bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder.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 Distinct genetic architectures underlie divergent thorax, leg, and wing pigmentation

2 between elegans and D. gunungcola

3 Jonathan H. Massey1,2, Jun Li1,3, David L. Stern2, Patricia J. Wittkopp1,4*

4

5 1Department of Ecology and Evolutionary Biology, University of Michigan, Ann Arbor, MI

6 48109, USA

7 2Janelia Research Campus of the Howard Hughes Medical Institute, Ashburn, VA 20147, USA

8 3Institute of Evolution and Ecology, School of Life Sciences, Central China Normal University,

9 Wuhan 430079, China

10 4Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann

11 Arbor, MI 48109, USA

12

13

14 *Corresponding author

15 [email protected]

16

17 Running head: Pigmentation divergence between D. elegans and D. gunungcola

18 Keywords: yellow, ebony, optomotor-blind

19 Word count: 4,553 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder.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.

20 Abstract

21 Understanding the genetic basis of species differences is a major goal in evolutionary biology.

22 Pigmentation divergence between Drosophila species often involves genetic changes in

23 pigmentation candidate genes that pattern the body and wings, but it remains unclear how these

24 changes affect pigmentation evolution in multiple body parts between the same diverging

25 species. Drosophila elegans and D. gunungcola show pigmentation differences in the thorax,

26 legs, and wings, with D. elegans exhibiting male-specific wing spots and D. gunungcola lacking

27 wing spots with intensely dark thoraces and legs. Here, we performed QTL mapping to identify

28 the genetic architecture of these differences. We find a large effect QTL on the X chromosome

29 for all three body parts. QTL on Muller Element E were found for thorax pigmentation in both

30 backcrosses but were only marginally significant in one backcross for the legs and wings.

31 Consistent with this observation, we isolated the effects of the Muller Element E QTL by

32 introgressing D. gunungcola alleles into a D. elegans genetic background and found that D.

33 gunungcola alleles linked near the pigmentation candidate gene ebony caused intense darkening

34 of the thorax, minimal darkening of legs, and minimal shrinking of wing spots. D. elegans ebony

35 mutants showed changes in pigmentation consistent with Ebony having different effects on

36 pigmentation in different tissues. Our results suggest that multiple genes have evolved

37 differential effects on pigmentation levels in different body regions.

38

39

40

41

42 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

43 Introduction

44 Pigmentation differences within and between species illustrate some of the most striking

45 examples of phenotypic evolution in nature. In , pigmentation intensity in the body, legs,

46 and wings varies widely, often distinguishing sexes, populations, and species (True, 2003;

47 Wittkopp et al., 2003). The genetic and developmental processes determining pigment patterning

48 are well understood, which has facilitated the use of pigmentation as a model system to

49 investigate the genetic basis of phenotypic evolution (Wittkopp et al., 2003; Kopp, 2009).

50 Multiple studies have revealed that the same genes have often evolved independently to cause

51 pigmentation variation; that genetic variation within these genes often explains the majority of

52 pigmentation differences; and that mutations affecting the expression of enzymes and

53 transcription factors cause pigmentation to evolve (Massey and Wittkopp, 2016). These results

54 support the hypothesis that the genetic architecture of phenotypic evolution is, at least to some

55 degree, predictable (Stern and Orgogozo, 2008).

56

57 The insect pigmentation synthesis pathway (Figure 1) therefore provides a roadmap for

58 predicting which genes likely contribute to pigmentation differences within and between insect

59 species. Differences in how dopamine is metabolized in this pathway ultimately lead to

60 combinations of black and brown melanins as well as yellow, tan, and colorless sclerotins that

61 diffuse into the developing cuticle in sex-, population-, and species-specific ways (reviewed in

62 True, 2003). The genes yellow, tan, and ebony in this pathway have repeatedly evolved to

63 contribute to these differences. These include both within and between species differences in

64 abdominal, thorax, and wing pigmentation (reviewed in Massey and Wittkopp, 2016). In each bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

65 case, mutations in cis-regulatory DNA underlie these phenotypic changes, emphasizing the

66 importance of gene regulatory evolution in generating pigmentation diversity.

67

68 Most studies identifying the genetic basis of divergent pigmentation have focused on the

69 divergence of one particular element of the pigment pattern within each species pair. It thus

70 remains unclear how often the same genes and/or mutations are responsible for divergent

71 pigmentation seen in multiple body parts. In some instances, closely related species show

72 striking pigmentation differences across most body regions. Pigmentation divergence between

73 Drosophila americana and D. novamexicana, for example, involves differences in thorax and

74 abdominal pigmentation intensity that are partly explained by evolution of the ebony gene (Lamb

75 et al., 2020). Divergence in pupal case coloration between D. americana and D. virilis, however,

76 is caused by evolution at the dopamine N-acetyltransferase (Dat) gene (Ahmed-Braimah and

77 Sweigart, 2015). In other instances, species differ not only in the pigmentation intensity across

78 multiple body parts but also in pigmentation patterning. Pigmentation evolution in redheaded

79 pine sawfly larvae, for example, involves changes in body coloration and spot patterning that is

80 partially explained by overlapping quantitative trait loci (QTL) (Linnen et al., 2018).

81

82 Species differences between D. elegans and D. gunungcola involve dramatic changes in

83 pigmentation intensity across multiple body parts, although body pigmentation is also

84 polymorphic in D. elegans (Bock and Wheeler, 1972; Hirai and Kimura, 1997). Populations of

85 D. elegans from Hong Kong and Indonesia have light brown thoraxes and legs, and males

86 possess dark black spots at the tips of their wings (Hirai and Kimura, 1997; Figure 2). In Japan

87 and Taiwan, populations have dark black thoraxes while still possessing male-specific wing bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

88 spots (Hirai and Kimura, 1997). In D. gunungcola, pigmentation is less variable than in D.

89 elegans (Sultana et al., 1999; Massey et al., 2020). Males and females have dark black thoraxes

90 like the D. elegans Japanese and Taiwanese populations but also show dark black-pigmented

91 legs, and males do not have wing spots (Yeh et al., 2006; Massey et al., 2020; Figure 2).

92 Although these two species diverged 2-2.8 million years ago (Prud'Homme et al., 2006), they

93 can produce fertile, female F1 hybrids in the laboratory (Yeh et al., 2006), allowing us to

94 investigate the genetic basis of pigmentation evolution across multiple body regions in the same

95 mapping population.

96

97 Here, we use quantitative trait locus (QTL) mapping in two backcross populations between the

98 light D. elegans morph from Hong Kong (HK) and D. gunungcola from Sukarami, Indonesia

99 (SK) to identify gene regions associated with thorax, leg, and wing pigmentation divergence. For

100 all three traits, we identified a major effect QTL on the X chromosome, where pigmentation

101 candidate genes omb and yellow are located. We also find evidence for QTL linked to Muller

102 Element E for all three traits, but these effects depend on genetic background. We attempted to

103 isolate these effects by crossing D. gunungcola alleles into a D. elegans genetic background and

104 by introgressing a ~1.5 Mb region containing the ebony locus, which is found on Muller Element

105 E. We find that homozygous D. gunungcola alleles linked to this introgression darkened thorax

106 pigmentation intensity to D. gunungcola levels, but only minimally affected leg and wing

107 pigmentation, consistent with the differences observed in QTL mapping data for the three traits.

108 These data suggest that during the evolution of pigmentation across multiple body parts, epistasis

109 influences the extent to which individual pigmentation loci affect pigmentation divergence in

110 each tissue. In the case of D. elegans and D. gunungcola, evolution of at least two loci is bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

111 necessary to cause species differences in leg and wing pigmentation, but evolution at a single

112 locus is sufficient for divergence in thorax pigmentation.

113

114

115 Materials and Methods

116 stocks

117 Drosophila elegans HK (Hong Kong) and D. gunungcola SK (Sukarami) species stocks were a

118 gift from John True (Stony Brook University). Stock maintenance and food recipes are described

119 in Massey et al. (2020). In brief, stocks were maintained at 23ºC on a 12 h light-dark cycle. At

120 the third instar larval stage of development, adults were transferred onto new food, and

121 Fisherbrand filter paper (cat# 09-790-2A) was added to the larval vials to facilitate pupariation.

122

123 Generating hybrid progeny

124 Male and female D. elegans HK will reproduce with male and female D. gunungcola SK in

125 the laboratory to produce fertile F1 hybrid female and sterile F1 hybrid male offspring (Yeh et al.,

126 2006). Creating these F1 hybrids in populations large enough for genetic analysis, however, is

127 difficult. Detailed methods are described in Massey et al. (2020). In brief, both D. elegans HK

128 and D. gunungcola SK species stocks were expanded to establish populations of more than

129 10,000 flies. Next, virgin males and females from each species were placed in heterospecific

130 crosses (D. elegans HK males with D. gunungcola SK females and vice versa) in groups of ten

131 males and females to generate fertile F1 hybrid female offspring. Dozens of these crosses were

132 set to create ~120 F1 hybrid female offspring. Since F1 hybrid males are sterile (Yeh et al., 2006;

133 Yeh and True, 2014), the F1 hybrid females were used to generate two backcross populations for bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

134 QTL mapping. Briefly, for the D. elegans HK backcross population, ~60 F1 hybrid females were

135 crossed in the same vial with ~60 D. elegans HK males and transferred onto new food every two

136 weeks for ~2.5 months, resulting in 724 recombinant individuals. For the D. gunungcola SK

137 backcross population, ~60 F1 hybrid females were crossed in the same vial with ~60 D.

138 gunungcola SK males and transferred onto new food every two weeks for ~2.5 months, resulting

139 in 241 recombinant individuals.

140

141 Pigmentation quantification

142 For thorax pigmentation QTL analysis, male recombinants from each backcross population were

143 organized into three thorax pigmentation classes. The lightest recombinants showed thorax

144 pigmentation intensities similar to D. elegans HK and were given a score of 0; recombinants with

145 intermediate thorax pigmentation intensities were given a score of 1; and recombinants with dark

146 thorax pigmentation intensities similar to D. gunungcola SK were given a score of 2 (Figure 3A).

147 To quantify the effects of the Muller Element E introgression region on thorax pigmentation

148 (Figure 4B,C), individuals were placed thorax-side up on Scotch double sided sticky tape on

149 glass microscope slides (Fisherbrand) (cat# 12-550-15) and imaged at the same exposure using a

150 Canon EOS Rebel T6 camera mounted to a Canon MP-E 65 mm macro lens equipped with a ring

151 light. The images were then imported into ImageJ software (version 1.50i) (Wayne Rasband,

152 National Institutes of Health, USA; http://rsbweb.nih.gov/ij/) and converted to 32 bit grayscale.

153 Using the “straight line segment” tool, a line was drawn between the anterior scutellar bristles

154 (Figure 4B white dashed box) to measure the mean grayscale value of the cuticle. Data shown in

155 Figure 4C were inverted so that higher values indicated darker pigmentation. Raw data for

156 Figure 3 and 4 are deposited on Dryad. bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

157

158 For leg pigmentation QTL analysis, methods were identical to the thorax procedures above

159 except recombinants were organized into light (0), intermediate (1), and dark (2) classes based

160 on the pigmentation intensity of the medial side of their right hindleg femur (Figure 3B). This

161 region of the leg was chosen because it contains few cuticular bristles that could obscure

162 pigmentation. To quantify the effects of the Muller Element E introgression region on leg

163 pigmentation (Figure 4B,E), the same methods as above were used except right medial hindleg

164 images were captured. In ImageJ, the “polygon selections” tool was used to draw a polygon

165 selection around the medial side of the right hindleg femur (Figure 4B red transparency

166 selection) to measure the mean grayscale value of the cuticle. Data shown in Figure 4E were

167 inverted so that higher values indicated darker pigmentation. Raw data for Figure 3 and 4 are

168 deposited on Dryad.

169

170 For wing spot pigmentation QTL analysis, methods are described in Massey et al., (2020).

171 Briefly, right wings from male recombinants from each backcross population were imaged, and

172 spots were quantified in ImageJ using the “polygon selections” tool to quantify wing spot size

173 relative to wing size (Figure 4B red transparency selection). The same procedure was used to

174 quantify the effects of the Muller Element E introgression region (Figure 4D). Raw data for

175 Figure 3 and 4 are deposited on Dryad.

176

177 Library preparation, sequencing, and genome assembly

178 Detailed methods for preparing and sequencing the genomic DNA (gDNA) libraries for the D.

179 elegans HK and D. gunungcola SK backcross populations and advanced introgression line were bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

180 described in Massey et al., (2020). Briefly, gDNA was extracted from male recombinants by

181 homogenizing individuals singly in each well of a 96-well plate (Corning, cat# 3879). Each

182 recombinant gDNA sample was then barcoded with unique adaptors, pooled into a single

183 multiplexed sequencing library, size selected, and sequenced in a single lane of Illumina HiSeq

184 by the Janelia Quantitative Genomics Team. Methods for assembling the D. elegans HK and D.

185 gunungcola SK genomes to facilitate marker generation were described in Massey et al. (2020).

186

187 Marker generation with Multiplexed Shotgun Genotyping

188 Chromosome ancestry “genotypes” for the backcross progeny and introgression line were

189 estimated with two Multiplexed Shotgun Genotyping (MSG) (Andolfatto et al., 2011) libraries,

190 following methods described in Cande et al., (2012). Briefly, reads generated from the Illumina

191 backcross sequencing library were mapped to the assembled D. elegans HK and D. gunungcola

192 SK parental genomes to estimate chromosome ancestry for each backcross individual. We

193 generated 3,425 and 3,121 markers for the D. elegans HK and D. gunungcola SK backcrosses,

194 respectively, for QTL analysis [markers, phenotypes, and procedures for QTL mapping are

195 deposited on Dryad (doi:10.5061/dryad.gb5mkkwm5)]. PDFs of chromosomal breakpoints for

196 each recombinant are available here:

197 https://deepblue.lib.umich.edu/data/concern/data_sets/j098zb17n?locale=en.

198

199 QTL analysis

200 QTL analysis was performed using R/qtl (Broman and Sen, 2009) in R for Mac version 3.3.3. (R

201 Core Team 2018). We imported ancestry data for both backcross populations into R/qtl using a

202 custom script (https://github.com/dstern/read_cross_msg). This script directly imports the bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

203 conditional probability estimates produced by the Hidden Markov Model (HMM) of MSG

204 (described in detail in Andolfatto et al., 2011). We performed genome scans with a single QTL

205 model using the “scanone” function of R/qtl and Haley-Knott regression (Haley and Knott, 1992)

206 for thorax, leg, and wing pigmentation. For QTL mapping using the D. elegans HK backcross

207 population, we excluded 18 and 20 individuals for thorax and leg pigmentation, respectively,

208 because fly samples were either too poor to image or sequencing reads were too shallow to map.

209 For the D. gunungcola SK backcross population, we excluded 12 and 9 individuals for thorax

210 and leg pigmentation, respectively, for the same reasons. For wing pigmentation QTL analysis,

211 we previously (Massey et al., 2020) reported on an ~400 kb fine-mapped region on the X

212 chromosome explaining the majority of variation for wing spot size. We also reported on QTL

213 underlying variation in wing spot size with spotless recombinants removed from the analysis

214 (Massey et al., 2020). Here, in Figure 3D, we show QTL underlying this variation to emphasize

215 the role of Muller Element E in spot size divergence. Significance of QTL peaks at α = 0.01 was

216 determined by performing 1000 permutations of the data.

217

218 Introgression of black body color alleles from D. gunungcola SK into D. elegans HK

219 To isolate individual QTL underlying body color differences between D. elegans HK and D.

220 gunungcola SK, F1 hybrid females and D. elegans HK backcross recombinants were generated

221 using the methods described above. Next, dark black/brown female recombinants were

222 repeatedly backcrossed with D. elegans HK males for four generations (BC3-BC6). Finally, we

223 generated a single BC6 dark black homozygous introgression line that we then genotyped using

224 MSG (Figure 4A,B)

225 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

226 Creating an ebony null allele in Drosophila elegans HK via CRISPR-Cas9 genome editing

227 Using methods described in Bassett et al. 2013, we in vitro transcribed (MEGAscript T7

228 Transcription Kit, Invitrogen) three single guide RNAs (sgRNAs) (Supplemental File S1) with

229 target sequences designed based on conserved sites between D. elegans HK and D. gunungcola

230 SK in exon 2 of the ebony gene. After transcription, sgRNAs were purified using an RNA Clean

231 and Concentrator 5 kit (Zymo Research), eluted with nuclease-free water, and quantified using a

232 Qubit RNA BR Assay Kit (Thermo Fisher Scientific). Next, mature (>2 weeks old) D. elegans

233 HK males and females were transferred to 60 mm embryo lay cages (GENESEE Part Number:

234 59-100) on top of 60 mm grape plates (3% agar + 25% grape juice + 0.3% sucrose) at high

235 densities (>300 flies) after brief CO2 anesthesia. After CO2 anesthesia, mature, mated D. elegans

236 HK females will often dispel an embryo from their abdomen where it sticks briefly to the anus.

237 Tapping the grape plate + embryo cage down on a hard surface 10 times causes the embryos to

238 stick to the grape agar. Flies were then transferred back into food vials and embryos were lined

239 up on glass cover slips taped to a glass microscope slide. For CRISPR/Cas9 injections into D.

240 elegans HK embryos, Cas9 protein (PNA Bio #CP01), phenol red, and all three sgRNAs were

241 mixed together at 400 ng/µl, 0.05%, and 100 ng/µl final injection concentrations, respectively.

242 All CRISPR injections were performed in-house, using previously described methods (Miller et

243 al, 2002). Finally, we screened for germline mutants based on body pigmentation and confirmed

244 loss of Ebony protein by western blot (Supplemental Figure S1). To the best of our knowledge,

245 these are the first gene editing experiments to succeed in D. elegans. We attempted the same

246 experiments in D. gunungcola SK, but failed to recover any mutants.

247

248 Western blotting bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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 Western blot methods were followed similar to Wittkopp et al. (2002). In brief, for each replicate

250 per genotype, four newly eclosed (within 60 min) male flies were homogenized in 100 µl of 125

251 mM Tris pH 6.8, 6% SDS and centrifuged for 15 min. The supernatant was then transferred to a

252 new protein low-bind Eppendorf tube with an equal amount of 2X Laemmli sample buffer [4%

253 SDS, 20% glycerol, 120 mM Tris-Cl (pH 6.8)], boiled for 10 min, and stored at -80ºC. Before

254 gel electrophoresis, samples were thawed at room temperature for 30 min, and 20 µl of each

255 sample was loaded into individual wells of an Invitrogen NuPAGE 4-12% Bis-Tris Gel. The gel

256 was run at 175 V for 60 min, washed, and transferred to an Invitrogen iBlot 2 PVDF Mini Stack

257 Kit. The mini stack was run on an iBlot 2 (ThermoFisher, Catalog Number: IB21001) to perform

258 western blotting transfer, and blocked using an Invitrogen Western Breeze anti-rabbit kit

259 (Catalog Number: WB7106). After two washes with diH2O, samples were incubated in 1:400

260 rabbit anti-Ebony (Wittkopp et al., 2002) overnight at 4ºC. Finally, samples were washed using

261 the Western Breeze wash solution, incubated in secondary antibody solution alk-phos. (Catalog

262 Number: WP20007) with conjugated anti-rabbit for 30 min, washed for 2 min with diH2O, and

263 prepared for chromogenic detection.

264

265 Statistics

266 Statistical tests were performed in R for Mac version 3.3.3 (R Core Team 2018). ANOVAs were

267 performed with post hoc Tukey HSD for pairwise comparisons adjusted for multiple

268 comparisons. See “QTL analysis” methods for statistical tests used during QTL mapping.

269

270

271 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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 Results

273

274 X-linked divergence contributed to pigmentation divergence of all three traits

275 We examined the genetic basis for divergence of thorax, leg, and wing pigmentation between D.

276 elegans HK and D. gunungcola SK (Figure 2A). First, to assess the contribution of the X

277 chromosome in pigmentation divergence, we compared F1 hybrid males from reciprocal crosses

278 between D. elegans HK and D. gunungcola SK, which inherited their X chromosome from either

279 their D. elegans HK (F1E) or D. gunungcola SK (F1G) mother, respectively. We observed

280 striking pigmentation differences in the thorax, legs, and wings between the hybrids (Figure 2B).

281 F1 hybrid males inheriting their X chromosome from D. elegans HK mothers possessed wing

282 spots (quantified in Massey et al., 2020) and had thorax and leg pigmentation similar to D.

283 elegans HK (Figure 2A,B). In contrast, F1 hybrids from D. gunungcola SK mothers lacked wing

284 spots and had thorax and leg pigmentation similar to D. gunungcola SK (Figure 2A,B). These

285 observations suggest that X-linked loci contributed to pigmentation divergence in multiple body

286 parts.

287

288 QTL on the X chromosome underlie pigmentation divergence

289 To assess the genome-wide distribution of genes contributing to pigmentation divergence for

290 each body part, we performed quantitative trait locus (QTL) mapping. We generated two

291 backcross populations by crossing F1 hybrid females to males from each parental species and

292 quantified pigmentation variation for thorax, leg, and wing spots (see Methods) (Figure 3). For

293 each trait in both backcrosses, we identified major effect QTL on the X chromosome (Figure 3;

294 Table 1). The X-linked QTL peaks for thorax and leg pigmentation appear broader than the wing bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

295 spot pigmentation QTL (Figure 3), which is also reflected in their support intervals (Table 1).

296 For the wing spot, this QTL was previously refined to a ~400 kb region containing the

297 pigmentation candidate gene optomotor-blind (omb) (Massey et al., 2020; Table 1), consistent

298 with the single, narrow peak for wing spot variation (Figure 3C). The broader QTL peaks for

299 variation in thorax and leg pigmentation may result from multiple X-linked genes affecting these

300 traits.

301

302 QTL on Muller Element E underlie pigmentation divergence

303 We detected major effect QTL for thorax and leg, but not wing, pigmentation on Muller Element

304 E (Figure 3A,B), with the strongest effect linked to the left end of the chromosome (Table 1). For

305 these traits, the QTL peaks cover the entire Muller element, which might indicate low

306 recombination rate or the presence of multiple genes on Muller Element E contributing to thorax

307 and leg pigmentation divergence. Nonetheless, this result indicates that at least one gene on

308 Muller Element E affects thorax and leg pigmentation divergence differently than wing spot

309 pigmentation. Prior work has also shown, however, that autosomal QTL on Muller Elements C

310 and E affect the size of wing spots even though the X chromosome is sufficient to control their

311 presence or absence (Massey et al, 2020; Figure 3D).

312

313 To investigate further whether autosomal variation might affect wing spot divergence, and

314 whether this variation might be hidden by a large epistatic effect of the X chromosome, we

315 repeated QTL analysis on a subset of backcross recombinants that displayed wing spots of any

316 size and excluded offspring that completely lacked spots (Figure 3D). This analysis revealed

317 small effect QTL linked to Muller Elements C and E that affect wing spot size (Figure 3D). Like bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

318 thorax and leg pigmentation QTL, the largest effect was linked to the left end of Muller Element

319 E.

320

321 ebony is a candidate gene for the Muller Element E QTL

322 The pigmentation gene ebony is located near the left end of Muller Element E, where we

323 observed linkage to thorax, leg, and wing spot size pigmentation differences in the backcross

324 populations, making it an excellent candidate for the locus harboring the genetic variant(s)

325 affecting these traits (Figure 3). To test this hypothesis, we attempted to create stable

326 introgression lines by repeatedly backcrossing D. gunungcola SK alleles into a D. elegans HK

327 genetic background. For each backcross, we selected for dark black female recombinants to cross

328 with virgin male D. elegans HK for six generations (see Methods). Our goal was to isolate

329 multiple, independent introgression lines to fine map QTL involved in pigmentation divergence;

330 however, the majority of introgressions failed to propagate. We were not able to maintain any

331 introgression lines containing D. gunungcola SK alleles linked to the X chromosome in a D.

332 elegans HK genetic background (Massey et al., 2020), suggesting that divergence on the X

333 chromosome also contributed to hybrid incompatibilities. We did, however, recover a single

334 introgression line with a dark black thorax that resembled D. gunungcola SK (Figure 4A,B).

335 Sequencing this line revealed that it was homozygous for an approximately 1.5 Mb region of the

336 left end of D. gunungcola SK Muller Element E loci that included the ebony gene (Figure 4A).

337 These data provide independent confirmation that variation in genes on the left end of Muller

338 Element E contribute to thorax pigmentation divergence and support the hypothesis that ebony

339 contributes to thorax pigmentation evolution.

340 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

341 We next quantified pigmentation variation of the Muller Element E introgression line to

342 determine whether genetic variation in this region is sufficient to explain the Muller Element E

343 QTL effects. We found that this introgression completely recapitulated the species difference in

344 thorax pigmentation, but it had only weak effects on leg pigmentation and wing spot size (Figure

345 4C-E). This result allowed us to partially separate the genetic effects on Muller Element E,

346 further implicating ebony as a candidate gene with a large effect on thorax pigmentation

347 divergence and a small effect on leg and wing spot pigmentation.

348

349 To determine whether ebony contributes to thorax pigmentation divergence, we attempted to

350 knock out the ebony gene in both species using CRISPR-Cas9 genome editing to perform a

351 reciprocal hemizygosisty test (Stern, 2014). Unfortunately, we were able to generate only a

352 single ebony knockout line in D. elegans HK (Figure 4F; Supplemental Figure S1). We found

353 that the D. elegans HK ebony line displayed a dark black thorax (Figure 4F) from loss of Ebony

354 activity, similar to ebony mutant effects in D. melanogaster (Wittkopp et al., 2002). This result

355 further supports the hypothesis that lower Ebony activity and/or expression in D. gunungcola SK

356 underlies its dark thorax color. Interestingly, loss of Ebony activity did not darken leg

357 pigmentation or expand the size of wing spots in D. elegans HK (Figure 4F). One explanation for

358 this observation is that ebony is not normally expressed in these tissues in D. elegans HK, so loss

359 of Ebony function is not reflected phenotypically in these regions. Alternatively, because

360 pigmentation reflects the balance of expression between ebony, tan, and yellow (Wittkopp et al.,

361 2002; Wittkopp et al., 2009), differences in expression of these other genes might explain why

362 we did not observe the accumulation of dark black melanins in these tissues of D. elegans HK

363 ebony mutants (Figure 1). This observation could explain why genetic variation linked near the bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

364 ebony locus seems to affect thorax pigmentation divergence more than leg and wing

365 pigmentation divergence.

366

367

368 Discussion

369 We found that pigmentation divergence in the thorax, legs, and wings between D. elegans HK

370 and D. gunungcola SK is primarily influenced by genes on the X chromosome and Muller

371 Element E. Despite the fact that QTL regions overlap for these three body regions, several

372 observations suggest that pigmentation divergence results from evolution of multiple genes that

373 act in different ways on the three body regions examined. First, the X-linked QTLs that influence

374 thorax and leg pigmentation are broader than the QTL for wing spots. Second, QTL on Muller

375 Element E have a strong effect on thorax and leg pigmentation but minimal effect on wing spots.

376 Third, a 1.5 Mbp introgression of Muller Element E had a strong effect on thorax pigmentation

377 but minimal effects on leg pigmentation and wing spots. Thus, there is evidence for independent

378 genetic variation influencing pigmentation in each of these three body regions.

379

380 Evidence from previous studies is similar to our results. In D. melanogaster, for example, genetic

381 variants in the tan gene affect pigmentation in multiple abdominal segments, but genetic variants

382 in other genes affect each segment independently (Dembeck et al., 2015). Similarly, a single,

383 large effect QTL changes bristle number across multiple body regions in D. melanogaster, but

384 QTL on different chromosomes have independent effects on bristle number in the thorax versus

385 the abdomen (Mackay, 1995). Genetic variation in individual genes, therefore, can have bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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 pleiotropic effects on phenotypic differences in multiple body parts, whereas genetic variation in

387 less pleiotropic genes can have minimal effects in specific body parts.

388

389 Multicellular organisms develop distinct body parts through the action of gene regulatory

390 networks (Davidson, 2010). Differences in which (and how) genes are used during phenotypic

391 evolution likely reflect the activity of these networks (Stern and Orgogozo, 2008). If a gene is

392 not expressed in a given tissue, for example, then genetic variation at this gene will not be

393 reflected phenotypically. In our study, QTL on the X chromosome influenced pigmentation

394 differences in all three body parts (Figure 3), but QTL linked to the left end of Muller Element E

395 mainly affected thorax pigmentation (Figure 4). We predict, therefore, that species differences in

396 the spatial expression pattern of genes with distinct functions in the insect sclerotization and

397 pigmentation synthesis pathway (Figure 1) underlie these results.

398

399 In Drosophila, pigmentation coloring and patterning is determined by the action of multiple

400 enzymes that are regulated by multiple transcription factors (reviewed in Massey and Wittkopp,

401 2016). The enzymes Yellow and Ebony both require dopamine to synthesize dark black or light

402 yellow pigments, respectively, and transcription factors determine where these enzymes are

403 expressed (Figure 1; Massey and Wittkopp, 2016). Mutations that disrupt Yellow activity cause

404 flies to develop light yellow pigments, because Ebony converts excess dopamine into lightly

405 colored sclerotins. Reciprocally, mutations that disrupt Ebony activity cause flies to develop dark

406 black pigments, because Yellow converts excess dopamine into dark black melanins (Figure 1).

407 The phenotypic effect of mutations in one enzyme, therefore, require the reciprocal activity of

408 the other (Wittkopp et al., 2002). When we introgressed D. gunungcola SK Muller Element E bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

409 alleles containing the ebony gene into a D. elegans HK genetic background, we observed

410 darkening of the thorax but not the leg or wing (Figure 4B). This likely resulted from the

411 combined effects of low D. gunungcola SK ebony expression with high D. elegans HK yellow

412 expression. We hypothesize, then, that the D. elegans HK yellow allele is not normally expressed

413 in the legs or in regions outside the wing spot. Instead, the D. gunungcola SK yellow allele likely

414 evolved an expanded expression pattern in the legs and a restricted expression pattern in the

415 wings (Prud’homme et al., 2006). Dissecting these mechanisms in detail will shed light on how

416 the evolution of gene regulatory networks shape phenotypic divergence and vice versa.

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

432 Acknowledgements

433 We thank members of the Wittkopp, Stern, and Rebeiz labs for helpful discussions. For fly

434 strains, we thank J. True (Stony Brook University). For guidance with CRISPR/Cas9 genome

435 editing and embryo injections, we thank Abigail Lamb. For helpful advice on creating F1

436 hybrids, we thank Shu-Dan Yeh (National Central University). Funding was provided by

437 University of Michigan, Department of Ecology and Evolutionary Biology, Peter Olaus

438 Okkelberg Research Award, National Institutes of Health (NIH) training grant T32GM007544,

439 and Howard Hughes Medical Institute Janelia Graduate Research Fellowship to JHM; NIH R01

440 GM089736 and 1R35GM118073 to PJW.

441

442 Competing Interests

443 The authors declare no competing interests.

444

445 Data Archiving

446 All supporting data can be accessed at University of Michigan Deep Blue

447 (https://deepblue.lib.umich.edu/data/concern/data_sets/j098zb17n? locale=en) and Dryad.

448

449

450

451

452

453

454 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

455 References

456 Ahmed-Braimah, Y. H., & Sweigart, A. L. (2015). A single gene causes an interspecific

457 difference in pigmentation in Drosophila. Genetics, 200(1), 331-342.

458

459 Andolfatto, P., Davison, D., Erezyilmaz, D., Hu, T. T., Mast, J., Sunayama-Morita, T., & Stern,

460 D. L. (2011). Multiplexed shotgun genotyping for rapid and efficient genetic mapping. Genome

461 research, 21(4), 610-617.

462

463 Bock, I. R., and Wheeler, M. R. (1972) The Drosophila melanogaster species group. University

464 of Texas Publication 7213, 1–102.

465

466 Broman, K. W., & Sen, S. (2009). A Guide to QTL Mapping with R/qtl (Vol. 46). New York:

467 Springer.

468

469 Davidson, E. H. (2010). The regulatory genome: gene regulatory networks in development and

470 evolution. Elsevier.

471

472 Dembeck, L. M., Huang, W., Magwire, M. M., Lawrence, F., Lyman, R. F., & Mackay, T. F.

473 (2015). Genetic architecture of abdominal pigmentation in Drosophila melanogaster. PLoS

474 Genet, 11(5), e1005163.

475 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

476 Geyer, P. K., & Corces, V. G. (1987). Separate regulatory elements are responsible for the

477 complex pattern of tissue-specific and developmental transcription of the yellow locus in

478 Drosophila melanogaster. Genes & Development, 1(9), 996-1004.

479

480 Haley, C. S., and S. A. Knott. 1992. A simple regression method for mapping quantitative trait

481 loci in line crosses using flanking markers. Heredity, 69:315–324.

482

483 Hirai, Y., and Kimura, M. T. (1997) Incipient reproductive isolation between two morphs of

484 Drosophila elegans (Diptera: ). Biol. J. Linn. Soc., 61, 501–513.

485

486 Jeong, S., Rebeiz, M., Andolfatto, P., Werner, T., True, J., & Carroll, S. B. (2008). The evolution

487 of gene regulation underlies a morphological difference between two Drosophila sister species.

488 Cell, 132(5), 783-793.

489

490 Kopp, A. (2009). Metamodels and phylogenetic replication: a systematic approach to the

491 evolution of developmental pathways. Evolution, 63(11), 2771-2789.

492

493 Lamb, A. M., Wang, Z., Simmer, P., Chung, H., Wittkopp, P. J. 2020. ebony affects

494 pigmentation divergence and cuticular hydrocarbons in Drosophila americana and D.

495 novamexicana. Front. Ecol. Evol. doi: 10.3389/fevo.2020.00184

496 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

497 Linnen, C. R., O’Quin, C. T., Shackleford, T., Sears, C. R., & Lindstedt, C. (2018). Genetic basis

498 of body color and spotting pattern in redheaded pine sawfly larvae (Neodiprion lecontei).

499 Genetics, 209(1), 291-305.

500

501 Liu, Y., Ramos-Womack, M., Han, C., Reilly, P., Brackett, K. L., Rogers, W., ... & Rebeiz, M.

502 (2019). Changes throughout a genetic network mask the contribution of Hox gene evolution.

503 Current Biology, 29(13), 2157-2166.

504

505 Mackay, T. F. (1995). The genetic basis of quantitative variation: numbers of sensory bristles of

506 Drosophila melanogaster as a model system. Trends in Genetics, 11(12), 464-470.

507

508 Massey, J. H., & Wittkopp, P. J. (2016). The genetic basis of pigmentation differences within

509 and between Drosophila species. In Current topics in developmental biology (Vol. 119, pp. 27-

510 61). Academic Press.

511

512 Massey, J. H., Rice, G. R., Firdaus, A., Chen, C. Y., Yeh, S. D., Stern, D. L., Wittkopp, P. J.

513 (2020). Co-evolving wing spots and mating displays are genetically separable traits in

514 Drosophila. Evolution, DOI: doi:10.1111/evo.13990.

515

516 Miller, D. F., Holtzman, S. L., & Kaufman, T. C. (2002). Customized microinjection glass

517 capillary needles for P-element transformations in Drosophila melanogaster. Biotechniques,

518 33(2), 366-375.

519 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

520 Prud'Homme, B., Gompel, N., Rokas, A., Kassner, V. A., Williams, T. M., Yeh, S. D., ... &

521 Carroll, S. B. (2006). Repeated morphological evolution through cis-regulatory changes in a

522 pleiotropic gene. Nature, 440(7087), 1050-1053.

523

524 R Core Team. 2018. R: a language and environment for statistical computing. Available via

525 http://www.r-project.org/

526

527 Rebeiz, M., Pool, J. E., Kassner, V. A., Aquadro, C. F., & Carroll, S. B. (2009). Stepwise

528 modification of a modular enhancer underlies adaptation in a Drosophila population. Science,

529 326(5960), 1663-1667.

530

531 Stern, D. L. (2014). Identification of loci that cause phenotypic variation in diverse species with

532 the reciprocal hemizygosity test. Trends in Genetics, 30(12), 547-554.

533

534 Stern, D. L., & Orgogozo, V. (2008). The loci of evolution: how predictable is genetic

535 evolution?. Evolution, 62(9), 2155-2177.

536

537 Sultana, F., Kimura, M. T., & Toda, M. J. (1999). Anthophilic Drosophila of the elegans species-

538 subgroup from Indonesia, with description of a new species (Diptera: Drosophilidae).

539 Entomological Science, 2, 121-126.

540

541 True, J. R. (2003). Insect melanism: the molecules matter. Trends in Ecology & Evolution,

542 18(12), 640-647. bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

543

544 Wittkopp, P. J., True, J. R., & Carroll, S. B. (2002). Reciprocal functions of the Drosophila

545 Yellow and Ebony proteins in the development and evolution of pigment patterns. Development,

546 129(8), 1849-1858.

547

548 Wittkopp, P. J., Carroll, S. B., & Kopp, A. (2003). Evolution in black and white: genetic control

549 of pigment patterns in Drosophila. Trends in Genetics, 19(9), 495-504.

550

551 Wittkopp, P. J., Stewart, E. E., Arnold, L. L., Neidert, A. H., Haerum, B. K., Thompson, E. M.,

552 ... & Shefner, L. (2009). Intraspecific polymorphism to interspecific divergence: genetics of

553 pigmentation in Drosophila. Science, 326(5952), 540-544.

554

555 Yeh, S. D., Liou, S. R., & True, J. R. (2006). Genetics of divergence in male wing pigmentation

556 and courtship behavior between Drosophila elegans and D. gunungcola. Heredity, 96(5), 383-

557 395.

558

559

560

561

562

563

564

565 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

566 Figure Legends

567 568 Figure 1 Insect sclerotization and pigmentation synthesis pathway

569 Pigmentation enzymes (shown in red) convert substrates (shown in black) into pigmentation

570 precursors that polymerize into darkly colored melanins or lightly colored sclerotins. Synthesis

571 of black melanin depends on Yellow function, whereas Ebony converts dopamine into lightly

572 colored sclerotins. Tan catalyzes the reverse reaction of Ebony, converting NBAD molecules

573 back into dopamine.

574

575 Figure 2 Differences in reciprocal F1 male hybrids suggest X-linked sequence divergence in

576 pigmentation evolution between Drosophila elegans HK and D. gunungcola SK

577 (A) D. elegans HK (Hong Kong) and D. gunungcola SK (Sukarami) show divergent thorax, leg,

578 and wing pigmentation. (B) F1 male hybrids inheriting their X chromosome from D. elegans HK

579 mothers (F1E, bottom left) possess wing spots and have thorax and leg pigmentation similar to D.

580 elegans HK. F1 male hybrids inheriting their X chromosome from D. gunungcola SK mothers

581 (F1G, bottom right) lack wing spots and have thorax and leg pigmentation similar to D.

582 gunungcola SK.

583

584

585 Figure 3 Quantitative trait locus (QTL) mapping of thorax, leg, and wing pigmentation

586 divergence

587 (A) Thorax pigmentation was organized into three classes (light, intermediate, and dark). QTL

588 map for thorax pigmentation is shown for D. elegans HK backcross (red) and D. gunungcola SK

589 backcross (blue). (B) Leg pigmentation was organized into three classes (light, intermediate, and bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

590 dark). QTL map for leg pigmentation is shown for D. elegans HK backcross (red) and D.

591 gunungcola SK backcross (blue). (C) Wing spot pigmentation was quantified relative to wing

592 size (see Methods). QTL map for wing spot pigmentation is shown for D. elegans HK backcross

593 (red) and D. gunungcola SK backcross (blue) [Reprinted from Massey et al. (2020), Copyright

594 (2020), with permission from John Wiley and Sons]. (D) Similar to (C), wing spot pigmentation

595 was quantified relative to wing size, however, recombinants completely lacking wing spots were

596 excluded (red “X”) from the QTL analysis. QTL map for wing spot size is shown for D. elegans

597 HK backcross (red) and D. gunungcola SK backcross (blue) [Reprinted from Massey et al.

598 (2020), Copyright (2020), with permission from John Wiley and Sons]. Transparent yellow bars

599 indicate the chromosomal location of pigmentation candidate genes. LOD (logarithm of the

600 odds) is indicated on the y-axis. The x-axis represents the physical map of Muller Elements X, B,

601 C, D, E, and F based on the D. elegans HK assembled genome (see Methods). D. elegans HK

602 and D. gunungcola SK have six chromosomes (Yeh et al. 2006; Yeh and True 2014) that

603 correspond to D. melanogaster chromosomes as follows: X = X, B = 2L, C = 2R, D = 3L, E =

604 3R, and F = 4. Individual SNP markers are indicated with black tick marks along the x-axis.

605 Horizontal red and blue lines mark P = 0.01 for the D. elegans HK and D. gunungcola SK

606 backcross, respectively.

607

608 Figure 4 D. gunungcola SK alleles linked to ebony on Muller Element E have varied effects

609 on thorax, leg, and wing pigmentation divergence

610 (A) Multiplexed Shotgun Genotyping (MSG) (Andolfatto et al., 2011) was used to estimate

611 genome-wide ancestry assignments for a single introgression line generated by repeated

612 backcrossing of dark D. gunungcola SK recombinant females with D. elegans HK males (see bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

613 Methods). The posterior probability that a region is homozygous for D. elegans HK (red) or D.

614 gunungcola SK (blue) ancestry is plotted along the y-axis. (B) Images highlighting pigmentation

615 and chromosome differences between species and the Muller Element E introgression line. The

616 white dashed box highlights the scutellar region of the thorax used to quantify thorax

617 pigmentation in (C) (see Methods). Dashed red transparencies highlight regions of the wing and

618 leg used to quantify wing spot and leg pigmentation in (D) and (E), respectively (see Methods).

619 The black arrowhead emphasizes the wing spot region that has shrunk in the introgression line.

620 (C) Quantification of thorax pigmentation differences between species and the Muller Element E

-10 621 introgression line (One-way ANOVA F3,35 = 33.9; P = 1.84 x 10 ; groups not sharing the same

622 letter are significantly different based on post-hoc Tukey HSD at P < 0.00001). (D)

623 Quantification of wing spot pigmentation differences between species and the Muller Element E

-16 624 introgression line (One-way ANOVA F2,88 = 78.6; P < 2.0 x 10 ; groups not sharing the same

625 letter are significantly different based on post-hoc Tukey HSD at P = 0.02) [Reprinted from

626 Massey et al. (2020), Copyright (2020), with permission from John Wiley and Sons]. (E)

627 Quantification of leg pigmentation differences between species and the Muller Element E

-16 628 introgression line. (One-way ANOVA F3,33 = 481; P < 2.0 x 10 ; groups not sharing the same

629 letter are significantly different based on post-hoc Tukey HSD at P < 5 x 10-7). Gray triangles

630 represent individual replicates for (C-E). (F) Images highlighting effects of knocking out the

631 pigmentation candidate gene ebony in D. elegans HK.

632

633 Supplementary Figure S1 Quantification of Ebony band intensity in western blot

634 (A) Western blot using anti-Ebony antibody to detect Ebony protein levels (black arrowhead) in

635 newly eclosed D. elegans HK, the Muller Element E introgression line, and the D. elegans HK bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

636 ebony knockout line. Each number denotes a unique biological replicate involving four

637 homogenized flies. A second band just below the ~94 kDa Ebony band appeared in both the

638 wild-type and ebony knockout extracts (as observed in Wittkopp et al., 2002). (B) Using this

639 secondary band as a loading control, we quantified Ebony protein levels for each genotype and

640 observed no statistically significant differences. The western blot image was imported into

641 ImageJ software (version 1.50i) (Wayne Rasband, National Institutes of Health, USA;

642 http://rsbweb.nih.gov/ij/) and converted to 32 bit grayscale. Using the “straight line segment”

643 tool, the mean grayscale value of each band was quantified and inverted so that higher values

644 indicated darker band intensity (Student’s t-test; t = -0.495; df = 2.75; P = 0.646; two-tailed).

645 Gray triangles represent individual replicates, ns = not significant.

646

647

648

649

650 651 bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

Dark pigment (brown) Dark pigment (black) Dopamine Dopamine melanin melanin PO Pale DDC Yellow Tyrosine DOPA Dopamine Ebony Tan N-acetyl dopamine N-β-alanyl dopamine (NADA) (NBAD) PO PO

NADA sclerotin NADA sclerotin Colorless pigment Light pigment (yellow-tan) bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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 D. elegans HK D. gunungcola SK

X B C D E F X B C D E F

B F1E hybrid F1G hybrid (♀ D. ele x ♂ D. gun) (♀ D. gun x ♂ D. ele)

X B C D E F X B C D E F bioRxiv preprint doi: https://doi.org/10.1101/2020.06.28.176735; this version posted June 29, 2020. 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.

omb, D. elegans backcross (N = 706) tan yellow ebony A 40 D. gunungcola backcross (N = 229) 30 20 LOD LOD 10

Thorax 0

XA B C D E F B 40 D. elegans backcross (N = 704) Chromosome 30 D. gunungcola backcross (N = 232) 20 LOD LOD Leg 10 0 C XA B C D E F 200 D. elegans backcross (N = 656) Chromosome 150 D. gunungcola backcross (N = 199) 100 LOD LOD 50

Wing Spot 0

AX B C D E F D 7 7 6 6 50 D. elegans backcross (N =5 362) 5 Chromosome4 4 LOD LOD 3 3 40 D. gunungcola backcross2 (N = 128) 2 1 1 30 0 0 A B CA DB EC FD E F LOD LOD 20 X Chromosome Chromosome 10

Wing Spot 0

XA B C D E F Chromosome Chromosome ebony A Prob. Homozygouspar1 D. elegans HK Prob. Homozygous D. gunungcola SK par2 0 X 0 B 0 C 0 D 0 E F0 6.25 Mb A B C D E F

B D. elegans HK D. gunungcola SK Introgression D. elegans HK Introgression

X B C D E F X B C D E F X B C D E F X B C D E F

C 200 b b F D. ele HKebonyKO 180 a a 160

(mean grayscale) (mean 140

Thorax pigmentation pigmentation Thorax HK SK F6 F1_het D 0.16 a c 0.14 ) 2 0.12 b 0.10 (mm 0.08 Wing spot X B C D E F

Wing spot size size spot Wing absent * 0.06 * E HK SK F6_dark HET 200 b 180 c 160 a 140 a 120 (mean grayscale) (mean Leg pigmentation HKHK SKSK Intro.F6 F1_hetHK Intro. Table 1 QTLs detected for thorax, leg, and wing pigmentation divergence

Trait Backcross Chr. QTL interval QTL peak (bp) LOD Candidate (bp)a (bp) Thorax D. elegans HK X 8,996,888- 9,309,667 34.2 omb: pigmentation 10,120,567 10,700,000 yellow: 11,412,900

Thorax D. elegans HK E 244,349- 409,133 27.8 ebony: pigmentation 6,527,255 1,580,200

Leg D. elegans HK X 8,727,567- 9,456,223 34.0 omb: pigmentation 14,805,348 10,700,000 yellow: 11,412,900

Leg D. elegans HK C 14,797- 27,017 5.27 ? pigmentation 7,500,000

Leg D. elegans HK E 7,600- 606,610 27.9 ebony: pigmentation 6,588,062 1,580,200

Wing spot D. elegans HK X 10,297,836- 10,304,581 220 omb: 10,744,020b 10,700,000

Wing spotc D. elegans HK X 10,117,675- 10,303,766 49.1 omb: 10,748,235b 10,700,000

Thorax D. gunungcola SK X 6,969,794- 7,704,294 20.0 omb: pigmentation 9,711,406 10,700,000 yellow: 11,412,900

Thorax D. gunungcola SK E 2,083,292- 3,766,760 30.3 ebony: pigmentation 3,843,775 1,580,200

Leg D. gunungcola SK X 10,029,176- 10,117,133 39.5 omb: pigmentation 11,595,407 10,700,000 yellow: 11,412,900

Leg D. gunungcola SK E 517,241- 11,251,902 6.41 ebony: pigmentation 27,451,006 1,580,200

Wing spot D. gunungcola SK X 10,474,499- 11,223,359 38.9 omb: 11,584,862b 10,700,000

Wing spotc D. gunungcola SK C 6,655,757 - 8,420,192 4.37 Dll: 12,279,025b 5,221,911

Wing spotc D. gunungcola SK E 10,907- 12,292 6.85 ebony: 4,009,870b 1,580,200

a LOD drop 1.5 support interval (the region where the LOD score is within 1.5 of its maximum; Broman et al., 2003).

b Previously reported in Massey, J. H. et al. (2020). Co-evolving wing spots and mating displays are genetically separable traits in Drosophila. Evolution doi: 10.1111/evo.13990.

c Variation in wing spot size with spotless recombinant wings excluded.

A HK HK HK Intro. Intro. Intro. eKO (1) (2) (3) (1) (2) (3)

B ns 1.00 0.95 0.90 0.85 0.80 No band

(mean grayscale) (mean 0.75 present

Ebony band intensity intensity band Ebony HK F6 E - HK HK e ele D. ele Introgression D.