted of miniMos cargo vector at 20 ng/µl, Mos1

187 transposase-expressing plasmid at 50 ng/µl, red or green fluorescent co-injection markers

188 each at 10 ng/µl. pZZ203 was added to 80 ng/µl with a total DNA concentration of 170

189 ng/µl. Plasmids pMA122, pZZ185, and pZZ196 were built to test their use as a negative

190 selection marker. Neomycin stock solution of 12.5 mg/ml was made with G418 powder

191 (ThermoFisher) in water. For detailed transformation steps, see Fig. 1.

192 Quantification of P0 insertion frequency

193 Three days after microinjection, plates without F1 transgenic progeny were discarded and

194 excluded from the number of injected P0 animals. The F2 animals were screened for

195 miniMos insertion under stereo-microscope based on reporter expression or phenotypic

196 rescue, and the absence of co-injection markers. Animals confirmed with transgene

197 insertion from the same P0 plate were treated as a single independent insertion. The P0

198 insertion frequency was measured through dividing the number of independent insertions

199 by the total number of injected P0 animals. Mapping of insertion site was performed as

200 described (Frøkjær-Jensen et al., 2014). The C. briggsae cb4 genome (Ross et al., 2011)

201 was used to report the chromosomal coordinates.

202 Selection marker for transformation

203 For selection with Neomycin (NeoR), injected animals were allowed to recover overnight

9 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

204 (12 to 16 hours) followed by heat shock treatment on a 35 °C water bath for three hours.

205 The heat shock treated plates were incubated at room temperature for half hour to recover.

206 Then 500 μl of 12.5 mg/ml G418 solution was added to each plate. Plates were incubated

207 at 25 °C for 6 to 7 days. Twelve F2 animals were singled onto a new NGM plate from the

208 P0 plates crowded with animals. Homozygosity was checked based on reporter expression

209 in F3.

210 For selection with cbr-dpy-5, all rescued dpy-5(+) F1 animals were singled onto individual

211 plates. Twelve F2 animals were singled onto new plate from each F1 plate in which over

212 50% animals were dpy-5(+) ones. F2 plates with 100% dpy-5(+) animals that did not

213 express visible co-injection markers were kept for genotyping.

214 Development of negative selection marker for extrachromosomal array

215 A fusion PCR product was made between a cassette consisting of peel-1::tbb-2 UTR or

216 gfp::tbb-2 UTR and the C. briggsae syntenic region of hsp-16.41 promoter with the primers

217 listed in Table S1. Additional details of C. briggsae syntenic region of hsp-16.41 promoter

218 sequence screening are available in Fig. S2. Injection mixture was made with pMA122,

219 pZZ185, pZZ196 or the fusion PCR product at 70 ng/µl, fluorescent co-injection marker

220 vector pZZ161 at 20 ng/µl, pGH8, pCFJ90, pCFJ104 at 10 ng/µl each, and pZZ203 at 50

221 ng/µl.

222 To examine the lethality in C. briggsae after heat shock, an array-bearing line was

223 generated and embryos harvested for synchronization. The synchronized L1 animals were

224 divided into two groups, one was subjected to heat shock treatment at 35 °C for 3 hours

225 before transferring into 25 °C incubator, and the other was incubated at 25 °C without heat

226 shock treatment as a control. C. elegans heat shock treatment was performed at 34 °C for

10 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

227 3 hours. The ratio of surviving adults carrying an array out of all progeny were calculated

228 in both control and heat-shock treated animals.

229

230 Results

231 Heat shock treatment significantly increased the efficiency of transgene insertion

232 To improve the transgenesis efficiency using miniMos in C. briggsae, we started the

233 transgenesis by repeating the steps basically as described previously (Frøkjær-Jensen et al.,

234 2014), i.e., injection of a vector carrying cbr-Pmyo-2::gfp along with a transposase-

235 expressing vector into cbr-unc-119 deletion mutant strain RW20000 (Zhao et al., 2010).

236 One of major complications associated with the cbr-unc-119 as an injection selection

237 marker was the low transformation efficiency even for the formation of extrachromosomal

238 array. This was the case using the rescuing fragment from either C. elegans or C. briggsae

239 (data not shown). We speculated that the low insertion rate associated with the selection

240 marker could be related to the low transformation efficiency of single copy insertion.

241 To improve transformation efficiency, we generated another injection selection marker

242 cbr-dpy-5 using CRISPR/Cas9. This was based on the observation that dpy-5 had been

243 used as a very efficient selection marker in C. elegans microinjection (Hunt-Newbury et

244 al., 2007; Zhao, Sheps, Ling, Fang, & Baillie, 2004). As expected, cbr-dpy-5 worked as an

245 effective selection marker for microinjection. We successfully obtained three independent

246 transgenic lines with single copy insertion using the marker after injecting 35 animals.

247 However, due to lack of negative selection maker against extrachromosomal array, it is

248 tedious to screen for the rescued animals out of all progeny.

11 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

249 We then performed miniMos-based single-copy insertion using neomycin as a selection

250 marker as described (Frøkjær-Jensen et al., 2014; Giordano-Santini et al., 2010). The

251 method provided a key advantage in simplifying selection of the transformed animals

252 because all un-rescued animals were killed, but the overall efficiency of obtaining single-

253 copy insertion was not improved compared with previous studies (Fig. 2).

254 Given that heat shock treatment significantly increased transposition frequency (Ryan et

255 al., 2016), we reasoned that inclusion of a heat shock step might boost the successful rate

256 of transgenesis in C. briggsae based on miniMos, which is a modified transposon. As

257 expected, inclusion of a step of heat shock treatment, i.e., 35 °C for 3 hrs, we were able to

258 significantly increase the frequency of single-copy insertion of Pmyo-2::gfp or Pmyo-

259 2::mCherry (Figs. 1 & 2). Insertion frequency increased from lower than 5% without heat

260 shock to approximately 25% with heat shock for both constructs, indicating that the heat

261 shock treatment was essential for increasing transposition rate.

262

263 A large collection of single copy insertions expressing Pmyo-2::mCherry were generated

264 in C. briggsae

265 To complement the genetic resources mainly consisting of GFP insertions over the C.

266 briggsae we generated previously (Bi et al., 2015), we decided to generate a cohort of

267 insertions randomly distributed over the C. briggsae genome with a focus on generation of

268 insertion expressing a fluorescent protein other than GFP, i.e., Pmyo-2::mCherry. This

269 would be particularly useful for isolation of interacting loci responsible for hybrid male

270 sterility as detailed below. To this end, we generated a total of 54 insertions consisting of

271 9 Pmyo-2::gfp and 45 Pmyo-2::mCherry that were able to be uniquely mapped to the C.

12 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

272 briggsae genome (Fig. 3 and Table S2). The markers show roughly random distribution

273 over the C. briggsae genome.

274

275 Dual-coloured marking system facilitates screen for interacting loci responsible for HI

276 loci

277 We demonstrated that genetic interaction between X Chromosome and autosome was

278 essential for hybrid male sterility between C. briggsae and C. nigoni (Bi et al., 2019).

279 Identification of such interacting loci is challenging without proper markers on interacting

280 chromosomes. We showed that the interaction between independent GFP-labelled C.

281 briggsae introgression fragments on two different chromosomes was essential for male

282 fertility in C. nigoni, but it was difficult to pinpoint these interacting loci systematically.

283 Availability of such dual-colour labelled strains would greatly facilitate the identification

284 of such interactions. To help illustrate this point, we tried to recapitulate the interaction we

285 identified earlier by using two C. nigoni strains ZZY10377 and ZZY10382 that carry a

286 mCherry-labelled introgression fragment derived from the right arm of the C. briggsae X

287 and chromosome II (Fig. 4), respectively, between which an interaction was found (Bi et

288 al., 2019). We crossed two introgression strains, one labelled with GFP and the other with

289 mCherry in reciprocal way. We then examined the fertility of the males that simultaneously

290 carried both introgressions. We found that the males carrying the both loci were mostly

291 fertile, whereas the males carrying a single introgression of GFP or mCherry inserted on

292 the X chromosome were sterile, indicating that the dual-colour labelled introgressions did

293 recapitulate the interaction. Presence of the dual-coloured markers paves the way for

13 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

294 genome-wide identification of any other interacting loci responsible for the hybrid

295 incompatibilities.

296

297 C. briggsae appeared to develop native immunity against PEEL-1

298 For transgene insertion using MosSCI or miniMos in C. elegans, a sperm-derived toxin

299 gene, peel-1, is used as a negative selection marker to effectively kill extrachromosomal

300 array-bearing worms (Frøkjær-Jensen et al., 2012, 2014; Seidel et al., 2011), greatly

301 reducing the burden of screening for transgenic strains carrying an insertion out of all

302 transgenic animals, including those carrying an extrachromosomal array.

303 To adopt the negative selection marker in C. briggsae, we first compared the killing effect

304 of PEEL-1 in C. elegans and C. briggsae through its forced expression driven by a C.

305 elegans heat shock promoter, Phsp-16.41. We generated transgenic lines carrying

306 extrachromosomal array consisting of the PEEL-1-expressing vector in both C. elegans

307 and C. briggsae. The synchronized L1 animals were subjected to heat shock at 34 °C and

308 35 °C for C. elegans and C. briggsae, respectively. The killing effect in C. elegans was

309 comparable to that reported previously (Seidel et al., 2011), but no apparent killing was

310 observed in C. briggsae after heat shock treatment (Fig. 5A).

311 To further investigate what caused the failure of PEEL-1 to kill C. briggsae, we replaced

312 C. elegans heat shock promoter, Phsp-16.41, with its C. briggsae equivalent (Fig. S2) and

313 generated the transgenic lines. Again, we did not observe significant increase in killing (Fig.

314 5A). We reasoned that the C. briggsae heat shock promoter might not be a functional

315 equivalent of Phsp-16.41. To examine whether the C. briggsae syntenic heat shock

316 promoter responds to heat shock treatment, we generated transgenic lines carrying

14 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

317 extrachromosomal array consisting of a fusion between the C. briggsae heat shock

318 promoter and GFP. We did see induced GFP expression in the transgenic animals after heat

319 shock treatment in both C. elegans and C. briggsae (data not shown), suggesting that C.

320 briggsae somehow develops immunity against PEEL-1. To further investigate C.

321 briggsae’s immunity against PEEL-1, we generated transgenic C. briggsae animals

322 expressing peel-1 driven by an effective promoter, Peft-3, which was known to able to drive

323 expression in C. briggsae (Frøkjær-Jensen et al., 2012, 2014). Again, we did not observe a

324 significant killing effect (Fig. 5B). Taken together, it seems that C. briggsae develops

325 native immunity against PEEL-1, preventing it from being used as a negative selection

326 marker against extrachromosomal array.

327 Limited success of using sup-35 as a negative selection marker for extrachromosomal

328 array

329 In addition to peel-1, a maternal-effect toxin, SUP-35, that kills developing embryo, has

330 recently been identified in C. elegans (Ben-David et al., 2017). To test the killing efficiency

331 of SUP-35 in C. briggsae, we made a sup-35-expressing vector driven by the C. elegans

332 eft-3 promoter, Peft-3 (Fig. S1). We injected the construct along with fluorescence co-

333 injection markers into both C. elegans and C. briggsae. As expected, nearly all F1 embryos

334 expressing the fluorescence markers arrested at late embryogenesis or as early larvae for

335 both C. elegans and C. briggsae (Fig. 5C), indicating that SUP-35 is an effective toxin in

336 C. briggsae and has a potential to be developed as a negative selection marker against

337 extrachromosomal array in C. briggsae. To this end, we generated transgenic strains in C.

338 elegans and C. briggsae that carry an extrachromosomal array, which consists of the sup-

339 35-expressing vector driven again by C. briggsae equivalent of hsp-16.41 promoter (Fig.

15 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

340 S2). However, we did not observe any significant killing effect in C. elegans after heat

341 shock treatment (Fig. 5D). Because C. elegans expresses both PEEL-1 and its antidote

342 ZEEL-1 only transiently in embryo, but expresses SUP-35 and its antidote PHA-1

343 throughout its life cycle (Gerstein et al., 2010). It is possible that the postembryonic PHA-

344 1 is sufficient to neutralize the toxicity of SUP-35 induced by heat shock treatment in C.

345 elegans. It is also possible that the C. briggsae heat shock promoter may not respond to

346 heat shock as effectively as its C. elegans equivalent in C. elegans. Consistent with this,

347 the heat shock treatment showed a significant killing effect in C. briggsae carrying the

348 same construct, i.e., from 66.4% to 39.6% though it is not efficient enough to serve as a

349 negative selection marker.

350

351 Discussion

352 Given the similar morphology, physiology and developmental program between C. elegans

353 and C. briggsae, methods for C. elegans transgenesis are expected to be transferrable to C.

354 briggsae with minimal modification. However, there is an exception to this, which is the

355 case for miniMos mediated single-copy insertion (Frøkjær-Jensen et al., 2014).

356 Potential cause of low efficiency of transgenesis in C. briggsae mediated by miniMos.

357 It has been well established that transposon mobility can be significantly increased across

358 species after exposure to biotic or abiotic stress (Bouvet, Jacobi, Plourde, & Bernier, 2008;

359 Liu, Chu, Choudary, & Schmid, 1995; Walbot, 1999). Heat shock proteins serve as

360 molecular chaperone to facilitate folding of other proteins into their appropriate

361 conformations following heat exposure, thus buffering the subsequent phenotypic changes

362 (Erlejman, Lagadari, Toneatto, Piwien-Pilipuk, & Galigniana, 2014). Perturbation of heat

16 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

363 shock protein also increases transposition frequency in a similar way to that of heat

364 treatment itself. It is possible that under optimal growth condition, C. briggsae has a more

365 robust system to maintain genome integrity in its germline than C. elegans, which prevents

366 its genome from environment-induced damage or editing. Improvement of transgenesis

367 efficiency by heat shock treatment or inhibition of heat shock pathway could be at the cost

368 of compromised genome integrity. One important way to preserve genome integrity in the

369 germline is through Piwi-interacting RNA (piRNA), which constitutes one of the major

370 regulatory molecules that curb the transposon activities (Ishizu, Siomi, & Siomi, 2012).

371 Consistent with this, functional perturbation of Hsp90 attenuates the piRNA silencing

372 mechanism, leading to transposon activation and the induction of morphological mutants

373 in (Specchia et al., 2010). Therefore, the transgenic strains generated using heat

374 shock treatment should be subjected to rigorous backcrossing with wild isolate before any

375 functional characterization.

376

377 Differential responses to PEEL-1 and SUP-35 between C. elegans and C. briggsae

378 It is intriguing that C. briggsae responds differentially to the overexpressed paternal toxin

379 PEEL-1 and maternal toxin SUP-35 from C. elegans. In C. elegans, PEEL-1 was believed

380 suspected to function as a calcium pump on the cell membrane by generating a membrane

381 pore, which leads to the release of intracellular calcium (Seidel et al., 2011). The maternal-

382 effect toxin, SUP-35, kills developing embryos, but the molecular mechanism of its toxicity

383 remains unclear (Ben-David et al., 2017). Our data showed that C. briggsae somehow

384 develops native immunity against PEEL-1 but not SUP-35 (Fig. 5). However, forced

385 expression of SUP-35 driven by the C. briggsae heat shock promoter did not mediate

17 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

386 complete killing of C. briggsae after heat shock treatment (Fig. 5D). Two possible reasons

387 account for the ineffective killing. First, time windows for SUP-35 killing seems to be

388 narrow, i.e., at certain stage of embryogenesis. This is why a constitutive promoter, i.e.,

389 eft-3 promoter is able to drive SUP-35 expression and mediate complete killing. For the

390 heat shock promoter, its response to heat shock may be more effective during larval

391 development than embryogenesis. Second, heat-shock induced expression of SUP-35 may

392 not provide enough dosage to kill C. briggsae larvae. Further optimization of heat shock

393 condition or choice of a different heat-shock promoter is necessary to develop an effective

394 negative selection marker against extrachromosomal array in C. briggsae.

395

396 Acknowledgments

397 We thank Ms Cindy Tan for logistic support and members of Zhao’s lab for helpful

398 comments. Some strains were provided by the CGC, which is funded by NIH Office of

399 Research Infrastructure Programs (P40 OD010440).

400

401 Reference

402 Ben-David, E., Burga, A., & Kruglyak, L. (2017). A maternal-effect selfish genetic 403 element in Caenorhabditis elegans. Science, 356(6342), 1051–1055. 404 https://doi.org/10.1126/science.aan0621 405 Bi, Y., Ren, X., Li, R., Ding, Q., Xie, D., & Zhao, Z. (2019). Specific Interactions 406 Between Autosome and X Cause Hybrid Male Sterility in Caenorhabditis Species . 407 Genetics, genetics.302202.2019. https://doi.org/10.1534/genetics.119.302202 408 Bi, Y., Ren, X., Yan, C., Shao, J., Xie, D., & Zhao, Z. (2015). A Genome-wide hybrid 409 incompatibility landscape between and C. nigoni. PLoS 410 Genetics, 11(2), e1004993. https://doi.org/10.1371/journal.pgen.1004993 411 Bouvet, G. F., Jacobi, V., Plourde, K. V., & Bernier, L. (2008). Stress-induced mobility 412 of OPHIO1 and OPHIO2, DNA transposons of the Dutch elm disease fungi. Fungal 413 Genetics and Biology. https://doi.org/10.1016/j.fgb.2007.12.007 414 Chiu, H., Schwartz, H. T., Antoshechkin, I., & Sternberg, P. W. (2013). Transgene-Free 415 Genome Editing in Caenorhabditis elegans Using CRISPR-Cas. Genetics, 195(3),

18 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

416 1167–1171. https://doi.org/10.1534/genetics.113.155879 417 Cho, S. W., Lee, J., Carroll, D., Kim, J.-S., & Lee, J. (2013). Heritable gene knockout in 418 Caenorhabditis elegans by direct injection of Cas9-sgRNA ribonucleoproteins. 419 Genetics, 195(3), 1177–1180. https://doi.org/10.1534/genetics.113.155853 420 Dickinson, D. J., & Goldstein, B. (2016). CRISPR-based methods for caenorhabditis 421 elegans genome engineering. Genetics, 202(3), 885–901. 422 https://doi.org/10.1534/genetics.115.182162 423 Dickinson, D. J., Pani, A. M., Heppert, J. K., Higgins, C. D., & Goldstein, B. (2015). 424 Streamlined genome engineering with a self-excising drug selection cassette. 425 Genetics, 200(4), 1035–1049. https://doi.org/10.1534/genetics.115.178335 426 Erlejman, A. G., Lagadari, M., Toneatto, J., Piwien-Pilipuk, G., & Galigniana, M. D. 427 (2014). Regulatory role of the 90-kDa-heat-shock protein (Hsp90) and associated 428 factors on gene expression. Biochimica et Biophysica Acta - Gene Regulatory 429 Mechanisms. https://doi.org/10.1016/j.bbagrm.2013.12.006 430 Friedland, A. E., Tzur, Y. B., Esvelt, K. M., Colaiácovo, M. P., Church, G. M., & 431 Calarco, J. A. (2013). Heritable genome editing in C. elegans via a CRISPR-Cas9 432 system. Nature Methods, 10(8), 741–743. https://doi.org/10.1038/nmeth.2532 433 Frøkjær-Jensen, C., Davis, M. W., Ailion, M., & Jorgensen, E. M. (2012). Improved 434 Mos1-mediated transgenesis in C. elegans. Nature Methods. 435 https://doi.org/10.1038/nmeth.1865 436 Frøkjær-Jensen, C., Davis, M. W., Sarov, M., Taylor, J., Flibotte, S., LaBella, M., … 437 Jorgensen, E. M. (2014). Random and targeted transgene insertion in Caenorhabditis 438 elegans using a modified Mos1 transposon. Nature Methods, 11(5), 529–534. 439 https://doi.org/10.1038/nmeth.2889 440 Frøkjær-Jensen, C., Wayne Davis, M., Hopkins, C. E., Newman, B. J., Thummel, J. M., 441 Olesen, S. P., … Jorgensen, E. M. (2008). Single-copy insertion of transgenes in 442 Caenorhabditis elegans. Nature Genetics, 40(11), 1375–1383. 443 https://doi.org/10.1038/ng.248 444 Gerstein, M. B., Lu, Z. J., Van Nostrand, E. L., Cheng, C., Arshinoff, B. I., Liu, T., … 445 Waterston, R. H. (2010). Integrative analysis of the Caenorhabditis elegans genome 446 by the modENCODE project. Science. https://doi.org/10.1126/science.1196914 447 Giordano-Santini, R., Milstein, S., Svrzikapa, N., Tu, D., Johnsen, R., Baillie, D., … 448 Dupuy, D. (2010). An antibiotic selection marker for nematode transgenesis. Nature 449 Methods, 7(9), 721–723. https://doi.org/10.1038/nmeth.1494 450 Hochbaum, D., Ferguson, A. a, & Fisher, A. L. (2010). Generation of transgenic C. 451 elegans by biolistic transformation. Journal of Visualized Experiments : JoVE, (42), 452 1–5. https://doi.org/10.3791/2090 453 Hunt-Newbury, R., Viveiros, R., Johnsen, R., Mah, A., Anastas, D., Fang, L., … 454 Moerman, D. G. (2007). High-throughput in vivo analysis of gene expression in 455 Caenorhabditis elegans. PLoS Biology. https://doi.org/10.1371/journal.pbio.0050237 456 Ishizu, H., Siomi, H., & Siomi, M. C. (2012). Biology of Piwi-interacting RNAs: New 457 insights into biogenesis and function inside and outside of germlines. Genes and 458 Development. https://doi.org/10.1101/gad.203786.112 459 Kozlowska, J. L., Ahmad, A. R., Jahesh, E., & Cutter, A. D. (2012). Genetic variation for 460 postzygotic reproductive isolation between caenorhabditis briggsae and 461 caenorhabditis sp. 9. Evolution. https://doi.org/10.1111/j.1558-5646.2011.01514.x

19 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

462 Li, R., Ren, X., Bi, Y., Ho, V. W. S., Hsieh, C. L., Young, A., … Zhao, Z. (2016). 463 Specific down-regulation of spermatogenesis genes targeted by 22G RNAs in hybrid 464 sterile males associated with an X-Chromosome introgression. Genome Research, 465 26(9), 1219–1232. https://doi.org/10.1101/gr.204479.116 466 Liu, W. man, Chu, W. ming, Choudary, P. V., & Schmid, C. W. (1995). Cell stress and 467 translational inhibitors transiently increase the abundance of mammalian SINE 468 transcripts. Nucleic Acids Research. https://doi.org/10.1093/nar/23.10.1758 469 Lo, T.-W., Pickle, C. S., Lin, S., Ralston, E. J., Gurling, M., Schartner, C. M., … Meyer, 470 B. J. (2013). Precise and heritable genome editing in evolutionarily diverse 471 using TALENs and CRISPR/Cas9 to engineer insertions and deletions. 472 Genetics, 195(2), 331–348. https://doi.org/10.1534/genetics.113.155382 473 Praitis, V., Casey, E., Collar, D., & Austin, J. (2001). Creation of low-copy integrated 474 transgenic lines in Caenorhabditis elegans. Genetics, 157(3), 1217–1226. Retrieved 475 from http://www.ncbi.nlm.nih.gov/pubmed/11238406 476 Radman, I., Greiss, S., & Chin, J. W. (2013). Efficient and Rapid C. elegans Transgenesis 477 by Bombardment and Hygromycin B Selection. PLoS ONE, 8(10). 478 https://doi.org/10.1371/journal.pone.0076019 479 Ren, X., Li, R., Wei, X., Bi, Y., Ho, V. W. S., Ding, Q., … Zhao, Z. (2018). Genomic 480 basis of recombination suppression in the hybrid between Caenorhabditis briggsae 481 and C. nigoni. Nucleic Acids Research, 46(3), 1295–1307. 482 https://doi.org/10.1093/nar/gkx1277 483 Ross, J. A., Koboldt, D. C., Staisch, J. E., Chamberlin, H. M., Gupta, B. P., Miller, R. 484 D., … Haag, E. S. (2011). Caenorhabditis briggsae recombinant inbred line 485 genotypes reveal inter-strain incompatibility and the evolution of recombination. 486 PLoS Genetics. https://doi.org/10.1371/journal.pgen.1002174 487 Ryan, C. P., Brownlie, J. C., & Whyard, S. (2016). Hsp90 and physiological stress are 488 linked to autonomous transposon mobility and heritable genetic change in 489 nematodes. Genome Biology and Evolution, 8(12), 3794–3805. 490 https://doi.org/10.1093/gbe/evw284 491 Seidel, H. S., Ailion, M., Li, J., van Oudenaarden, A., Rockman, M. V., & Kruglyak, L. 492 (2011). A novel sperm-delivered toxin causes late-stage embryo lethality and 493 transmission ratio distortion in C. elegans. PLoS Biology, 9(7), e1001115. 494 https://doi.org/10.1371/journal.pbio.1001115 495 Seidel, H. S., Rockman, M. V., & Kruglyak, L. (2008). Widespread genetic 496 incompatibility in C. elegans maintained by balancing selection. Science, 319(5863), 497 589–594. https://doi.org/10.1126/science.1151107 498 Semple, J. I., Garcia-Verdugo, R., & Lehner, B. (2010). Rapid selection of transgenic C. 499 elegans using antibiotic resistance. Nature Methods, 7(9), 725–727. 500 https://doi.org/10.1038/nmeth.1495 501 Specchia, V., Piacentini, L., Tritto, P., Fanti, L., Dalessandro, R., Palumbo, G., … 502 Bozzetti, M. P. (2010). Hsp90 prevents phenotypic variation by suppressing the 503 mutagenic activity of transposons. Nature. https://doi.org/10.1038/nature08739 504 Stein, L. D., Bao, Z., Blasiar, D., Blumenthal, T., Brent, M. R., Chen, N., … Waterston, 505 R. H. (2003). The genome sequence of Caenorhabditis briggsae: A platform for 506 comparative genomics. PLoS Biology, 1(2). 507 https://doi.org/10.1371/journal.pbio.0000045

20 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

508 Thompson, O., Edgley, M., Strasbourger, P., Flibotte, S., Ewing, B., Adair, R., … 509 Waterston, R. H. (2013). The million mutation project: A new approach to genetics 510 in Caenorhabditis elegans. Genome Research. https://doi.org/10.1101/gr.157651.113 511 Tyson, J. R., O’Neil, N. J., Jain, M., Olsen, H. E., Hieter, P., & Snutch, T. P. (2018). 512 MinION-based long-read sequencing and assembly extends the Caenorhabditis 513 elegans reference genome. Genome Research, 28(2), 266–274. 514 https://doi.org/10.1101/gr.221184.117 515 Waaijers, S., Portegijs, V., Kerver, J., Lemmens, B. B. L. G., Tijsterman, M., van den 516 Heuvel, S., & Boxem, M. (2013). CRISPR/Cas9-targeted mutagenesis in 517 Caenorhabditis elegans. Genetics, 195(3), 1187–1191. 518 https://doi.org/10.1534/genetics.113.156299 519 Walbot, V. (1999). UV-B damage amplified by transposons in maize. Nature. 520 https://doi.org/10.1038/17043 521 Wood, A. J., Lo, T. W., Zeitler, B., Pickle, C. S., Ralston, E. J., Lee, A. H., … Meyer, B. 522 J. (2011). Targeted genome editing across species using ZFNs and TALENs. 523 Science. https://doi.org/10.1126/science.1207773 524 Woodruff, G. C., Eke, O., Baird, S. E., Félix, M. A., & Haag, E. S. (2010). Insights into 525 species divergence and the evolution of hermaphroditism from fertile interspecies 526 hybrids of Caenorhabditis nematodes. Genetics. 527 https://doi.org/10.1534/genetics.110.120550 528 Yan, C., Bi, Y., Yin, D., & Zhao, Z. (2012). A Method for Rapid and Simultaneous 529 Mapping of Genetic Loci and Introgression Sizes in Nematode Species. PLoS ONE, 530 7(8). https://doi.org/10.1371/journal.pone.0043770 531 Yin, D., Schwarz, E. M., Thomas, C. G., Felde, R. L., Korf, I. F., Cutter, A. D., … Haag, 532 E. S. (2018). Rapid genome shrinkage in a self-fertile nematode reveals sperm 533 competition proteins. Science, 359(6371), 55–61. 534 https://doi.org/10.1126/science.aao0827 535 Zhao, Z., Boyle, T. J., Bao, Z., Murray, J. I., Mericle, B., & Waterston, R. H. (2008). 536 Comparative analysis of embryonic cell lineage between Caenorhabditis briggsae 537 and Caenorhabditis elegans. Developmental Biology, 314(1), 93–99. 538 https://doi.org/10.1016/j.ydbio.2007.11.015 539 Zhao, Z., Flibotte, S., Murray, J. I., Blick, D., Boyle, T. J., Gupta, B., … Waterston, R. H. 540 (2010). New tools for investigating the comparative biology of Caenorhabditis 541 briggsae and C. elegans. Genetics, 184(3), 853–863. 542 https://doi.org/10.1534/genetics.109.110270 543 Zhao, Z., Sheps, J. A., Ling, V., Fang, L. L., & Baillie, D. L. (2004). Expression analysis 544 of ABC transporters reveals differential functions of tandemly duplicated genes in 545 Caenorhabditis elegans. Journal of Molecular Biology. 546 https://doi.org/10.1016/j.jmb.2004.09.052 547

548

21 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

549

550

551 Figure legend

552 Fig.1 Schematics of optimized protocol for miniMos-based transgene insertion in C.

553 briggsae. Injected animal was individually placed on a 55 mm crossing plate seeded with

554 OP50 (~6.7 ml NGM per plate) and allowed to recover at 25 °C overnight (roughly 12 to

555 16 hrs). For heat shock treatment, plates with injected animals were sealed with parafilm

556 and incubated on the water bath pre-heated to 35 °C for 3 hrs. Remove parafilm and any

557 water droplet on the inner lid. Incubate plates at room temperature (approximately 22 °C)

558 for half an hour. Add 0.5 ml 12.5 mg/ml neomycin (G418) solution pre-warmed up at room

559 temperature onto each plate. Gently rotate plates to allow neomycin solution to fully cover

560 plate surface. Air-dry the plates and incubate the dried plates in a 25 °C incubator for 6-7

561 days. Due to the presence of selection marker NeoR in the injected miniMos plasmids

562 (pZZ160 and pZZ184), single 10 to 12 survived L4 or young adult animals from the plates

563 crowded with animals that do not express co-injection marker onto individual plates to be

564 incubated at 25 °C. After 2-3 days, harvest animals with homozygous transgene based on

565 the reporter expression and lack of expression of visible co-injection marker for genotyping.

566

567 Fig. 2 Comparison of insertion frequencies in C. briggsae between this (brown) and

568 previous study (grey) (Frøkjær-Jensen et al., 2014). Y-axis denotes insertion frequency in

569 P0 animals. X-axis indicates various promoters used in the previous and this study that

570 drive Mos1 transposase expression. Each fluorescent marker in this study was tested for at

571 least three replicates. Animals with or without heat shock are indicated with hs(+) and hs(-),

22 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

572 respectively. Student’s t-test was performed for comparison between animals or without

573 heat shock. Error bars represent standard error.

574

575 Fig. 3 Genomic position of mapped insertions of visible marker Pmyo-2::gfp or Pmyo-

576 2::mCherry over C. briggsae genome (cb4). Insertions landed onto contigs unassigned to

577 a specific chromosome were not shown. Insertions of individual GFP (green) or mCherry

578 (purple) marker are indicated with inverted triangle along the genome in scale, detailed

579 insertion information is listed in Table S2. Insertion site was determined using two rounds

580 inverse PCR as described previously.

581

582 Fig. 4 Use of dual-colour fluorescent markers in mapping interacting loci responsible for

583 rescue of HI phenotype.

584 A. Schematics of introgression strains expressing GFP marker (green) generated previously

585 with biolistic bombardment or mCherry marker (red) generated with miniMos in this study.

586 Strains ZZY10330 and ZZY10377 carry an X-linked introgression derived from C.

587 briggsae that produces GFP or RFP expression, respectively, and HI phenotype in an

588 otherwise C. nigoni background, i.e., hybrid male sterility. Strains ZZY10353 and

589 ZZY10382 carry a Chromosome II-linked introgression derived from C. briggsae that

590 produces GFP or RFP expression, respectively, and HI phenotype in an otherwise C. nigoni

591 background, i.e., homozygous inviable (not shown).

592 B. Schematics of crossing strategy in mapping interacting loci between X chromosome and

593 Chromosome II. We previously demonstrated that presence of the introgression fragment

594 from ZZY10353 rescued the male sterility of ZZY10330. However, this demands tedious

23 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

595 genotyping of ZZY10353 to ensure its presence because it is impossible to distinguish the

596 two introgressions both expressing GFP. Substitution of ZZY10353 with ZZY10382

597 expressing RFP greatly facilitated the process for screening for simultaneous presence of

598 the two introgressions.

599 C. Reciprocal crossing between strains with autosome- and X-linked introgressions

600 expressing GFP and RFP, respectively, serves as the same purpose as in B.

601

602 Fig. 5 Comparison of killing efficiencies of PEEL-1 (A & B) and SUP-35 (C & D) in C.

603 elegans (N2) and C. briggsae (AF16). (A) Fraction of surviving array-bearing animals with

604 (grey) and without (black) heat shock in C. briggsae (AF16), which carries a peel-1 driven

605 by heat shock promoter as an extrachromosomal array as indicated. Note, no significant

606 change in survival using hsp-16.41 promoter of from C. elegans and CBG19187(cbr-hsp-

607 16.60) promoter from C. briggsae. (B) Fraction of arrested F1 animals presumably carrying

608 an injection marker derived from P0, which was injected with peel-1 driven by a

609 constitutively expressing promoter, Peft-3. Note the incomplete killing of PEEL-1

610 expressing animals in C. briggsae. (C) Fraction of arrested F1 animals presumably carrying

611 a co-injection marker derived from P0, which was injected with sup-35 driven by Peft-3.

612 Note the complete killing of sup-35 expressing animals in both C. elegans and C. briggsae.

613 (D) Fraction of surviving array-bearing animals with (grey) and without (black) heat shock

614 in C. elegans (N2) and C. briggsae (AF16), which carries a sup-35 driven by a heat shock

615 promoter from CBG19187(cbr-hsp-16.60) as an extrachromosomal array as indicated.

616 Note, no significant change in survival in C. elegans and a significant drop in C. briggsae

617 survival after heat shock.

24 bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

618

619 Supporting information

620 Figure legends for supplemental figures.

621 Fig. S1 Overview of plasmid components.

622 Fig. S2 Identification of equivalent of hsp-16.41 promoter in C. briggsae.

623 A. Schematics of Steps in identifying C. briggsae heat shock promoter. The hsp-16.41

624 promoter sequence was derived from pMA122, which is shared by two adjacent gene pair,

625 hsp-16.41 and hsp-16.2. Genomic sequences spanning the two genes were aligned against

626 C. briggsae genome “cb4” to identify their homology pairs. The cbr-HSP pair1 (grey)

627 shares the highest similarity in coding sequence to the C. elegans pair, in which CBG19185

628 is the best hit of hsp-16.41. However, cbr-HSP pair2 (brown) shows the highest similarity

629 in promoter sequence to the C. elegans promoter (data not shown).

630 B. Sequence of C. briggsae heat shock promoter tested in this study.

25

bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Microinjection

P0 Single animals onto NGM plate with food

Recover at 25°C for 12~16 hours

Heat shock Heat shock at 35°C for 3 hrs

Add neomycin Add 500 μl (12.5 mg/ml) G418 solution

F1 Wait for 6~7 days

F2 Single F2 from plates crowded with animals to individual plates

Check homozygosity based on F3 reporter expression

Isolate homozygous lines bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

40 hs(-) hs(+) 35 ** ** 30 n.s. Frøkjær-Jensen 25 et al., 2014 This study 20 15

GFP 10 n.s.

mCherry Insertion frequency(%) 5

0 GFP eft−3 eft−3 eft−3 eft−3 eft−3 −Ppie−1 cbr−P cbr cel−P cel−P cel−P cel−P bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

I mCherry GFP II

III

IV Chromosome V

X

0 5 Mb 10 Mb 15 Mb 20 Mb bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

A Chr. X introgression strains Chr. II introgression strains ZZY10330 ZZY10353

male sterile male fertile

female fertile female fertile

ZZY10377 ZZY10382

male sterile male fertile

female fertile female fertile

B

ZZY10382 ZZY10330

F1 C. nigoni

50% fertile C

ZZY10353 ZZY10377

F1 C. nigoni

32% fertile C. nigoni autosome C. briggsae introgression GFP C. nigoni X RFP bioRxiv preprint doi: https://doi.org/10.1101/704569; this version posted July 16, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

A B 100% Without heatshock 100% ** With heatshock 80% Phsp:: cbr-Phsp:: 80% Peft-3::PEEL-1 PEEL-1 PEEL-1 60% n.s. 60% n.s. 40% 40%

20% 20% Feaction of arrested animals Fraction of array-bearing animals 0% 0% AF16 AF16 N2 AF16 C n.s. D 100% 100% Peft-3::SUP-35 cbr-Phsp::SUP-35 n.s. 80% 80% ** 60% 60%

40% 40%

20% 20% Fraction of arrested animals Fractoin of array-bearing animals 0% 0% N2 AF16 N2 AF16